Next Article in Journal
Direct Steady-State Calculation of Electromagnetic Devices Using Field-Circuit Models
Next Article in Special Issue
Paper Review of External Integrated Systems as Photovoltaic Shading Devices
Previous Article in Journal
Analysis of GHG Emission from Cargo Vehicles in Megacities: The Case of the Metropolitan Zone of the Valley of Mexico
Previous Article in Special Issue
Enhancing Smart Home Design with AI Models: A Case Study of Living Spaces Implementation Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Approaching the Theoretical Maximum Performance of Highly Transparent Thermochromic Windows

1
The Netherlands Organisation for Applied Scientific Research (TNO), High Tech Campus 25, 5656 AE Eindhoven, The Netherlands
2
Brightlands Materials Center, Urmonderbaan 22, 6167 RD Geleen, The Netherlands
3
Design and Synthesis of Inorganic Materials (DESINe), Institute for Materials Research, Hasselt University, Agoralaan Building D, 3590 Diepenbeek, Belgium
4
IMEC vzw, IMOMEC Associated Laboratory, Wetenschapspark 1, 3590 Diepenbeek, Belgium
5
EnergyVille, Thor Park 8320, 3600 Genk, Belgium
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(13), 4984; https://doi.org/10.3390/en16134984
Submission received: 23 May 2023 / Revised: 21 June 2023 / Accepted: 22 June 2023 / Published: 27 June 2023
(This article belongs to the Special Issue Smart and Innovative Solutions for Adaptive Facade Systems)

Abstract

:
Thermochromic window coatings represent a promising technology to improve the energy efficiency of buildings in intermediate climates. With the technology approaching market introduction it is important to investigate its performance limits within smart windows and to identify existing development challenges. Here we analyze the theoretical maximum performance of thermochromic window coatings that modulate IR transmission whilst retaining high visible transparency. The set limitations lead to a theoretical maximum solar modulation of 39.1%. Within an insulated glazing unit (IGU), where at least 2 glass panes and a conventional low-e coating are required, this value is further reduced to 12.9%. We show that by carefully selecting a low-e coating with the highest compatibility to a thermochromic coated glass and by allowing 10% of modulation in the visible spectral range, the theoretical maximum can be increased to 23.1%, illustrating the importance to codesign and match both coatings within a smart window to reach optimum performance. Furthermore, we compared our current best-performing VO2:SiO2 composite coating within an IGU to the theoretical maximum. The analysis shows that with a solar modulation of 13.4%, the coating is currently at 59% of the theoretical maximum. Finally, we propose and discuss several strategies to proceed further toward the theoretical maximum.

1. Introduction

In the process of increasing the energy efficiency of buildings, most building components require optimization to minimize interaction with the outside environment. Buildings are made air-tight, and the insulation of walls, roofs, and floors is further increased to realize a static indoor climate [1]. The part of the building skin for which interaction with the outside is still desired, are windows. For windows, capturing natural light and solar heat is desired to a certain extent. It has been demonstrated that sunlight and solar heat have a positive influence on people’s health and well-being [2] and all solar energy that can be directly used in a building for lighting and heating reduces the overall energy costs [3]. However, when certain irradiance levels are surpassed, glare becomes an issue, and shading is desired. In addition, solar heat transmission through windows accounts for 60% of the total solar heat gain in buildings. As a result, overheating of buildings and air-conditioning demand to reduce indoor temperatures are largely based on solar transmission through windows [4]. Therefore, the requirements for visible (Tvis) and total solar (Tsol) transmission vary depending on the climate zone and the amount of window surface [3,5]. The glass industry is approaching these different demands by offering static glazing with different Tvis and g values (the g value equals the total solar energy transmitted through a window, radiative and conductive) [6]. A wide range of different products are available which have been optimized over the past decades. However, since in many climate regions a building’s demands on sunlight and solar heat transmission change during the day and between the seasons, static glazing can never capture the full potential of sunlight.
Traditionally, the inability of windows to adapt to the changing requirement for sunlight and solar heat transmission has been accommodated by using external or internal shading devices [7]. Since all these shading devices themselves have certain limitations, like not being able to distinguish between visible light and solar heat transmission, recent development has been focused on smart windows that can adapt their interaction with sunlight directly upon an external stimulus [8]. Electrochromic windows are the most advanced of all smart window technologies, with several products already available on the market [9,10]. Due to their complexity and the necessity for wiring and special frames, these technologies are expensive and therefore still limited to niche applications. Since the costs cannot be earned back via energy savings in a reasonable amount of time, electrochromic smart windows are focused on increasing comfort via actively adjusting the shading of the glass and therefore providing privacy, as well as reducing glare and solar heat gain [11]. Another smart window technology that is being developed is photochromic windows. Photochromic glazing technology is already commercially applied in sunglasses that increase shading at high solar irradiance levels. In the same way, photochromic windows can be used to increase shading and reduce glare [8]. As already inherent to the technology, photochromic materials are foremost applicable for glare control. It is difficult to optimize photochromic windows for minimum energy demand, due to the fact that irradiance levels are not the best trigger to determine when to transmit or block solar heat. For example, on sunny winter days, the high irradiance levels can induce tinting, although low ambient temperatures make solar heat gain still beneficial for the energy demand. Thermochromic materials are best suited to be optimized for energy efficiency alone [12,13]. They can be designed to minimize the impact on Tvis, whilst modulating infrared (IR) transmission and therefore solar heat gain. Temperature as the external trigger is best suited to couple with a building’s thermostat by adjusting the transition temperature in the thermochromic material. Additionally, the material does not require special equipment to trigger the switch and can be installed like regular windows. This way, thermochromic windows can be designed for use in regular insulated glazing units (IGUs) enabling optimum use of solar heat whilst retaining high visible transparency.
Because of their high application potential, thermochromic materials—mainly vanadium dioxide (VO2)—have been a research focus in the last years [12,13,14]. Materials for thermochromic smart windows have been optimized by increasing solar modulation (ΔTsol) values at high Tvis [15,16,17,18]. Switching temperatures have been optimized to cover a wide range between 0 and 68 °C [19,20], increasing the potential to tune the properties to requirements in various climate zones. Furthermore, processes are being scaled up and production costs are being reduced [17,21]. The potential of thermochromic smart windows has been analyzed in several simulation studies [22], including the impact of switching temperature and hysteresis [23], and it has been shown that these new technologies surpass commercial glazing in energy efficiency in intermediate climates [3]. Nevertheless, most of the studies optimizing the performance of thermochromic windows, as well as many simulation studies, are studying the optical properties of the thermochromic material isolated from the real application in smart windows. But windows are not only important for their solar heat gain, but they are also the most vulnerable part of the building skin with respect to radiator heat loss. To increase the insulation value of windows, the number and thickness of glass panes as well as the number, width and atmosphere of gas spaces between the glass panes can be adjusted. In addition, the glass panes can be coated with a low emissivity (low-e) coating, significantly reducing the thermal conductivity of an IGU [24]. Since these low-e coatings not only impact infrared radiation between 5–25 µm, and therefore the emissivity, but also near-IR transmission, they have to be accounted for when analyzing the impact of thermochromic glazing. To realize insulation values (u values) of thermochromic smart windows in agreement with current standards, a combination of thermochromic glass and low-e coated glass is required. We have shown in a recent simulation study that only with such a combination the positive effect of changing solar heat gain can be realized. However, the low-e coating properties of partly blocking solar near-IR light reduces the solar modulation of the complete IGU [3]. This shows the necessity of codesigning all components of the IGU and investigating a perfect combination of thermochromic and low-e properties.
Since thermochromic smart windows are approaching market introduction, it is important to investigate the performance limit of smart windows, to identify existing development challenges [25]. For that purpose, we study practical combinations of our thermochromic glass with best-in-class optical performance (combination of visible transmission and solar modulation) [15] with selected commercially available low-e glasses. Here we report a detailed study on the maximum theoretical solar modulation when focusing solely on the IR part of the solar spectrum. Furthermore, we investigate how the theoretical maximum changes when applied in an IGU with different low-e coatings, and when including 10% of visible transmission in the modulated range. Additionally, we compare one of our high-performance thermochromic smart window coatings to these theoretical maximum values to investigate which major challenges are to overcome, and we give an outlook on how to optimize these coatings to come closer to their theoretical maximum performance.

