Dingkun
Wang†
,
Guoqi
Chen†
and
Jun
Fu
*
Key Laboratory of Polymeric Composite and Functional Materials of Ministry of Education, Guangdong Engineering Technology Research Centre for Functional Biomaterials, School of Materials Science and Engineering, Sun Yat-Sen University, Guangzhou, 510275, China. E-mail: fujun8@mail.sysu.edu.cn
First published on 25th April 2024
Windows are an important part of buildings and transmit light between indoors and outdoors. Frequent heat exchange through windows increases building energy consumption. Smart windows can change optical properties and modulate solar radiation, which are recognized as frontrunners in building energy saving. Among various smart windows, thermochromic windows usually passively regulate light transmittance in response to environmental temperature and have showed great potential for practical applications. Thermochromic materials are key to constructing thermochromic smart windows. Usually, a reversible phase transition takes place for thermoresponsive materials near the critical transition temperature, leading to changes in transmittance over different spectrum bands. Representative thermochromic materials include metal oxides, hydrogels, perovskites, ionic liquids, liquid crystals, etc. The intrinsic phase transition temperature, luminous transmittance, and solar modulation ability are among the critical parameters defining the performance of smart windows. New strategies have been developed to modulate the performance of thermochromic materials and smart windows to meet demands from different environments and climates. Such endeavors have boosted smart windows to modulate full-spectrum solar regulation and to achieve efficient all-climate building energy saving. Next generation smart windows will not only modulate solar transmission, but also convert and store solar energy through new power technologies such as thermoelectricity conversion and solar cells. Challenges and future prospects of smart windows are discussed to inspire future building energy saving.
Smart windows have been developed to intelligently regulate indoor solar radiation in response to external stimuli and consequently reduce heat exchange to conserve energy,5,6 including electrochromic,7 mechanochromic,8 magnetochromic,9 photochromic,10,11 and thermochromic smart windows.12,13 The former three types are known as active smart windows that need extra energy input to trigger and maintain the “ON” or “OFF” state. The latter two types are categorized as passive smart windows that spontaneously respond to environmental conditions without extra power input.
Thermochromic windows can modulate indoor solar irradiation by reversibly regulating the transmittance or reflectance of ultraviolet (UV, 250–380 nm), visible (380–780 nm) and near-infrared (NIR, 780–2500 nm) solar radiation.14,15 Thermoresponsive materials that undergo phase transitions and reversible transparent-to-opaque transitions upon temperature changes have been utilized for thermochromic smart windows for solar modulation.16,17 To date, many inorganic and organic thermochromic materials have been used for smart windows, including metal oxides13,18,19 and hydrogels.20–22 VO2, for example, experiences a phase transition at 68 °C, above which the IR transmittance is decreased, while the transmittance in the visible band barely changes.12,13,15 Besides, the solubility and/or hydrophilicity/hydrophobicity of polymer chains in thermochromic hydrogels can change at different temperatures, which causes a phase transition, chain conformation change, or precipitation of polymer chains.20,23 Such transitions can cause variations in color or transparency to modulate solar modulation.24,25 Other thermochromic materials, including ionic liquids, perovskites and liquid crystals, have also been used for thermochromic smart windows.26
The performances of smart windows are defined by several parameters including phase transition temperature (τc), luminous transmittance (Tlum), and solar modulation ability (ΔTsol). τc determines the optimal working conditions and potential additional methods needed to trigger its transition. A transition temperature close to the target application conditions is desired. Tlum is important for daily use. High Tlum is desired when smart windows are not triggered to block solar radiation. ΔTsol is defined as the solar transmittance change before and after the phase transition, which characterizes the ability to block solar energy. For light-harvesting smart windows, the power conversion efficiency (η) is a vital indicator that reflects the electrical output capability by converting the incident energy from the sun. The values of these parameters can be calculated using the following equations:
![]() | (1) |
![]() | (2) |
ΔTsol = Tsol(T < τc) − Tsol(T > τc) | (3) |
![]() | (4) |
Next generation smart windows are expected to harness solar energy instead of simply blocking and wasting it through scattering.27,28 New concepts to integrate thermoelectricity conversion and solar cells in smart windows can harness solar energy and convert it into electricity, thus maximizing building energy exploitation.28 Therefore, the next generation design principle of thermochromic smart windows requires intelligent management of incident sunlight according to temperature and rational utilization of solar light for power generation through photovoltaic technologies.
