Introduction

The world's available energy resources have been diminishing since the beginning of the 21st century and saving energy is one of the most effective ways to solve this energy shortage problem. It is estimated that buildings are responsible for more than 40% of primary energy use in most countries1, primarily for space heating and cooling for indoor comfort. Energy-efficient wall or window coatings are considered to be the first step for reducing heat transfer between the indoor and outside environments. Energy efficient commercial products have already been developed, such as the Low-E window that possesses a low emissivity to achieve high thermal reflectivity of infrared light. Although the Low-E window has been widely used in the building industry, one of its disadvantages is that it cannot change its optical properties in response to environmental temperature changes or different human needs2. Based on the metal-to-semiconductor phase transition (MST, Fig. 1), vanadium dioxide (VO2) smart windows can exhibit automatic solar/heat control in response to environmental temperatures without the use of any external switching device, in comparison with other chromogenic windows, such as electrochromic or gasochromic and it has recently attracted considerable attention. VO2 undergoes a reversible MST at ~68°C, that is accompanied by a structural phase transition from monoclinic (P21/c, M1 phase) to tetragonal (P42/mnm, R phase) symmetry with an abrupt change in optical properties from transmitting to highly reflecting particularly in the infrared spectral region3,4. Thermochromic coatings have recently been developed using physical or chemical vapour deposition (e.g., sputtering5 or CVD6), polymer-assisted deposition7,8,9,10 and flexible foils by soft chemical methods11,12 and some prototype products have also been produced by our research group11,13.

Figure 1
figure 1

The crystallographic structure of VO2(M1) and VO2(R).

After metal-to-semiconductor transition (MST), linear and equal-distant (2.880Å) V-V chains along CR axis in tetragonal R phase become distorted with alternative distances at 3.132 Å and 2.662 Å with unit cell doubled in monoclinic M phase. Such a structural transition was accompanied with a small volumetric change of 0.044%.

Furthermore, environmental pollution as a by-product of rapid industrial development is another serious problem in addition to the energy shortage. Air quality is worsening in large cities and consequently, for example, the windows/walls of buildings are becoming dirtier more easily and require more frequent, labour-intensive cleanings that require extra organic detergents which cause secondary contamination. It is known that photocatalysts are effective for the degradation of organic dirt compared to the use of detergents. Among the various types of photocatalysts, titania (TiO2) is the most suitable material because of its excellent photocatalytic degradation ability for organic pollutants and photo-induced hydrophilicity14.

From the above background discussion, it is hypothesised that a composite material with a combination of VO2 and TiO2 would present both energy-efficiency and environmental protection properties. This proposed composite is a thrilling concept and of great value for applications in functional films/coatings. In fact, a multilayer structure of TiO2/VO2/glass deposited using PVD15,16 (magnetron sputtering) and a VO2/TiO2(anatase) composite film deposited using APCVD6,17 have been proposed for smart window applications; however, the above deposition processes are either expensive or hardly demonstrate photocatalytic performance.

A series of studies by our group11,18 have demonstrated that a nanopowder-based solution coating process may likely be a superior alternative to the PVD or CVD methods because of some advantages including simple coating systems, flexibility for substrate selection, ease for large-scale production and low-cost. Note that such functional plastic foils are considerably more suitable for windows on vehicles and old buildings; the latter represents a very large market. Because of the polymorphism of VO2 and the metastable valence of vanadium19, controlling the synthesis of VO2 mono-dispersed particles with a pure crystalline phase and a uniform morphology has long been a challenge. The majority of the developed methods can only produce either B/A metastable phases or M1/R phases with non-uniform morphologies20. Fortunately, there have recently been exciting breakthroughs in the synthesis of VO2 mono-dispersed single phase nanopowders by the hydrothermal method12,18,21,22,23. These technologies make it possible to achieve flexible foils.

The high chemical stability, the modified colour of VO2 film and a narrow hysteresis are three key requirements for VO2 thermochromic coatings. However, VO2 is gradually oxidised into toxic V2O5 in wet ambient atmospheres24, which leads to a considerable deterioration of its thermochromic properties. Enhancing the chemical stability of VO2 against oxidation is urgently required. VO2 films with grain size below 600 nm usually hold a yellowish colour, which originates from visible absorption specially in the ~300–400 nm region9. Reducing the yellow colour or changing to more acceptable colours favours applications for VO2. Hysteresis problem is rarely considered in powder-based films at present, however in principle, 20~26°C is considered to be a favorable temperature range for human comfort, so hysteresis higher than ~6°C will be inappropriate for real applications. However, films prepared using powders in reported studies commonly exhibited a very wide hysteresis9,10. For smart coatings deposited by PVD, it has been reported that a TiO2 layer coated on a VO2 film enhanced the luminous transmittance through an anti-reflecting (AR) effect, which may also likely protect the VO2 film against oxidation15,16.

Similarly in the case of nanopowders, core/shell nanotechnology is an available and facile method to obtain multifunctional VO2/TiO2 powders. Some pioneering works using VO2 as a core for the formation of core-shell composites have been reported. Jin, P.25 and Gao, Y. F. et al11 prepared VO2@SiO2 core-shell nanoparticles and Gao successfully made them into flexible foils with excellent solar energy modulation abilities and highly enhanced chemical stability. Although TiO2 exhibits advantages over SiO2 because of its photocatalytic properties, there are no reports on the production of high performance VO2@TiO2 composite nanopowders with both thermochromic and photocatalytic properties and a high chemical stability.