2. Methods

2.1. Theoretical Investigation

For investigation of potential transmission modulation via thermochromic smart windows for several combinations of different glazing options, the UV-vis-NIR transmission spectrum of a commercial glass or combination of two glasses in an IGU was compared to a manipulated version of the respective spectrum, where transmission was set to 0 in wavelength regions > 780 nm. The two spectra were compared by calculating the total solar transmission (Tsol) and the solar modulation (ΔTsol). Tvis, Tsol and ΔTsol were calculated using the following formulas:
T v i s , s o l = ϕ v i s , s o l λ T λ d λ ϕ v i s , s o l λ d λ
Δ T s o l = T s o l , c o l d T s o l , h o t
with T(λ) as the transmittance at wavelength λ, Φvis(λ) as the photopic spectral sensitivity of human vision, and Φsol(λ) as the AM 1.5 solar irradiance spectrum. For IGUs also modulations of g values (Δg) were calculated. Here g values were calculated according to EN410 [26]. For each IGU the following assumptions were made (Figure S1): (i) double glazing with 2 glass panes and 1 gas space; (ii) glass thickness 4 mm; (iii) gas space 12 mm; (iv) gas space atmosphere argon; (v) adaptive glass as outer glass pane with coating on surface 2; (vi) coating of inner glass pane on surface 3.
The following glasses and glazing systems have been used in the study (Table 1).
Optical properties of coated (Scheuten Isolide and TNO SunSmart) and uncoated (Pilkington Optifloat) glass plates have been measured via UV-vis-NIR spectrophotometry using a Perkin Elmer Lambda 750 (Perkin Elmer, Inc., Waltham, MA, USA) with integrating sphere and a customized module comprising a heating element equipped with a module for measuring transmission and reflection in the wavelength regions between 300 nm and 2400 nm. Optical data of commercial low-e coated glass (Saint-Gobain ECLAZ and Pilkington SunCool) have been used from the International Glazing Database (IGDB) within the WINDOW program (version 7.6) by Berkeley Labs [27]. The transmission spectra of all glasses used in this study are presented in Figure 1. Tvis and Tsol of glazing systems have been calculated using the Optics program by Berkeley Labs [27] and transmission spectra have been calculated using the following formula:
T g l a z i n g   s y s t e m λ = T g l a s s   p a n e   1 λ × T g l a s s   p a n e   2 λ
with T(λ) as the transmittance at wavelength λ.

2.2. Coating Preparation

The thermochromic coating used in this study was prepared in analogy to a previously published procedure [15]. One side of a barrier coated glass substrate was masked using d-c-fix® self-adhesive foil. Subsequently, the thermochromic coating was applied to the non-covered, tin side using dip coating. The applied liquid coating formulation consisted of a mixture of vanadyl oxalate solution and silica at a SiO2 to VO2 ratio of 2.2:1 w/w. The xerogel coating was applied to the glass substrates at a withdrawal speed in the range of 8.0 mm·s−1 to achieve the desired thickness. After drying for 5 min under ambient conditions, the adhesive foil was removed. The non-coated side was then placed onto a 6 inch silica wafer, and annealed in a Jipelec Jetfirst PV Rapid Thermal Processor (ECM, Grenoble, France), with the coated side facing the IR radiators. The coated glass plates were first heated to 270 °C in air for 480 s, and subsequently in N2 at 450 °C for 30 s.