There are several review articles on thermochromic smart windows. Long's group has reviewed progress on thermochromic smart windows based on VO2,13 hydrogels,20 and some other thermochromic materials.12 Jin's group discussed thermochromic VO2-based smart windows for practical applications.29 Those review articles mainly focussed on single thermochromic materials for smart windows. Systematic and comparative investigations on smart windows based on different materials are needed to further discuss the influence of chemical structures, nanostructures, thermochromic properties, and device design on the thermochromic performance of smart windows, which will inspire novel strategies to develop high performance and functional smart windows to match application demands. In this review, we present a comprehensive review of high performance smart windows based on thermochromic materials. The intrinsic properties and response mechanisms of representative thermochromic materials, including VO2, hydrogels, ionic liquids, perovskites, and liquid crystals, and their applications to smart windows are discussed in detail (Fig. 1). Novel strategies ranging from molecule engineering to hybridizing, copolymerization, compositing, and nano-/micro-structure engineering to develop high performance thermochromic materials to match application demands under diverse conditions and climates are introduced. In each section, current challenges and potential development directions are analyzed. Next generation smart windows by utilizing thermoelectricity conversion and solar cell technology to harness reflected solar energy for power generation are discussed (Fig. 1). Future development of thermochromic smart windows in building energy saving is discussed. This review aims to inspire research interest and innovations to promote the development of smart windows for building energy saving.
![]() | ||
Fig. 2 (a) Schematic of the atomic structure for the metallic VO2 R phase and the insulating VO2 M phase. The V–V distances in each crystal structure are highlighted. Reproduced with permission from Whittaker et al., J. Phys. Chem. Lett. 2, 745 (2011). Copyright 2011 American Chemical Society.34 (b) Band structures of the metallic R and insulating M phases of VO2 depicted by molecular orbital diagrams. Reproduced with permission from Cui et al., Joule 2, 1707 (2018). Copyright 2018 Elsevier, Inc.13 |
Currently, diverse mechanisms have been proposed for the phase transition of VO2.37–40 The electron–phonon interaction mechanism deems that the appearance of metallic rutile phase bandgap (R) comes from the distortion of the monoclinic phase (M).41 The electron correlation mechanism suggests that the strong electron–electron correlation produces an insulating band gap.42–44 The reversible phase transition of VO2 (M/R) is described by using crystal field and molecular orbital theory (Fig. 2b). Briefly, in the rutile VO2 (R), the 2p orbital of O2− and 3d orbital of V4+ are hybridized to generate a narrow anti-bonding orbital π* and a wide bonding π. Besides, V4+ offers an anti-bonding orbital d// parallel to the c axis. The semi-filled d// and π* bands partially overlap at the Fermi level, showing metallic characteristics. But in the monoclinic VO2 (M), the d// and π* orbitals separate. The d// band splits into a filled orbital and an empty orbital with a band gap of 0.7 eV. Therefore, the electrons are trapped in the low energy d// orbital and VO2 (M) becomes insulating.
Thermochromic properties of representative VO2-based materials for smart windows are summarized in Table 1. Elemental doping can lower the τc of VO2 to about 20–50 °C, depending on the elements used. Meanwhile, the ΔTsol of the VO2 film can be improved to around 10–20% by constructing multilayer films. The Tlum of the VO2 film can be increased to about 40–85% through fabricating nanocomposite films or nano-/microstructures. These results represent significant progress toward practical applications of VO2-based smart windows.