In this paper, we report for the first time synthesis of novel VO2@TiO2 core/shell nanoparticles using the consecutive hydrothermal synthesis of VO2, the sol-gel to prepare the titania precursor coating and a strictly controlled vacuum-annealing for crystallising the titania shell (Fig. 2). The VO2@TiO2 core/shell nanoparticles exhibit multiple functions: thermochromism from the VO2 core for solar energy modulation, photocatalytic properties from the TiO2 shell for environmental purification and highly enhanced chemical stability of VO2. Furthermore, it was observed that a TiO2 layer with a proper thickness can increase the visible transmittance of the coating using VO2@TiO2 nanoparticles and modify the yellow colour to a more pleasant blue one. Meanwhile, the coated VO2 showed a narrower hysteresis. The above technique breakthroughs will play a role in obtaining materials with integrated energy-savings and environmental protection functions for building applications.

Figure 2
figure 2

Illustration of coating and annealing procedures.

By sol-gel reaction at 80°C with reflux condenser for 2 h, TBOT was hydrolysed and precipitated onto the surface of VO2 nanorods and then collected by filtration. Then dried powders were annealed in vacuum quartz tube under proper temperature and pressure with the shell transformed into crystallised TiO2.

Results

Formation of core/shell nanostructure

Pure, single crystalline VO2(M1) powders were obtained after hydrothermal treatment. Fig. 3 presents the SEM, TEM images and the XRD pattern of the as-prepared VO2 nanorods. The SAED pattern in Fig. 3-b demonstrated that the as-prepared VO2 has a typical preferential growth direction along [100] (//aM1 axis), which is in accordance with previous reports12,19. The as-synthesised VO2 nanorods have a uniform morphology with an average short diameter of 500–600 nm.

Figure 3
figure 3

As-synthesized VO2 nanorods.

(a), (b) and (c) showed the SEM, TEM images and XRD pattern for as-synthesized VO2 nanorods, respectively. TEM showed a growing direction along aM1<100>.

The sol-gel process was used to produce TiO2 precursor shells with controllable thicknesses (Fig. 2). Detailed experimental parameters for some samples are listed in Table 1. The Ti(OR)4-H2O-alcohol (R = -CH3,-C2H5,-C3H7,-C4H9) system has been used for titania coating in the past decade on both particles26,27 and thin films28,29. Ti(OR)4 is hydrolysed and then peptized into titania colloids by condensation. The general reactions for these two processes are schematically described below:

Table 1 Experimental parameters for different samples (all at 80°C and react for 2 hours)

In this study, a carefully-controlled moderate hydrolysing process was achieved by utilising a relatively low concentration of reagents (Ti(OR)4 and H2O). The slow and dropwise addition of the H2O/ethanol mixture under vigorous stirring is important for nucleating and precipitating hydrolysed species on the VO2 surface to ensure a homogenous gel coating (Results of comparative experiments with increased concentrations and rapid addition (N6 and N7) are shown in supplementary Fig. S1 and the Tyndall phenomenon was used to determine whether isolated titania species are formed). Previous sol-gel synthesis studies have revealed that the TBOT-H2O-ethanol system with an excessive water content, typically with a TBOT:H2O molar ratio at 1:430 and 1:22031, readily resulted in homogeneous precipitation. In our process, homogeneous precipitation was largely diminished (the filtrate solution did not exhibit the Tyndall effect) by adopting the above mentioned lower concentration of reagents and the slow addition method to avoid a large degree of supersaturation. Therefore, titania species will preferentially precipitate onto the surface of the core VO2 rods to form shell coatings.

To obtain a crystalline coating, the annealing procedure was carefully investigated because of the ease of oxidation of VO2 in air or oxygen-rich atmospheres at relatively high temperatures (~>350°C)32. The results from TG-DTA measurements (see supplementary Fig. S2) indicate that samples N1 and N2 initially experienced a considerable mass loss until 300~350°C that resulted from the oxidation of organic -OR (R = (CH2)3CH3), which was followed by an abrupt weight gain caused by the oxidation of VO2. Previous studies have demonstrated that the transition from the amorphous titania gel precursor to anatase commonly begins at approximately ~350°C and a higher temperature, such as ~400°C29,31,33, favours an increase in crystallinity. Oxygen is essential for the removal of organic species (-OR) in amorphous titania shells, but an excessive amount of O2 leads to the oxidation of VO2. The two critical synthesis parameters, annealing temperature (T) and air pressure (P), should be carefully controlled to crystallise titania without oxidising VO2.

To determine the appropriate annealing parameters, the annealing performance of VO2 powders was investigated (see supplementary Fig. S3). At temperatures greater than 500°C, pure VO2 can be oxidised into high-valence vanadium oxides (e.g., V3O7 and V2O5) at a rather high pressure (~30 torr, an oxygen-rich atmosphere) or change into amorphous or partially reduced states under oxygen-deficient conditions (~48 mtorr). Similar results have been reported in the study of the synthesis of a VO2 film by reduction from V2O534, in which a temperature and pressure-dependent dynamic equilibrium between V2O5 and its reduced derivatives was revealed. This redox behaviour is a result of the multi-valence nature of the V-O system and a rather narrow solubility range of oxygen in the VO2 lattice35. This result implies that an appropriate oxidative atmosphere (with a certain proportion of oxygen at an air pressure of 48 mtorr~3 torr) with a proper temperature (400°C ~450°C) are preferable for obtaining the crystallisation of titania while keeping the VO2 core intact.

Based on the above discussion, the four composite samples were annealed (supplementary Fig. S3) and the optimised temperature and air pressure were ~400°C and ~3 torr, respectively. All samples under discussion in the following sections of this paper were treated using this same protocol. Fig. 4 presents the XRD patterns of samples N1–N4 before and after annealing. In the VO2@TiOx (amorphous gel) samples before annealing (Fig. 4-a), only peaks belonging to VO2 (JCPDF card No. 82-0661) were detected, which indicates that the gel coatings in all samples are amorphous or poorly crystallised. After annealing, the amorphous gel coatings were crystallised into anatase (JCPDF card No. 21-1272) without changes of the VO2 crystalline phase. However, the peaks that correspond to titanium oxides were not clearly detected for sample N1 because of a limited thickness, which will be discussed in the following.