3. Results and Discussion

3.1. Investigation of Theoretical Maximum Performance

To investigate the theoretical maximum solar modulation for adaptive glazing, we started by analyzing the wavelength dependent light transmission of clear glass. Here we measured the UV-vis-NIR transmission of a 4 mm Pilkington Optifloat Clear glass and calculated the solar properties. The wavelength dependent transmission spectrum can be seen in Figure 1. The glass plate showed a Tvis of 90.3% and a Tsol of 85.4% (Table 2). According to EN 410 [22], visible transmission is defined as the region between 380–780 nm. Therefore, if the solar modulation should be limited to the IR region, all light transmitted >780 nm could theoretically be blocked. To simulate this theoretical blocking state, the light transmission between 780–2500 nm was set to 0, whilst retaining the wavelength dependent transmission between 300–780 nm of the Optifloat Clear glass. The thereby obtained theoretical transmission spectrum of a total IR blocking state (blocking >780 nm) was used to calculate the theoretical solar properties. Since Tvis was not influenced by the data manipulation, the same value as for the Optifloat Clear glass was obtained, whereas the Tsol value dropped to 46.3% (Table 2). Directly comparing the transmission spectrum of Optifloat Clear glass with the theoretical IR blocking state shows that a large part of the transmission can be modulated even by solely focusing on the IR (Figure 2). Since solar intensity drops significantly in higher wavelength regions and the IR only accounts for approximately 45% of the total solar radiation, the maximum theoretical solar modulation ΔTsol between these two states is 39.1% (Table 2).
The analyzed optical properties of Optifloat Clear glass and the potential solar modulation of 39.1% for an adaptive glazing would be representative of single glazing. But this type of glazing has a very low insulation value of 5.8 W/m2 K and therefore leads to significant radiator heat loss. Current standard glazing for newly built homes in the Netherlands is so-called HR++ glazing with an insulation value of 1.1 W/m2 K, with similar standards for windows in most European countries. These are insulated glazing units (IGUs) made of 2 glass panes of which one is coated with a low-e coating and a gas space commonly filled with a low thermal conductivity gas, such as argon. This type of glazing has different optical properties from a single Optifloat Clear glass, especially in the IR. Therefore, to assess the theoretical maximum solar modulation of an adaptive glazing within a current state-of-the-art IGU, we investigated the optical properties of an IGU combining an Optifloat Clear glass with a Scheuten Isolide Superplus glass. In the UV-vis-NIR transmission spectrum, it can be seen that the Isolide Superplus glass still shows a high visible transmission, but the transmission in the IR drops significantly, especially in longer wavelength regions (Figure 1). In the IGU the transmission over the whole wavelength region drops slightly due to the transmission of the Optifloat glass between 80–90% for all wavelengths between 350–2500 nm (Figure 3). Overall, the representative HR++ glazing shows solar properties of Tvis = 79.8% and Tsol = 51.3% (Table 2). This means that the combination Optifloat x Isolide already blocks 48.7% of solar radiation from entering the building, which significantly reduces the amount of sunlight that can be modulated. To investigate the amount of IR light that can still be modulated within such an IGU, we simulated a theoretical IR blocking state with 0% IR transmission. To calculate the optical properties of such a glazing system, we combined the Isolide glass with the theoretical blocking > 780 nm glass. Such a combination showed again unchanged Tvis of 79.8% and a reduced Tsol of 38.4%. Comparing the UV-vis-NIR transmission spectrum of the Optifloat × Isolide combination with the theoretical blocking > 780 nm × Isolide state, it can be seen that the amount of IR light that can be modulated is reduced significantly in comparison to the Optifloat single glass (Figure 2 and Figure 3). Overall the theoretical maximum ΔTsol drops to 12.9% for the HR++ glass.
Even though the Scheuten Isolide is a representative of the most common low-e coated glasses used in IGUs for residential buildings, there are other low-e coatings on the market that block more or less IR light. Since adaptive glazing can in general be combined with any low-e coated glass to reach an insulation value of 1.1 W/m2 K, it is important to select the best suited low-e coated glass. The suitability here can be equated to the maximum possible solar modulation. To investigate the theoretical maximum solar modulation of IGUs with different low-e coated glasses and their suitability to be used in adaptive glazing systems, we calculated the solar properties of 2 additional IGUs with a highly transparent low-e coating and a so-called solar control coating, that blocks a large part of IR light. As highly transparent low-e coating we selected the Saint-Gobain ECLAZ series. Combined with an Optifloat clear glass in an IGU, a Tvis of 83.2% and Tsol of 60.3% were obtained (Table 2). The solar control glazing, where we selected Pilkington’s SunCool series as a representative case study, resulted in Tvis values of 73.7% and Tsol of 35.8% (Table 2). Here it could already be observed that the solar control coating blocks a large part of IR light and therefore leaves little room for modulation via adaptive glazing (Figure 1). To calculate ΔTsol values and to compare the different low-e coatings, we again calculated transmission spectra where we combined the individual low-e coated glasses with the theoretical blocking > 780 nm spectrum. By comparing the IGUs with different low-e coatings with each other as well as with the respective theoretical IR blocking states, it was observed that the combination Optifloat × ECLAZ has the highest potential modulation range in the IR (Figure 4). Here, for the blocking state, a Tsol of 41.3% was calculated, which results in a ΔTsol of 19.0% (Table 2). For the solar control glazing the blocking state had a Tsol of 32.7% which is not much lower than for the regular combination Optifloat × SunCool and therefore leads only to a ΔTsol of 3.1%. The potential solar modulation of the HR++ glazing lies at 12.9% between those two. Since the highly transparent low-e coating (Saint-Gobain ECLAZ) transmits the largest part of IR light for all 3 studied low-e coatings it has the highest potential and best suitability to be combined with adaptive glazing.
Even though VO2 based thermochromic coatings are aimed at staying highly transparent in the visible and solely modulating the IR region of sunlight, especially high-performing nanocomposite coatings [15] or coatings within anti-reflective coating stacks [28] also show a slight modulation in the visible range. Here, it has been shown that adding modulation of a few percent in the visible to a large modulation in the IR can significantly increase the overall ΔTsol of a thermochromic coating. Furthermore, if the change in visible transmission is kept at a few percent, the change is barely visible to the human eye and the coated glass still appears highly transparent. To investigate the potential solar modulation of an IGU with an ECLAZ low-e coated glass, where full IR modulation is combined with partial modulation of visible light, we simulated a second theoretical blocking transmission spectrum. In accordance with the blocking > 780 nm spectrum, we started from the Optifloat Clear transmission spectrum whilst setting transmission > 780 nm to 0. In addition, we reduced the transmission of the Optifloat transmission spectrum ≤ 780 nm by 10%. This theoretical blocking > 780 + 10% vis spectrum was combined with the ECLAZ spectrum to simulate a theoretical blocking state of the IGU (Figure 5). For this blocking state, a Tvis of 74.9% and Tsol of 37.2% were calculated (Table 2). Comparing this with the transmission of the Optifloat × ECLAZ IGU, it can be seen that Tvis is reduced by 8.3% and a ΔTsol of 23.1% is obtained. Since a larger modulation in the visible is not desired and the ECLAZ low-e coating shows the highest potential for solar modulation, it can be concluded that within these limitations a ΔTsol of 23.1% is the theoretical maximum that can be achieved with VO2 based thermochromic coatings within an IGU.