Category | Materials/structures | T lum (%) | ΔTsol (%) | τ c (°C) | Ref. |
---|---|---|---|---|---|
Element doping | W-doped VO2 coating | 45.0 | 10.0 | 22.0 | 58 |
W-doped VO2 film | 54.4 | 10.7 | 39.0 | 54 | |
Zr-doped VO2 film | 61.4 | 10.3 | — | 59 | |
Mg-doped VO2 film | 59.4 | 9.5 | — | 59 | |
Si-doped VO2 film | 54.7 | 13.9 | — | 46 | |
H-doped VO2 | — | — | 30.0 | 60 | |
W/Zr co-doped VO2 film | 60.7 | 10.6 | 46.9 | 61 | |
SiO2/W co-doped VO2 film | 60.0 | — | — | 62 | |
W/Mg co-doped VO2 film | 46.2 | 10.8 | 36.9 | 63 | |
Hf/W co-doped VO2 film | 41.1 | 13.1 | 38.9 | 64 | |
SiO2/VO2 double-layered film | 55.0 | — | 70.0 | 65 | |
VO2/TiO2 double-layered film | 61.5 | 15.1 | — | 66 | |
VO2/TiO2 double-layered structure | 49.0 | 7.0 | — | 67 | |
Si–Al/VO2 double-layered coating | 44.0 | 18.9 | — | 68 | |
Multilayer films | TiO2/VO2/TiO2/VO2/TiO2 multilayered film | 45.0 | 12.1 | 60.0 | 55 |
SiNx/NiCr/NiCrOx/VOx/NiCrOx/NiCr/SiNx multilayered film | 40.5 | 18.4 | 54.0 | 48 | |
SiNx/VO2/SiNx multilayered film | 40.4 | 14.5 | — | 49 | |
SiO2/VO2/SiO2/polymer multilayered coating | 54.0 | 16.4 | — | 69 | |
WO3/VO2/WO3 multilayered structure | 55.4 | — | 52.0 | 70 | |
Nanocomposite films | VO2 nanoparticles in a Ni-based thermochromic system | 73.4 | 18.2 | — | 71 |
VO2 nanoparticles and Sb-doped SnO2 nanoparticles | 60.1 | 20.0 | 66.3 | 72 | |
VO2 nanoparticles and Sb-doped SnO2 nanoparticles | 84.4 | 11.6 | 84.5 | 72 | |
HfO2/VOx nanocomposite film | 51.6 | 15.4 | 60.6 | 73 | |
VO2 nanoparticles with SiO2 aerogel particles | 41.2 | 18.4 | — | 74 | |
VO2 nanoparticles in a polyurethane matrix (single layer) | 54.0 | 14.5 | — | 50 | |
VO2 nanoparticles in a polyurethane matrix (three layers) | 46.8 | 20.0 | — | 50 | |
VO2 nanoparticles in polyvinyl butyral | 43.4 | 17.3 | — | 51 | |
Core–shell VO2@SnO2 nanoparticles | 47.5 | 25.0 | ∼65.0 | 75 | |
VO2 nanoparticles in a polyvinyl pyrrolidone matrix | 57.3 | 13.8 | — | 76 | |
W-doped VO2 nanoparticle film | 50.0 | 10.0 | 32 | 77 | |
Double-layered nanoparticle array of VO2 | 46.1 | 13.2 | — | 78 | |
Nanostructure/microstructure | Three-dimensional ordered microporous VO2 | 71.1 | 10.8 | — | 52 |
Nanoporous VO2 film | 78.0 | 14.1 | — | 79 | |
VO2 nanoparticles film with VO2 clusters | 46.3 | 11.2 | — | 80 | |
VO2 mesh film | 86.0 | — | — | 81 | |
Micro-patterned VO2 film | 43.3 | 14.9 | — | 53 |
![]() | ||
Fig. 3 (a) The V–V chains along the c axis in pure and H-doped VO2 (R). Reproduced with permission from Cui et al., Phys. Chem. Chem. Phys. 17, 20998 (2015). Copyright 2015 Royal Society of Chemistry.60 (b) The calculated super-cell structures of pure V32O64 and Hf4W1V27O64. Reproduced with permission from Wang et al., Appl. Phys. Lett. 118, 192102, (2021). Copyright 2021 American Institute of Physics.64 |
![]() | ||
Fig. 4 (a) Schematic illustration of the refractive index-tunable bilayer film with nano-VO2. Reproduced with permission from Liu et al., J. Alloys Compd. 731, 1197 (2018). Copyright 2018 Elsevier, Inc.68 (b) XRD pattern and cross-section SEM image of a multilayer SiNx/VO2/SiNx film. Reproduced with permission from Long et al., Sol. Energy Mater Sol. Cells 189, 138 (2019). Copyright 2019 Elsevier, Inc.49 (c) Three-layered Ag/VO2/Pt film and Marilyn Monroe's image composed of different SiO2 thicknesses at different temperatures. Reproduced with permission from Kim et al., Appl. Surf. Sci. 565, 150610 (2021). Copyright 2021 Elsevier, Inc.86 |