Figure 4
figure 4

XRD patterns for pure VO2 and composite samples (N1 to N4) before (a) and after (b) annealing.

Before annealing, only VO2 can be detected, indicating the shell was amorphous or poorly crystallized. After annealing, peaks ascribing to anatase TiO2 can be distinguished from pre-existing VO2 peaks.

Fig. 5 (a) to (d) present the SEM images of the annealed core-shell particles N1–N4 with different thicknesses. Thick coatings (100~200 nm for N3 and N4) contain a considerable amount of apparent cracks that result from the relaxation of thermal stress, which can hardly be observed in the thinner coatings (N1 and N2). By increasing the loading amount of VO2 (m/g), the mean coating thickness (d/nm) become stably reduced from ~200 nm (0.03 g VO2, sample N4) to ~20 nm (0.2 g VO2, sample N1) with a nearly inverse proportional d-m relationship (refer to supplementary Fig. S4); this relationship is made possible because the precipitation rate was maintained constant by decreasing the homogeneous precipitation. Fig. 6 presents the TEM and EDS results for the N1 (VO2/TiO2 (~20 nm)) and N2 (VO2/TiO2 (~70 nm)) core/shell particles, respectively. A clear image contrast between the amorphous shell and the crystallised VO2 core provides estimated coating thicknesses of ~20 nm (Fig. 6-a) and ~70 nm (Fig. 6-c) for samples N1 and N2, respectively. The coatings cover the surface of the rods with a uniform thickness. After annealing, the core/shell interface become obscure (Fig. 6-b and d), which indicates that shell coating has crystallised.

Figure 5
figure 5

Morphology of the as-prepared VO2@TiO2 core/shell nanoparticles.

SEM images of the annealed samples N1(a), N2(b), N3(c) and N4(d) showed clear core/shell structure and thickness-dependent shell morphology. White rings in c and d indicate that the rather thick shell tend to form cracks.

Figure 6
figure 6

Microstructure probing of core/shell composite particles by TEM.

TEM images for sample N1 and N2. (a) and (b) are images for sample N1 before and after annealing. P1) is SAED pattern for the dot denoted in b, the HRTEM image for dot R1 were shown in Fig. 6R1. (c) and (d) are images for sample N2 before and after annealing. P2 and P3 are the SAED pattern for corresponding dots shown in Fig. 6d. R2,R3 are HRTEM images for two areas in Fig. 6d.

A more detailed microstructural analysis further confirmed that all the coatings crystallised into anatase TiO2. SAED patterns can distinguish two different materials with good resolving ability. For pure VO2 single rods, only one set of two-dimensional diffraction rings appeared (Fig. 6-b), whereas another set of SAED patterns with dispersed diffraction rings (red dashed circle) can be well indexed to anatase for core/shell samples N1 (Fig. 6-P1) and N2 (Fig. 6-P2), which is in agreement with the XRD results. The Raman spectra (Fig. 7) of N1 and N2 also indicate that the anatase shell was crystallised; a Raman shift at 143 cm−1 can be attributed to the dominating Eg(1) vibrational mode in anatase TiO236. Other peaks denoted by black figures can be attributed to the vibration modes in VO2(M1)37. The results from the EDS, Raman and SAED analyses confirm that the coating is primarily composed of anatase. The absence of anatase peaks in the XRD pattern of N1 should primarily be attributed to a relatively small crystal size (Fig. 6-R1, 5~10 nm) and volume fraction.

Figure 7
figure 7

Compositonal and phase identification by Raman spectroscopy.

Raman spectra of composite samples N1 (20 nm TiO2 coating) and N2 (70 nm TiO2 coating). Peaks denoted by red figures can be indexed to anatase TiO2, other peaks denoted by black figures can be ascribed to VO2(M1).

For N1 with a very thin TiO2 coating, the surface anatase particles have rather distinguishing morphologies. Some dispersed bright areas appeared on the relatively rough surface of N1 (Fig. 6-b), which should be a result of inhomogeneous expansion and shrinking of the coating during annealing and cooling. High resolution TEM images of the thin edge parts of the coating N1 and N2 are shown in images Fig. 6-R1, R2 and R3. N1 presents an anatase crystal size of ~5 nm (Fig. 6-R1), whereas N2 exhibits a single crystal size of >40 nm (R2) with an incident direction along the zone axis [-121] of anatase, which is confirmed by a set of periodic bright two-dimensional lattice in the SAED image; the interplanar distance exhibited by the lattice fringes is in accordance with that of the (101) plane of anatase. For point R3, the crystals are isolated with a rather small (~a few nanometres) size. Furthermore, the interface between adjacent anatase particles in N1 is very compact (Fig. 6-R1) compared to N2 (Fig. 6-R3).

Phase transition

DSC is a very commonly used technique for characterising the first-order phase transition12,19,22,23,38 of VO2 by monitoring the changes in enthalpy. In our work, the transition temperature (Tc, extreme values in heating and cooling cycles) and hysteresis (ΔT, the difference of the Tc in the heating and cooling cycles) for pure VO2 nanorods are very close to previously reported values (~53.6°C in a cooling cycle and ~64.6°C in a heating cycle, Fig. 8-a)12,16,17,35,36. After annealing (above dashed curves), sample N1 and N2 exhibited an increased Tc in both the heating and cooling cycles, which varied for different samples: 71.6°C (heating)/60.8°C (cooling) for N1 and 70.3°C (heating)/ 62.4°C (cooling) for N2, representing a Tc increase of 6~7°C compared to pure VO2. In addition, the width of the hysteresis loop (ΔT) of N2 was reduced after annealing from 11°C to 7.9°C, which is not observed for pure VO2 and is inconspicuous in sample N1. Thirdly, the enthalpy change (for unit mass) during the transition was reduced after coating (Fig. 8-b). This result is expected because TiO2 does not exhibit a phase transition in this temperature range. Through integration over the peak range, the titania mass content can be quantitatively calculated using the following formula: W(TiO2) = {1−I(sample)/I(VO2)} × 100%, where I denotes the integral intensity over the peak range. Based on this calculation (Fig. 8-c), the TiO2 shell accounts for ~29.3 wt% and ~51.0 wt% in samples N1 and N2, respectively.