3.2. Current Record Performance

After assessing in detail the theoretical maximum solar modulation of a VO2 based thermochromic coating within an IGU, we compared the properties of our current best-performing coating to these values. The coating we selected is a VO2:SiO2 composite coating, where the composition and coating thickness is optimized to realize an optimum combination of high Tvis of 67.1% and large ΔTsol of 22.9%. A high Tvis is obtained via the integration of VO2 domains in an SiO2 matrix, which increases transparency and simultaneously reduces reflection via a reduced refractive index of the composite layer. In addition, the composition is optimized to create small evenly distributed VO2 domains with an average size < 30 nm, which leads to increased light absorption in the rutile state (hot state) in wavelength regions between 800–1300 nm. This effect for small VO2 domains within a coating matrix has been attributed in literature to an occurring plasmonic absorption of the metallic VO2 phase [29,30]. The increased absorption in this wavelength region has a big impact on the overall Tsol in the hot state, since the solar IR intensity is highest between 780–1360 nm (Figure S2). Since solar IR intensity decreases >1400 nm significantly, the higher transmission of our SunSmart coating in the hot state in these wavelength regions does not negatively impact the overall ΔTsol. By carefully designing the transmission of the thermochromic coating in both cold and hot states, the large ΔTsol of 22.9% can be reached. To investigate the optical properties of a Smart Window IGU, we calculated the transmission spectra by combining our thermochromic coated glass with an ECLAZ low-e coated glass (Figure 6). For the cold state, we obtained Tvis values of 61.7% and Tsol of 45.7%, whereas for the hot state, we obtained a reduced Tvis of 52.8% and Tsol of 32.3% (Table 2). Overall we obtained a reduction in Tvis of 8.9%, which is similar to the blocking > 780 +10% vis × ECLAZ combination, and a ΔTsol of 13.4%. A significant reduction in ΔTsol in comparison to the single thermochromic coated glass can be observed, which is due to the IR-blocking properties of the ECLAZ coating. However, here also the designed optical properties with an increased modulation in the region between 800–1300 nm are beneficial, since in this region the ECLAZ coating still has a relatively high transparency, whereas >1400 nm the transparency drops significantly. Comparing the performance of the SunSmart IGU with the theoretical maximum performance, it can be concluded that the ΔTsol of 13.4% is at 59% of the theoretical maximum.

3.3. Optimization Strategies to Approach Theoretical Maximum

With 59% of theoretical maximum performance, the current SunSmart coating already shows very good properties. However, optimization of the coating and the active material can further improve the performance to bring the resulting solar modulation closer to its theoretical maximum. By careful analysis of the composite coating, as well as optimization strategies in literature, we have identified four strategies to further optimize the solar modulation that will be elaborated in more detail.

3.3.1. Purity of Active Material

VO2 can exist in an amorphous, as well as in various crystalline phases, of which only the monoclinic phase shows thermochromic properties [31]. Furthermore, vanadium can exist in several oxidation states ranging from +2 up to +5 [32]. Therefore, VO2 in the oxidation state +4, can be easily oxidized or reduced. Since all vanadium species show distinct absorption in the visible, but only VO2(M) has thermochromic properties, the Tvis of a coating is impacted by the total vanadium content, whereas ΔTsol is only impacted by the VO2(M) content. Therefore, to optimize the combination of high Tvis with high ΔTsol it is important to increase the share of active VO2(M) in the total vanadium content of a thermochromic coating.
In an earlier study, we investigated the purity of VO2(M) in a VO2:SiO2 composite coating made from liquid precursors [15]. Here we could show that only 54% of the total vanadium content was crystalline VO2(M). Since no other crystal phases or oxidation states were found, the remaining content was most probably amorphous VO2. Therefore, further improving the purity of the VO2(M) content via careful adjustment of the coating formulation and thermal anneal process can further improve the thermochromic performance.
Another possibility to produce thermochromic VO2 coatings is by synthesizing VO2 pigments and integrating those in a coating matrix. Here the purity of the VO2(M) content can be more easily controlled, since the crystallization of VO2(M) is decoupled from the coating formation. However, also the synthesis of VO2 nanoparticles brings many challenges to achieving phase pure monoclinic material with a high switching enthalpy. Additionally, doping, which is required to reduce the switching temperature to application-oriented values, may also lead to more complex systems and other phases as side products which reduce the performance of the coating. Synthesizing VO2 nanoparticles via a top-down approach using wet bead milling, we obtained particles with a switching enthalpy of 28 J/g and an amorphous fraction between 30–40% [33]. Therefore, more than one-third of the nanoparticles are non-switching VO2. This will lead to a decrease in both the Tvis and ΔTsol. Using a bottom-up hydrothermal approach to synthesize VO2(M) nanoparticles, a switching enthalpy of 31 J/g with VO2(A) as a side fraction was obtained [33]. In top-down synthesis, the purity of the VO2(M) content can be improved via the synthesis of the VO2 powder via calcination. Furthermore, the bead milling can be further improved to reduce the loss in crystallinity that occurs during size reduction. Here, every percentage improvement in the purity of VO2(M) content within the nanoparticles will lead to improved thermochromic performance in the final coating.