Figure 8
figure 8

Phase transition property of the as-prepared VO2@TiO2 nanocomposites.

(a) DSC curves for: pure VO2(black curve), sample N1(red curve) and N2(blue curve) before (below solid curves) and after (above dash curves) annealing. (b) The magnified part of DSC peaks of annealed VO2, N1 and N2 during heating. Fig. 8c listed the value of integrated area for those three peaks in Fig. 8b and the mass ratio of TiO2 calculated based on the integral area (refers to the details in the corresponding text).

The transition between the semiconducting M1 phase and the metallic R phase generally has a ferroelastic feature39,40 that primarily includes two processes: the nucleation of a new phase in the original phase matrix and the boundary diffusion of the hetero-phase (M1&R). The nucleation is thermodynamically activated, which primarily depends on the microstructure (structural or chemical defects, where nucleation always preferentially initiates). The increased Tc indicates that the chemical stability of the M1 phase has increased compared to that of the R phase. Notably, pure VO2 also exhibits an increase of Tc after annealing (dashed black curves in Fig. 8-a). This result indicates that vacuum annealing can induce internal changes within the VO2 lattices. Intrinsically, non-stoichiometry32 can result in this change. Excess of oxygen can act as acceptor21 and thus increases the stability of monoclinic phase in Al/Cr-doped VO2 powders22. Recently, Zhang, S32 observed that a deficiency of oxygen (<30 mtorr) stabilised the rutile phase to temperatures as low as 103 K, whereas oxygen-rich growth conditions stabilised the monoclinic phases directly by confocal Raman nano-probing. Oxygen vacancies are commonly observed in VO2 polycrystalline films prepared under relatively oxygen-deficient reducing atmospheres. These oxygen defects can exert a considerable influence on the transition temperature41 because the new phase preferentially nucleates at the defect sites. In our case, the annealing under a relatively oxidative atmosphere (~3 torr) most likely increased the Tc.

With the exception of non-stoichiometry, interface effects may inevitably play an important role. Firstly, compositional diffusion at the VO2-TiO2 interface was confirmed by compositional scanning TEM analysis of the interfacial distribution of Ti and V elements using an EDS detector, as shown in Fig. 9-a to d. The curves for samples N1 and N2 before (a, b) and after (c, d) annealing exhibited a saddle-like distribution for the titanium content (red curves), with the two saddle points (marked by black sticks) at the interface between core and shell where the thickness of the titania coating reached a maximum. The relative intensity of the Kα1 signals of V and Ti reflected the compositional change before and after annealing. In the gel-coated samples (Fig. 9-a and b), the V intensity with respect to Ti outside of saddle points is very low. In contrast, the V intensity substantially increased after annealing, which implies that some amount of V diffused into the titania lattice. The impact of Ti doping on the dynamic phase transition of VO2 has been confirmed by previous studies42,43,44. V1−xTixO2 films exhibited lower hysteresis loop widths and increased Tc at higher Ti doping amounts (~9.4°C /at.%). A similar interfacial diffusion and Ti-doping can occur at 400°C45, 450°C46, or 500°C43 for TiO2/VO2 multilayer films. Ti-doping forms another metastable M2 phase and induces the formation of internal stress, which is reflected in an increase in the Mott phase transition temperature.

Figure 9
figure 9

Confirmation of the interfacial diffusion and Ti-V doping.

The compositional scanning TEM curves for: N1 (20 nm titania coating) and N2 (70 nm titania coating) before and after annealing. (a) N1, before annealing; (b) N2, before annealing; (c) N1, after annealing; and (d) N2, after annealing. Insets are dark-field TEM images of the corresponding samples. Scanning were conducted along the black line and signals of Ti and V Kα1 were collected for plotting.

In addition, the interfacial stress caused by the mismatch of thermal expansion/shrinkage between the core and shell materials may be another reason for the modified transition temperature. The average linear expansion coefficient (100~480°C) is 25 × 10−6 K−1 for VO2 along cR and 4.5 × 10−6 K−1 along aR47, whereas the value for titania is 7.38 × 10−6 K−1 (a) and 3.53 × 10−6 K−1 (a)48. This mismatch may result in an expansion (or shrinkage) prevailing along the rod length direction (because cR[001] corresponds to aM1[100]), which induces a compressive (when heating) and a tensile (when cooling) stress on the VO2 rod. For samples N1 and N2 with undamaged coatings, these thermal stresses may not be fully relaxed by microcracks. In our study, structural distortion was confirmed by slow XRD scans of the strongest VO2(M1) (011) peak (Fig. 10). There are apparent angular peak shifts in the annealed samples (dashed curves) compared to the unannealed ones (solid curves), which indicates that interfacial stress most likely induced the lattice distortion after annealing. The thermal stress changed the defect state and distribution on the local surface, which would have a significant influence on the phase transition behaviour of VO2. Recently, structural and electronic impacts were observed crossing phase boundaries at high-strain states49, whereas oxygen defects could be induced by boundary stress50, which jointly affect the phase transition behaviour.

Figure 10
figure 10

Confirmation of the structural change after annealing by XRD.

The XRD slow scanning curves from 27.3° to 28.3° (peak of (110)M1) for VO2(black), N1(red) and N2(blue) before (solid curves) and after (dash curves) annealing were shown and clear peak shifts indicated that crystallographic structure was changed after annealing.

From the above discussion, it appears that the annealing process not only changes the intrinsic stoichiometry of VO2 but also leads to interfacial diffusion and stress to some degree. Note that although the Tc was increased for our VO2@TiO2 particles, the transition temperature can be easily tuned by doping with W19 and/or Sb18. In fact, W-doped VO2@TiO2 particles were also synthesised with a W-doping content at ~1.34 at% and the transition temperature was reduced to 26.5 ~ 35.0°C with 3.5–6.0°C increase as compared to uncoated W-doped VO2 (see supplementary Figure S5). This temperature range is appropriate for real applications.

Thermochromic property

The assessments of the thermochromic property were only implemented on samples with uniform coatings (samples N1 and N2) to exclude the impact of macrocracks and other uncontrollable irregularities. In our study, the use of Teflon tape as a medium to form the powder coating has particular advantages. Fig. 11-a presents the SEM image of a typical surface of the film, in which the rods are patterned with only one single layer, which is useful for evaluating the properties of the rod with a uniform thickness while eliminating the influence of the dispersing media. Because the size of the nanorods is comparable to the wavelength of the incident radiation, UV-V is diffuse transmittance and reflectance (further denoted as Tr and Re) spectra including the scattering effect were measured and these results are shown in Fig. 11-b and c (FTIR transmittance and reflectance curves were refers to supplementary Fig. S6). The transmittance hysteresis loop at 2000 nm with various temperatures is shown in Fig. 11-d. Both N1 and N2 exhibited conspicuous infrared modulability and the transmittance modulation ability and luminous transmittance are summarised in Table 2.

Table 2 Summary of solar and luminous transmittance
Figure 11
figure 11

Thermochromic performance of as-prepared core/shell particles.

(a) is the SEM image of one typical film surface. Inset in (a) demonstrates films on glass made of pure VO2 (left, deep yellow-brown colour) and N2 (right, blue colour, with higher transparency). (b) and (c) are the diffuse transmittance and reflectance spectra in the wavelength range from 250–2500 nm. Samples are: VO2 (black curves), sample N1 (red curves) and N2 (blue curves) at 25°C (dash curves) and 90°C (solid curves). (d) is the temperature-varied transmittance of those three annealed samples at 2000 nm during heating (solid) and cooling (hollow) cycles.

Neither Teflon tape nor TiO2 have thermochromic properties; therefore, the change in the Tr and Re with temperature should be due to the phase transition of the VO2 particles. The entire spectral range in question can be divided into two regions: at a shorter wavelength region A the transmittance at 90°C is higher than 25°C; and at a longer wavelength region B the transmittance at 25°C is much higher than that at 90°C (the so-called IR switching) In the region B, the Tr behaviour is very conventional for VO2, in which the Tr is lower at temperatures above the Tc (metallic R phase) than that at lower temperatures (semiconducting M1 phase). The plasma frequency (ωp) for the R phase in polycrystalline thin films is reported to be 1.0 eV (~1.24 μm)51 or 1.6 eV (~775 nm)52, which is proportional to the free-electron concentration and depends on the temperature and the film/substrate interface state. Hence, the metallic R phase has a lower transmittance and higher reflectance in the IR region (ω < ωp). However, semiconducting VO2(M1) has a band gap at approximately 0.7 eV; therefore; it is highly transparent to photons with energies less than Eg (λ > ~1.77 μm). However, at the region A including the whole visible part, the R phase exhibited a higher transmittance than low-temperature M1 phase, although both transmittances were at a relatively low level (<30%) because of intrinsic absorption. This reverse transmittance cause undesirable minus modulation for solar energy as shown in Table 2. This behaviour deteriorate the whole efficiency and may be a considerable drawback for applications that require good visual senses; whereas, such a coating can be used for energy-saving wall paint/coatings or the rear windows of cars. As illustrated in Fig. 11-a, there are interspaces among rods, typically several tens of nanometres to submicrons. Therefore, the transmittance of our as-prepared samples in region A should primarily be attributed to light transmittance from the uncovered areas of the Teflon tape and the light scattered by the VO2 rods in the forward direction. Mie scattering predominates because the rod thickness (500–600 nm) was very similar to the wavelength of the incident visible light. According to Mie theory53, scattering rises to a maximum when it is the same as the characteristic length of the scattering centre (the rod thickness) and then it decays to zero towards longer wavelengths. Therefore, the Tr change before and after the Mott transition in this region should contribute to the different scattering abilities of the VO2 composite particles. The M1 phase scatters light more strongly due to its larger refractive index (n(VO2(M1)) = 2.7 ~ 2.8, n(VO2(R)) = 2.0 ~ 2.5)15,54, which can lead to smaller transmittance than that of the R phase. In particular, after being coated with highly-transparent TiO2, the composite films exhibited considerably improved luminous transparency. As shown in Fig. 11-a, the 70 nm TiO2 coating (N2) significantly increased the luminous transmittance, from ~8% to ~27% (Table 2). The transparent TiO2 coating accounts for a considerable part of the total film in thick N2 samples, which increases the total visible transmittance.

For a pure VO2 film, the modulation ability at 2 μm is ~70.6% with a large hysteresis loop width (more than 10°C). The TiO2-coated composites exhibited a modified optical switching property, both in the modulation level and hysteresis loop width (Fig. 11-b and d). A 70 nm TiO2 coating (sample N2) exhibits a largely increased transmittance in both the M1 and R states, but with somewhat deteriorated modulation ability (~58.6%). A thinner 20 nm coating (sample N1) can also increase the transmittance by a small degree with maintained modulability (~75.2%). Furthermore, the TiO2 coating can significantly decrease the transition hysteresis, which is in agreement with the DSC measurements. However, films prepared using powders in previous studies commonly exhibited a very wide hysteresis. We prepared films based on VO2 nanoparticles with large hysteresis (~21°C) and VO2@SiO2 composite particles with very wide transition range (~30°C). Lu, Z55 prepared a VO2/PEGDA film with hysteresis at ~25°C. Our composite particles with considerably reduced hysteresis will benefit future applications for rapidly responding and high-precision controlled materials. The diffusion of Ti into VO2 will weaken its IR switching property due to the introduction of lattice defects45. N2 with considerable Ti-V diffusion will clearly exhibit reduced modulability. However, the significantly enhanced transmittance and reduced hysteresis may compensate for this weakness and further investigations on optimising the VO2 size and coating thickness will be required to achieve better integrated performance.