3.3.2. Size and Distribution of Active Domains within the Matrix (Including Layer Thickness)

Particle size plays a key role in improving the optical properties of VO2 films and coatings. To avoid Mie scattering the particles are required to be smaller than 100 nm [34]. With bead milling experiments we have increased both the Tvis and ΔTsol by decreasing the particle size below 120 ± 10 nm [33]. Furthermore, the distribution of nanoparticles or active domains within the matrix of the coating or films is important to achieve an additional localized surface plasmon resonance (LSPR) in the metallic state of VO2, which leads to a pronounced minimum at λ ≈ 1200 nm in the transmission spectrum [30,35]. Simulation work from Zhu et al. [30] illustrates the pronounced transmission minimum when 40 nm sized VO2 particles are homogeneously dispersed in a coating matrix, whereas agglomeration leads to the disappearance of the LSPR absorption peak and optical properties that can be associated with a continuous VO2 layer. Due to the high solar IR intensity between 780–1360 nm, the additional absorption associated with the LSPR leads to a significant increase in ΔTsol.
VO2 content within a film or coating as well as layer thickness also play a crucial role in optimization of thermochromic properties. We have shown that changing the ratio between VO2 and SiO2 in a VO2:SiO2 composite coating at similar thickness, leads to an increased Tvis with reduced VO2 content, whilst retaining a similar ΔTsol [15]. Here the increased Tvis was associated with a change in vanadium rich domain size and morphology in TEM/EDX analysis, as well as with a lower overall VO2 content per surface area. Surprisingly, this lower VO2 content per surface area did not result in a significant reduction in ΔTsol, which indicates that the VO2 domain size and morphology have a significant impact on the thermochromic performance, even though the dispersion in these coatings was not good enough to result in the occurrence of an LSPR. Furthermore, we showed that by increasing the layer thickness at a constant VO2:SiO2 ratio, the ΔTsol could be increased significantly without leading to a major reduction in Tvis. This was already predicted by Schläfer et al. in a previous theoretical study on VO2:SiO2 composite coatings [36].

3.3.3. Integration into Antireflective Coating Stack

A common strategy within the literature to improve the thermochromic performance of VO2 coatings is to integrate a continuous VO2 layer into a multi-layer antireflective coating stack [12]. VO2 coatings in their monoclinic phase have a refractive index around 2.5, which leads to light reflection at the interfaces around 10% over the whole solar wavelength region between 300–2500 nm. In addition, the refractive index of rutile VO2 after phase transition is reduced to around 2 [37]. This leads, especially in the visible, to lower reflection and in continuous VO2 coatings to a higher visible transmission in the hot state than in the cold state [38]. Since solar intensity per wavelength is much higher in the visible than in the IR, such an increase in Tvis upon switching leads to a significant reduction in ΔTsol. By integrating a VO2 layer into an anti-reflective coating stack, the refractive indices and thicknesses of the individual layers within the stack can be matched to significantly increase the transmission in the monoclinic state. Since only exact matching leads to destructive interference and reduced reflection, the same effect does not occur when the refractive index of the VO2 layer changes upon switching. Therefore, the transmission in the hot state remains low, whilst the transmission in the cold state is increased. This simultaneously increases Tvis and ΔTsol and can lead to high solar modulations of up to 21.8% [28]. Even though in a composite coating, the refractive index of the layer is already adjusted via the coating matrix to get rid of the effect of increasing Tvis upon switch, a further reduction in reflection via integration into an anti-reflective coating stack can also improve the thermochromic properties.

3.3.4. Other Strategies

Recently, more advanced strategies to improve the thermochromic performance of VO2 coatings have been reported. Here a strategy is to combine VO2 with another thermochromic material that changes transmission in the visible [39]. This way the overall solar modulation can be improved significantly. However, with tinting in the visible an additional effect is added to the coating, which changes the application area and is therefore not directly comparable.
Another strategy is to take advantage of the changing solar angle of incidence in summer and winter. Here, periodically patterned VO2 coatings are prepared that enable transmission of sunlight at low angles of incidence without blocking by the VO2 layer, whereas at higher angles of incidence, the light hits the patterned surface at the VO2 domains and is therefore blocked [40]. Theoretically very high ΔTsol. values of up to 25.8% can be achieved with this strategy. However, the concept still has to be validated in real-life experiments.
Furthermore, optimization of thermochromic properties within an IGU can be achieved by focusing on optimizing the optical properties of the low-e coating. In this study, we have already selected the most suitable commercially available low-e coating to be combined with a thermochromic coated glass. However, still, a significant reduction in solar modulation is observed due to the IR-blocking properties of the low-e coating. To further improve the thermochromic properties of the complete IGU, it is also possible to further increase the transmission of the low-e coating. Here it is crucial to retain the low emissivity properties, to not reduce the insulation value of the glazing whilst increasing the thermochromic performance.

4. Conclusions

In conclusion, we have investigated the theoretical maximum performance of thermochromic VO2 coatings within a state-of-the-art IGU in detail. We have demonstrated that by focusing on the IR region of sunlight alone, the potential modulation range is limited to 39.1% for a single clear glass pane. We have shown that this range is further reduced by the low-e coated glass in HR++ glazing, which is necessary to reach the market standard insulation properties of IGUs. A careful selection of the most suitable low-e coating can limit this reduction and still enables a potential solar modulation of 19%, which can be increased to 23.1% when a small modulation of 10% in the visible is allowed. Larger modulation of visible light is not desired since the appearance of the IGU should remain (nearly) unchanged. In addition, we have demonstrated that our current best-performing VO2 coating shows a solar modulation of 13.4% in combination with a low-e coated glass within an IGU. This represents a performance of 59% of the theoretical maximum. Lastly, we have discussed various strategies to further improve the thermochromic performance of such an IGU to proceed toward the theoretical maximum performance. Here especially increasing the purity of active material, and optimizing active domain size, size distribution, and layer thickness seem to be the best strategies to reach this goal.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en16134984/s1, Figure S1: Built up of IGU; Figure S2: AM1.5 solar spectrum.

Author Contributions

D.M.: conceptualization, methodology, validation, writing—original draft, writing—review & editing, visualization; L.C.: investigation, validation, writing—original draft, writing—review & editing; C.P.K.Y.: validation, formal analysis, visualization; R.H.: validation; K.E.: conceptualization, reviewing; A.H.: supervision, resources, reviewing; M.K.V.B.: supervision, funding acquisition, reviewing; P.B.: conceptualization, writing—review & editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The research work in this paper was carried out within Brightlands Materials Center, a joint research initiative between TNO and the province of Limburg. The work is funded by the Dutch government as well as the province of Limburg. Additionally, the work was partly performed within the Interreg Vlaanderen-Nederland project Sunovate and the OP Zuid project LEEF.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Aditya, L.; Mahlia, T.M.I.; Rismanchi, B.; Ng, H.M.; Hasan, M.H.; Metselaar, H.S.C.; Muraza, O.; Aditiya, H.B. A review on insulation materials for energy conservation in buildings. Renew. Sustain. Energy Rev. 2017, 73, 1352–1365. [Google Scholar] [CrossRef]
  2. Beute, F.; de Kort, Y.A.W. Salutogenic Effects of the Environment: Review of Health Protective Effects of Nature and Daylight. Appl. Psychol. Health Well-Being 2014, 6, 67–95. [Google Scholar] [CrossRef]
  3. Mann, D.; Yeung, C.; Habets, R.; Vroon, Z.; Buskens, P. Comparative Building Energy Simulation Study of Static and Thermochromically Adaptive Energy-Efficient Glazing in Various Climate Regions. Energies 2020, 13, 2842. [Google Scholar] [CrossRef]
  4. Tsikaloudaki, K.; Theodosiou, T.; Laskos, K.; Bikas, D. Assessing cooling energy performance of windows for residential buildings in the Mediterranean zone. Energy Convers. Manag. 2012, 64, 335–343. [Google Scholar] [CrossRef]
  5. Jaber, S.; Ajib, S. Thermal and economic windows design for different climate zones. Energy Build. 2011, 43, 3208–3215. [Google Scholar] [CrossRef]
  6. Solar Control. Available online: https://www.pilkington.com/en-gb/uk/products/product-categories/solar-control (accessed on 30 November 2022).
  7. Kirimtat, A.; Koyunbaba, B.K.; Chatzikonstantinou, I.; Sariyildiz, S. Review of simulation modeling for shading devices in buildings. Renew. Sustain. Energy Rev. 2016, 53, 23–49. [Google Scholar] [CrossRef]
  8. Ke, Y.; Chen, J.; Lin, G.; Wang, S.; Zhou, Y.; Yin, J.; Lee, P.S.; Long, Y. Smart Windows: Electro-, Thermo-, Mechano-, Photochromics, and Beyond. Adv. Energy Mater. 2019, 9, 1902066. [Google Scholar] [CrossRef]
  9. SageGlass. Available online: https://www.sageglass.com/ (accessed on 30 November 2022).
  10. HALIO. Available online: https://halioinc.com/ (accessed on 30 November 2022).
  11. Granqvist, C.G.; Arvizu, M.A.; Bayrak Pehlivan, I.; Qu, H.-Y.; Wen, R.-T.; Niklasson, G.A. Electrochromic materials and devices for energy efficiency and human comfort in buildings: A critical review. Electrochim. Acta 2018, 259, 1170–1182. [Google Scholar] [CrossRef] [Green Version]
  12. Cui, Y.; Ke, Y.; Liu, C.; Chen, Z.; Wang, N.; Zhang, L.; Zhou, Y.; Wang, S.; Gao, Y.; Long, Y. Thermochromic VO2 for energy-efficient smart windows. Joule 2018, 2, 1707–1746. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, Z.; Zhang, L.; Zhou, Y.; Cui, Y.; Chen, Z.; Liu, Y.; Li, J.; Long, Y.; Gao, Y. Thermochromic Energy Efficient Windows: Fundamentals, Recent Advances, and Perspectives. Chem. Rev. 2023, 123, 7025–7080. [Google Scholar] [CrossRef]
  14. Chang, T.-C.; Cao, X.; Bao, S.-H.; Ji, S.-D.; Luo, H.-J.; Jin, P. Review on thermochromic vanadium dioxide based smart coatings: From lab to commercial application. Adv. Manuf. 2018, 6, 1–19. [Google Scholar] [CrossRef] [Green Version]
  15. Yeung, C.P.K.; Habets, R.; Leufkens, L.; Colberts, F.; Vroon, Z.; Mann, D.; Buskens, P. Phase separation of VO2 and SiO2 on SiO2-Coated float glass yields robust thermochromic coating with unrivalled optical properties. Sol. Energy Mater. Sol. Cells 2021, 230, 111256. [Google Scholar] [CrossRef]
  16. Mann, D.; Yeung, C.P.K.; Habets, R.; Van Zandvoort, R.; Leufkens, L.; Hupperetz, J.; Out, D.; Valckenborg, R.; Vroon, Z.; Buskens, P. SunSmart—The First Affordable, Energy Optimized Smart Window. IOP Conf. Ser. Earth Environ. Sci. 2022, 1085, 012060. [Google Scholar] [CrossRef]
  17. Chang, Q.; Wang, D.; Zhao, Z.; Ling, C.; Wang, C.; Jin, H.; Li, J. Size-Controllable M-Phase VO2 Nanocrystals for Flexible Thermochromic Energy-Saving Windows. ACS Appl. Nano Mater. 2021, 4, 6778–6785. [Google Scholar] [CrossRef]
  18. Okada, M.; Kuno, M.; Yamada, Y. Fast synthesis of monoclinic VO2 nanoparticles by microwave-assisted hydrothermal method and application to high-performance thermochromic film. Sol. Energy Mater. Sol. Cells 2023, 255, 112311. [Google Scholar] [CrossRef]
  19. Calvi, L.; Leufkens, L.; Yeung, C.P.K.; Habets, R.; Mann, D.; Elen, K.; Hardy, A.; Van Bael, M.K.; Buskens, P. A Comparative Study on the Switching Kinetics of W/VO2 Powders and VO2 Coatings and Their Implications for Thermochromic Glazing. Sol. Energy Mater. Sol. Cells 2021, 224, 110977. [Google Scholar] [CrossRef]