As previously mentioned, VO2 films have an intrinsic yellow-brown colour and the pure VO2 film exhibits a dark colour, as shown in Fig. 11-a, inset (left). However, N2 with a 70 nm TiO2 coating exhibits a more transparent blue colour (Fig. 11-a, inset (right)). This result implies that the 70 nm titania coating modified the colour of the film. Calculations of the optical interference effect in the sandwiched 70 nmTiO2/500 nmVO2/70 nmTiO2 (nair < nTiO2 < nVO216) structure suggest that the 70 nm TiO2 coating can serve as an effective anti-reflecting coating for red light (λ ~ 700 nm, λ = 4 × n × d, n is the average refractive index of anatase, n ≈ 2.5 and d is the thickness of the TiO2 coating, d ≈ 70 nm). Furthermore, the incidental short wavelength light (~300–400 nm) is reduced by a large degree because of the existence of UV-absorbing TiO2. Therefore, the TiO2 coating changed the spatial distribution of the scattered visible light, by which we can observe the reflected blue colour. This coating effectively changed the appearance of the film. In summary, TiO2 coatings can effectively increase the luminous transmittance and modify the colour of films. The modified blue colour is considerably more favourable and acceptable as a functional and decorative coating.

For practical application, W-doped composites with lower transition temperature were also considered. The UV-Vis spectra were shown in Fig. 12. The transmittance curves and hysteresis loops showed a similar law as undoped VO2: after coating, the average transmittance was increased with deteriorated modulation ability; the loop width was also narrowed down by 5.7°C. Meanwhile, the coated sample showed a light blue colour in comparison with yellow one of uncoated VO2:W film. As compared with undoped films, the IR switching efficiency is diminished because of the incorporation of W dopant, as reported elsewhere. However, the core/shell particles show a modified colour of light blue, which is more comfortable for human sense.

Figure 12
figure 12

VO2:W@TiO2 nanocomposites with a solar modulation ability at room temperature.

By the sol-gel coating method, we successfully prepared VO2:W@TiO2 powders (a). The DSC curves are shown in Fig. 12b. After W-doping, the transition temperature has been reduced to 20~30°C and after coating the transition temperature was increased by 3.5°C (heating cycle) and 6°C (cooling cycle). The films of the VO2:W and VO2:W@TiO2 powders were prepared in the same method described in “Methods” section with the same solid loading content. The transmittance curves and hysteresis loops showed a similar law (Fig. 12c and d): after coating, the average transmittance was enhanced whereas with deteriorated modulation ability; the loop width was also narrowed down by 5.7°C with increased transition temperatures both in heating and cooling cycles. Meanwhile, the coated sample showed a modified light blue colour in comparison with the yellowish one of uncoated VO2:W film.

For energy-efficient smart window, the as-prepared particles are so large that scattering and absorption dominate the solar modulation, which results in a very small efficiency and may inappropriate for practical application by now. However, VO2 particles with much smaller size (short diameter < 50 nm) was synthesized according to the literature23. VO2@TiO2 nanoparticles with shell thickness 1~7 nm has been successfully prepared and showed an excellent solar efficiency of 10.27% (Fig. 13). However, as shown in Fig. 13-c and f, no colour modification or obvious interfacial effect on transition was observed by now. The effect of interfacial Ti-V doping and stress on the phase transition temperature were inconspicuous because of relatively thin coating.

Figure 13
figure 13

VO2@TiO2 nanoparticles with excellent solar modulation efficiency.

VO2@TiO2 nanoparticles with short diameter ~50 nm was prepared by ultrasonic-assisted hydrolysis. The shell thicknesses of sample P1 and P2 can be estimated from (a) and (b) to be 1 nm and 7 nm, respectively. The photograph of the films and the corresponding transmittance spectrum were shown as below in Fig. 13c and d, respectively, the solid and dash curves are corresponding to the spectra collected at 90°C and 25°C. However, no obvious colour change was found in this system. P1 and P2 with size ~<50 nm will exhibit relatively small scattering and absorption for visible light in comparison with the big composite particles N1 and N2. As shown in Fig. 13f, as high as 10.27% of solar modulation efficiency can be obtained by the film of P1 powders. However, P1 and P2 didn't exhibit much variation in the transition temperatures and hysteresis (Fig. 13e).

Photocatalytic properties

The photocatalytic properties of our samples were investigated by monitoring the decomposition of Rhodamine B (RhB) under simulated solar light produced by irradiation from a Xe-lamp (240 nm~2 μm). For comparison, 0.128 g of the N1 and 0.073 g of the N2 composite powders were used in each experiment to maintain a constant titania content (according to the titania mass ratio calculated from the previous DSC results). In addition, 0.1 g of VO2 was applied to determine whether VO2 itself has photocatalytic properties. Before light irradiation, the adsorption behaviours of RhB over the three samples were examined under dark conditions. The adsorption-desorption equilibriums could be established within 1 hour for all the samples. The amounts of RhB adsorbed on the surfaces of the samples are listed in Table 3. The TiO2 coated samples exhibited enhanced adsorption abilities. On the basis of the same initial RhB concentration, as high as 60%, 81% of RhB was adsorbed onto the catalyst surface over the N1 and N2 samples compared to only 22% over the uncoated VO2 sample. The enhanced adsorption abilities for RhB with respect to the pure VO2 can be attributed to the significantly increased surface areas of samples (Table 3) and particularly to the considerably stronger adsorption ability of porous titania.