  20. Calvi, L.; van Zandvoort, R.; Leufkens, L.; Hupperetz, J.F.B.; Habets, R.; Mann, D.; Meulendijks, N.; Elen, K.; Verheijen, M.A.; Hardy, A.; et al. Thermochromic glass laminates comprising W/VO2 nanoparticles obtained by wet bead milling: An in-depth study of the switching performance. Sol. Energy Mater. Sol. Cells 2023, 257, 112350. [Google Scholar] [CrossRef]
  21. Kim, Y.; Yu, S.; Park, J.; Yoon, D.; Dayaghi, A.M.; Kim, K.J.; Ahn, J.S.; Son, J. High-throughput roll-to-roll fabrication of flexible thermochromic coatings for smart windows with VO2 nanoparticles. J. Mater. Chem. C 2018, 6, 3451–3458. [Google Scholar] [CrossRef]
  22. Mann, D.; Yeung, C.P.K.; Habets, R.; Vroon, Z.; Buskens, P. Building energy simulations for different building types equipped with a high performance thermochromic smart window. IOP Conf. Ser. Earth Environ. Sci. 2021, 855, 012001. [Google Scholar] [CrossRef]
  23. Warwick, M.E.A.; Ridley, I.; Binions, R. The effect of variation in the transition hysteresis width and gradient in thermochromic glazing systems. Sol. Energy Mater. Sol. Cells 2015, 140, 253–265. [Google Scholar] [CrossRef]
  24. Hutchins, M.G.; Platzer, W.J. The thermal performance of advanced glazing materials. Renew. Energy 1996, 8, 540–545. [Google Scholar] [CrossRef]
  25. Cao, X.; Chang, T.; Shao, Z.; Xu, F.; Luo, H.; Jin, P. Challenges and Opportunities toward Real Application of VO2-Based Smart Glazing. Matter 2020, 2, 862–881. [Google Scholar] [CrossRef]
  26. EN 410; Glass in Building-Determination of Luminous and Solar Characteristics of Glazing. Comite Europeen de Normalisation: Brussels, Belgium, 2011.
  27. International Glazing Database (IGDB). WINDOW Software and Optics Software, Lawrence Berkley National Laboratory, University of California. Available online: https://windows.lbl.gov/software (accessed on 30 November 2022).
  28. Sol, C.; Portnoi, M.; Li, T.; Gurunatha, K.L.; Schläfer, J.; Guldin, S.; Parkin, I.P.; Papakonstantinou, I. High-Performance Planar Thin Film Thermochromic Window via Dynamic Optical Impedance Matching. ACS Appl. Mater. Interfaces 2020, 12, 8140–8145. [Google Scholar] [CrossRef]
  29. Zhou, Y.; Huang, A.; Li, Y.; Ji, S.; Gao, Y.; Jin, P. Surface plasmon resonance induced excellent solar control for VO2@SiO2 nanorods-based thermochromic foils. Nanoscale 2013, 5, 9208–9213. [Google Scholar] [CrossRef]
  30. Zhu, J.; Zhou, Y.; Wang, B.; Zheng, J.; Ji, S.; Yao, H.; Luo, H.; Jin, P. Vanadium Dioxide Nanoparticle-based Thermochromic Smart Coating: High Luminous Transmittance, Excellent Solar Regulation Efficiency and Near Room Temperature Phase Transition. ACS Appl. Mater. Interfaces 2015, 7, 27796–27803. [Google Scholar] [CrossRef]
  31. Srivastava, A.; Rotella, H.; Saha, S.; Pal, B.; Kalon, G.; Mathew, S.; Motapothula, M.; Dykas, M.; Yang, P.; Okunishi, E.; et al. Selective growth of single phase VO2 (A, B, and M) polymorph thin films. APL Mater. 2015, 3, 026101. [Google Scholar] [CrossRef]
  32. Kniec, K.; Marciniak, L. Different Strategies of Stabilization of Vanadium Oxidation States in Lagao3 Nanocrystals. Front. Chem. 2019, 7, 520. [Google Scholar] [CrossRef]
  33. Calvi, L.; van Geijn, R.; Leufkens, L.; Habets, R.; Gurunatha, K.L.; Stout, K.; Mann, D.; Papakonstantinou, I.; Parkin, I.P.; Elen, K.; et al. The Impact of Bead Milling on the Thermodynamics and Kinetics of the Structural Phase Transition of VO2 Particulate Materials and Their Potential for Use in Thermochromic Glazing. Sol. Energy Mater. Sol. Cells 2022, 242, 111783. [Google Scholar] [CrossRef]
  34. Lockwood, D.J. Rayleigh and Mie Scattering. In Encyclopedia of Color Science and Technology; Luo, M.R., Ed.; Springer: New York, NY, USA, 2016. [Google Scholar]
  35. Wang, D.; Guo, D.; Zhao, Z.; Ling, C.; Li, J.; Hong, S.; Zhao, Y.; Jin, H. Surface modification-assisted solvent annealing to prepare high quality M-phase VO2 nanocrystals for flexible thermochromic films. Sol. Energy Mater. Sol. Cells 2019, 200, 110031. [Google Scholar] [CrossRef]
  36. Schläfer, J.; Sol, C.; Li, T.; Malarde, D.; Portnoi, M.; Macdonald, T.J.; Laney, S.K.; Powell, M.J.; Top, I.; Parkin, I.P.; et al. Thermochromic VO2–SiO2 nanocomposite smart window coatings with narrow phase transition hysteresis and transition gradient width. Sol. Energy Mater. Sol. Cells 2019, 200, 109944. [Google Scholar] [CrossRef]
  37. Currie, M.; Mastro, M.A.; Wheeler, V.D. Characterizing the tunable refractive index of vanadium dioxide. Opt. Mater. Express 2017, 7, 1697–1707. [Google Scholar] [CrossRef]
  38. Mlyuka, N.R.; Niklasson, G.A.; Granqvist, C.G. Mg doping of thermochromic VO2 films enhances the optical transmittance and decreases the metal-insulator transition temperature. Appl. Phys. Lett. 2009, 95, 171909. [Google Scholar] [CrossRef] [Green Version]
  39. Zhu, J.; Huang, A.; Ma, H.; Tong, K.; Ji, S.; Bao, S.; Cao, X.; Jin, P. Composite Film of Vanadium Dioxide Nanoparticles and Ionic Liquid-nickel-chlorine Complexes with Excellent Visible Thermochromic Performance. ACS Appl. Mater. Interfaces 2016, 8, 29742–29748. [Google Scholar] [CrossRef]
  40. Zhou, C.; Li, D.; Tan, Y.; Ke, Y.; Wang, S.; Zhou, Y.; Liu, G.; Wu, S.; Peng, J.; Li, A.; et al. 3D Printed Smart Windows for Adaptive Solar Modulations. Adv. Opt. Mater. 2020, 8, 2000013. [Google Scholar] [CrossRef]
Figure 1. UV-vis-NIR transmission spectra of coated and uncoated glass plates used in this study.
Figure 1. UV-vis-NIR transmission spectra of coated and uncoated glass plates used in this study.
Energies 16 04984 g001
Figure 2. UV-vis-NIR transmission spectrum of Optifloat Clear glass (full line) and theoretical spectrum with 0% transmission > 780 nm (dotted line) with potential modulation range marked as dotted.
Figure 2. UV-vis-NIR transmission spectrum of Optifloat Clear glass (full line) and theoretical spectrum with 0% transmission > 780 nm (dotted line) with potential modulation range marked as dotted.
Energies 16 04984 g002
Figure 3. UV-vis-NIR transmission spectrum of Optifloat × Isolide glazing system (full line) and theoretical blocking > 780 nm × Isolide state with 0% IR transmission (dotted line) with potential modulation range marked as dotted.
Figure 3. UV-vis-NIR transmission spectrum of Optifloat × Isolide glazing system (full line) and theoretical blocking > 780 nm × Isolide state with 0% IR transmission (dotted line) with potential modulation range marked as dotted.
Energies 16 04984 g003
Figure 4. UV-vis-NIR transmission spectra of glazing systems with different low-e coated glasses (full line) and all whilst theoretically blocking all light > 780 nm (dotted line) with potential modulation range marked as dotted.
Figure 4. UV-vis-NIR transmission spectra of glazing systems with different low-e coated glasses (full line) and all whilst theoretically blocking all light > 780 nm (dotted line) with potential modulation range marked as dotted.
Energies 16 04984 g004
Figure 5. UV-vis-NIR transmission spectrum of Optifloat × ECLAZ glazing system (full line), theoretical blocking > 780 nm × ECLAZ state with 0% IR transmission (dotted line) and additional blocking in the visible (blocking > 780 + 10% vis; small dotted line) with potential modulation ranges marked as dotted.
Figure 5. UV-vis-NIR transmission spectrum of Optifloat × ECLAZ glazing system (full line), theoretical blocking > 780 nm × ECLAZ state with 0% IR transmission (dotted line) and additional blocking in the visible (blocking > 780 + 10% vis; small dotted line) with potential modulation ranges marked as dotted.
Energies 16 04984 g005
Figure 6. UV-vis-NIR transmission spectrum of SunSmart × ECLAZ glazing system in the cold (full line) and hot state (dotted line) with modulation range marked as dotted.
Figure 6. UV-vis-NIR transmission spectrum of SunSmart × ECLAZ glazing system in the cold (full line) and hot state (dotted line) with modulation range marked as dotted.
Energies 16 04984 g006
Table 1. Glasses and glazing systems used in this study.
Table 1. Glasses and glazing systems used in this study.
GlassGlazing System (Outer Glass Pane × Inner Glass Pane)
Product NameManufacturerType of GlassThickness
Optifloat ClearPilkington, Lathom, UKMonolithic4 mm
ECLAZSaint-Gobain, Courbevoie, FranceCoated4 mmOptifloat × ECLAZ
Isolide SuperplusScheuten Glas, Venlo, The NetherlandsCoated4 mmOptifloat × Isolide
SunCool 70/40ProPilkington, Lathom, UKCoated6 mm *Optifloat × SunCool
SunSmart 67/23TNO, Geleen, The NetherlandsCoated/thermochromic4 mmSunSmart × ECLAZ
* The SunCool series was only available in 6 and 12 mm thickness in the database. However the deviation in glass thickness is not expected to have a major impact on the conducted analysis.
Table 2. Solar properties of measured and theoretical glass and glazing systems with potential solar modulation for various scenarios.
Table 2. Solar properties of measured and theoretical glass and glazing systems with potential solar modulation for various scenarios.
Glass/Glazing SystemScenarioTvis [%]Tsol [%]ΔTsol [%]
Optifloat ClearFull transmission90.385.439.1
blocking > 780 nm90.346.3
Optifloat × ECLAZFull transmission83.260.319.0
blocking > 780 nm83.241.3
blocking > 780 + 10% vis74.937.223.1
Optifloat × IsolideFull transmission79.851.312.9
blocking > 780 nm79.838.4
Optifloat × SunCoolFull transmission73.735.83.1
blocking > 780 nm73.732.7
SunSmartCold state67.167.722.9
Hot state57.444.8
SunSmart × ECLAZCold state61.745.713.4
Hot state52.832.3
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

Mann, D.; Calvi, L.; Yeung, C.P.K.; Habets, R.; Elen, K.; Hardy, A.; Van Bael, M.K.; Buskens, P. Approaching the Theoretical Maximum Performance of Highly Transparent Thermochromic Windows. Energies 2023, 16, 4984. https://doi.org/10.3390/en16134984

AMA Style

Mann D, Calvi L, Yeung CPK, Habets R, Elen K, Hardy A, Van Bael MK, Buskens P. Approaching the Theoretical Maximum Performance of Highly Transparent Thermochromic Windows. Energies. 2023; 16(13):4984. https://doi.org/10.3390/en16134984

Chicago/Turabian Style

Mann, Daniel, Lavinia Calvi, Cindy P. K. Yeung, Roberto Habets, Ken Elen, An Hardy, Marlies K. Van Bael, and Pascal Buskens. 2023. "Approaching the Theoretical Maximum Performance of Highly Transparent Thermochromic Windows" Energies 16, no. 13: 4984. https://doi.org/10.3390/en16134984

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