Table 3 Adsorption ability and Specific surface area of pure VO2, N1 (VO2@20 nmTiO2) and N2 (VO2@70 nmTiO2) based on N2 adsorption measurements (Ame), or calculation with ideal closely-packed surface postulate (Acal)

The temporal UV-Vis spectral evolution of the RhB aqueous solution for these three samples during the photocatalytic degradation reactions are plotted in Fig. 14. Pure VO2 has negligible photocatalytic degradation ability for RhB (after excluding the impact of the natural degradation of RhB and adsorption under the same illumination conditions), whereas the coated samples N1 and N2 can both degrade the RhB dyes with a considerable efficiency at a given time (67.4%~69.6% for 2.5 h). The primary absorbance that maximised at ~554 nm markedly decreased with irradiation time. Comparison experiments demonstrated that the photolysis of RhB without a catalyst was very slow (brown curve) and that RhB could not be degraded in the presence of samples under dark conditions, which implies that the degradation of RhB in the present study is indeed through a photocatalytic process.

Figure 14
figure 14

Evaluation of photocatalytic property.

(a) The RhB concentration change with varied reaction time, for adding VO2 (black curve) and VO2/TiO2 composites N1 (red curve) and N2 (blue curve) samples, yellow curve is the natural degradation curve of RhB under the same illumination condition. (b), (c) and (d) are the UV-Vis spectra evolution of RhB aqueous solution in corresponding systems: pure VO2 (b), N1 (c), N2(d).

Furthermore, both of these composite samples have higher specific surface area values than the theoretical values obtained through a simple calculation based on an ideal close-packed coating (refers to supplementary Fig. S7), which indicates that the crystals of the sample surfaces are far from closely-packed. This result is in agreement with the microstructures observed previously, where the interface between adjacent anatase particles in N1 is relatively compact (Fig. 6-R1) compared to N2 (Fig. 6-R3). A certain amount of micro-/nanopores or cracks composed the majority of the total surface area, especially in sample N2. A larger specific area can ensure more active sites for adsorption and subsequent reactions, but desorption of the degraded species will be significantly hindered if they reside in the inner space, which will in turn reduce the entire degradation efficiency. This may be the reason why sample N2, which had a considerably larger specific surface area than sample N1, did not exhibit considerably enhanced degradation ability. In addition to the specific surface area, other factors such as surface defects, carrier concentrations and carrier transfer rate will also affect the final photocatalytic performance. Modifications of the microstructure can effectively adjust the photocatalytic performance. Compared to other core/TiO2 particles56,57, the photocatalytic efficiencies of the as-synthesised samples are not very satisfactory, which is most likely due to their uncompetitive surface areas. On-going research on VO2@TiO2 nanocomposite particles with smaller core size may provide more exciting performance, which will be reported in the near future.

To exclude the heating effect from the high power irradiation, we placed the sample into an ice water bath during the reaction. We note that VO2 remains chemically stable after 2 hours of light exposure. The XRD patterns of the irradiated N1 and N2 samples are shown in S8.

Enhanced chemical stability against oxidation

To explore whether the titania coating will improve the oxidation resistance of the as-prepared VO2@TiO2 composite particles, annealing in air at different temperatures for the same amount of time (15 min) was conducted with a heating rate of 10°C /min on three samples. The phase evolution after annealing is shown in Fig. 15-a. VO2 will readily oxidise into V3O7 or V2O5 at temperatures greater than 290°C. In contrast, the coated N1 and N2 samples remained chemically stable until 320°C, which suggests that the titania shell can act as an effective barrier layer for the diffusion of oxygen into the VO2 lattice.

Figure 15
figure 15

Enhanced oxidation-resistant ability after TiO2 coating.

(a) plots the XRD results of phase evolution for pure VO2 (black curve), N1 (red curve) and N2 (blue curve) after annealing at different temperatures in the air. (b) shows TG curves for pure VO2 (black curve), N1 (red curve) and N2 (blue curve) sample powders.

Thermogravimetric analysis (TGA, Netzsch STA449C, under air) provided a more subtle perspective for monitoring the complete oxidation process. Fig. 15-b presents the results from the TGA of three samples (pure VO2 (black curve), VO2@TiO2 (20 nm) (red curve) and VO2@TiO2 (70 nm) (blue curve)). The sample mass began to increase at certain temperature values (shown by arrows), which indicate the onset of oxidation. Compared to ~301°C for pure VO2, the characteristic temperature for the TiO2-coated N1 and N2 samples increased to ~336°C and 343°C, respectively. Furthermore, the anatase-coated samples exhibited no clear mass losses, which indicate that the titania species were almost crystallised and that residual organic species scarcely exist. The crystallised TiO2 shell can enhance the oxidative resistance of VO2, which is of significant importance for practical applications.

Discussion

Our results demonstrate a new approach to control the optical and Mott transition properties by modifying the surface/interface state and microstructure of VO2 particles with core/shell nanotechnology. TiO2 coating can not only enhance the chemical stability of VO2 but also exhibits the potential to control the transition temperature and hysteresis by designed "interfacial doping" with Ti through annealing. Such "interfacial doping" has the advantage of avoiding the introduction of a large amount of defects as compared with bulk doping. Bulk defect will deteriorate the modulation ability of VO2 to a large degree. We expect that other cationic or anionic ions for example, Mo, W, Cr, F etc, which can adjust the transition property by bulk doping may also be designed to achieve "interfacial doping" through core/shell nanotechnology, thus to modify the phase transition of VO2 without bringing in excessive defects and retain a comparably sound switching efficiency as in pure VO2.

By changing the thickness of TiO2 shell, the distribution of scattered and transmitted light can be tuned to satisfy an anti-reflecting effect on a micro single-particulate scale (300~600 nm). We can expect such a colour modification can be only realized in particles with size comparable to visible light where scattering plays an important role in the light-matter interactions. The colour of smaller VO2 particles (~<50 nm) cannot be effectively changed by TiO2 coating, as shown in Figure 13.

The photocatalytic property of anatase TiO2 coating can be affected by grain size, defection and boundary states. We expect a better photocatalytic property can be obtained by adjusting the annealing program. This integrated structure combining solar-light driven environmental purification with energy modulation can make better use of the solar energy. We forsee the potential of nanocomposite of VO2 with other functional materials, such as light harvesting/conversion materials, to get combined or mutually optimized multi-functions for solar energy conversion or modulation.

Methods

Preparation of VO2 nanorods

Single-crystal VO2 nanorods were prepared following a previous work by our group18. V2O5 (Sinopharm Chemical Reagent Co.Ltd, China) and oxalic acid powders (H2C2O4.2H2O, Wako, Japan) were directly added in a 1:2 molar ratio to 200 mL of deionised water ([V] = 0.037 M) at room temperature. The suspension was continuously stirred until a clear dark blue solution was formed. Then, 40 mL of the suspension was transferred into a 100 mL Teflon-lined autoclave after adjusting its pH to ~1.85 by adding a H2SO4 (0.5 M) aqueous solution. After the hydrothermal treatment at 260°C for 48 h, the autoclave was naturally cooled to room temperature. The resulting black-coloured precipitate was collected by filtration, washed alternatively 3 times with deionised water and alcohol and then dried at 60°C in air for 10 h.

Preparation of VO2@TiO2 core/shell nanorods

To prepare the VO2@TiO2 core/shell nanostructures, 0.03~0.25 g of the as-synthesised VO2 nanorod powders was ultrasonically dispersed in 80 mL of water-free ethanol (Sinopharm Chemical Reagent Co.Ltd, China) to obtain a well-dispersed suspension. Then, 500 μL~1 mL of titanium tetrabutoxide (TBOT, Sigma Aldrich, reagent grade, 97%) was rapidly added into the suspension under vigorous stirring in an ice water bath for 20 min. The suspension was then transferred into a three-neck flask that was immersed in a constant-temperature water bath at 80°C. Subsequently, the controlled hydrolysis of TBOT was initiated by adding ~20 mL deionised water/ethanol mixture (50 mL/L) dropwisely. As have shown in the results part, the thickness of the shell can be easily tuned by varying the experimental parameters, which primarily include the loading amount of VO2, the volume of TBOT and the concentration of H2O. The as-prepared core-shell powders were collected by filtration, washed twice with ethanol and then dried in a vacuum oven at 60°C for 6 h.

To obtain a crystalline coating, the as-synthesised powders were placed into an alumina crucible and then annealed in a vacuum quartz-tube furnace under a flow of air at specific air pressure and temperature. To avoid VO2 oxidation, the annealing process was optimised. Typically, with an air pressure of ~3 torr, the sample was heated from ambient temperature to 400°C at a heating rate of ~10°C /min and then held at 400°C for 15 min before naturally cooling to room temperature.

Characterisation

X-ray diffraction (XRD) analyses were conducted on a Rigaku Ultima IV diffractometer with Cu Kα radiation (λ = 1.5418 Å) using a voltage and current of 40 kV and 40 mA, respectively. Unless stated otherwise, all the samples were measured at a scanning rate of 3°/min. The microscopic morphology was obtained using a field-emission scanning electron microscope (FE-SEM, HITACHI S-3400, Japan) at an acceleration voltage of 15 kV. The microstructure and composition of the samples were further analysed using a transmission electron microscope (TEM, JEOL2010) with an energy-dispersive spectrometer (EDS) attachment. Scanning TEM was operated in dark field imaging mode. The Raman spectra of the composite samples were measured using a Raman microscope (Renishaw in Via) with a 514 nm laser source at an input power of 1 mW. This small power was applied to avoid local heating effects.The phase transition behaviours of the resulting products were measured by differential scanning calorimetry (DSC200F3, NETZSCH) over the temperature range from 0 to 90°C using a liquid nitrogen cooling unit. The heating and cooling rates were set at 10°C/min.

Photocatalytic reaction

~0.1 g of the powder samples were ultrasonically dispersed in 50 mL deionized water followed by the addition of 500 μL of a 0.01 M rhodamine B (RhB) aqueous solution. The mixture was then stirred under darkness for 2 h to achieve adsorption-desorption equilibrium. Subsequently, the suspension with continuous stirring was exposed under a Xe lamb (300 W) with an incident direction normal to the surface of the solution. At given irradiation intervals, ~3 mL aliquots of the suspension were collected and separated by centrifugation. The absorption spectrum of the supernatant was measured using a UV-Vis spectrometer (HITACHI U-3010). The concentration of RhB was determined by monitoring the changes in the absorbance maximum at ~554 nm. According to the Lambert-Beer law, the absorbance is proportional to the concentration of a dilute solution. The specific surface area was measured on an ASAP2010 (Micromeritics,U.S.A) instrument using the N2 adsorption method with BET analysis.

Evaluation of the thermochromic properties

Films made of the obtained samples were prepared by coating the powder onto a float glass substrate. For measuring, a piece of double-sided thermal-resistant teflon tape was sticked onto a glass substrate. A stock suspension with solid loading amount of 3wt% was prepared by dispersing a certain amount of powder ultrasonically in ethanol. Then a film was made by casting the suspension solution by anilox roll on the Teflon-tape. After natural drying, repeating this process for twice or 3 times to make sure the whole surface was coated by powders and then compressed air gun was used to blow away redundant powders which werenot sticked on the teflon surface. The diffuse transmittance (Tr) and reflectance (Re) measurements of the film at the wavelength range from 250 nm to 2600 nm at 25°C and 90°C were conducted using a UV-Vis spectrophotometer (HITACHI U-3010) with a integrating sphere and a temperature controlling unit. The transmittance curve of the sole tape-sticked glass is calibrated as baseline for transmittance measurements.