Multilayer structures balance the trade-off between ΔTsol and Tlum and further promote the solar modulation ability. A SiNx/NiCr/NiCrOx/VOx/NiCrOx/NiCr/SiNx multilayer film deposited on a glass substrate shows a solar modulation of 18.4% and luminous transmittance of 40.5%.48 The multilayer SiNx/VO2/SiNx film consisting of a ∼80 nm VO2 layer and two SiNx layers has enhanced solar modulation from 10.8% to 14.5% and luminous transmittance from 36.1% to 40.4% (Fig. 4b).49 Moreover, a dip-coated four-layered SiO2/VO2/SiO2/polymer film exhibits a solar modulation of 16.4% and luminous transmittance of 54.0%.69 Furthermore, a three-layer Ag/VO2/Pt film is fabricated as a metal–dielectric–metal structure (Fig. 4c).86 The trilayer film shows a VO2 thickness-dependent reflective color. As the insulating VO2 layer undergoes a phase transition, the reflective color changes in response to temperature changes, leading to a unique solar modulation ability (Fig. 4c).
SiO2 is often used to modify the optical performances of pure VO2 films because of its outstanding visible transparency and low-cost coating process.13 VO2/SiO2 composite films are constructed through a facial mechanical ball-milling technology (Fig. 5).74 The SiO2 aerogel particles are coated on VO2 particles to prevent aggregation, which enhances the solar modulation (ΔTsol = 18.4%) and luminous transmittance (Tlum = 41.2%). Similarly, VO2 is composited in polyurethane with good dispersion and exhibits enhanced solar modulation from 6.6% to 14.5% and a luminous transmittance of 54%.50 However, the abovementioned research focuses on enhancing luminous transmittance and solar modulation ability of VO2. The high phase transition temperature of VO2 is rarely addressed. New VO2-based thermoresponsive materials with both moderate transition temperatures and excellent dispersion could be anticipated by compositing element-doped VO2 nanoparticles into matrices.
![]() | ||
Fig. 5 The mechanism of solar transmission increment of a VO2/SiO2 composite film. Reproduced with permission from Kang et al., Appl. Surf. Sci. 573, 151507 (2022). Copyright 2022 Elsevier, Inc.74 |
Apart from ordered VO2 structures, random VO2 nanoporous structures with the LSPR effect can also improve luminous transmittance and solar modulation ability. VO2 films with spontaneous random nanoporous structures are prepared to enhance the optical performances (ΔTsol = 14.1% and Tlum = 78%) (Fig. 6a).79 The VO2 film with random teeth-like nanoparticles and nanopores enhances absorption of LSPR in the NIR range and thus increases the Tlum and ΔTsol. The Tlum of VO2 films at different incident angles is enhanced by structure design. Similarly, a VO2 film with random large pores and isolated optical interaction particles is obtained by intense pulsed light (IPL) sintering to regulate light transmittance performance (Fig. 6b).80 The film sintered at 2000 V presents a visible light transmittance of 63.6% at 550 nm and luminous transmittance over 46.3%.
![]() | ||
Fig. 6 (a) Schematic diagram of fabrication route for nanoporous VO2 and optical properties of nanoporous VO2 films. Reproduced with permission from Long et al., ACS Appl. Mater. Interfaces 11, 22692 (2019), Copyright 2019 American Chemical Society.79 (b) Schematic illustration of VO2 nanoparticle films prepared by intense pulsed light and spin coating. Reproduced with permission from Kim et al., Mater. Des. 176, 107838 (2019). Copyright 2019 Elsevier, Inc.80 |
Hydrogels may have the upper critical solution temperature (UCST) and the lower critical solution temperature (LCST). Above LCST or below UCST, polymer chains in the hydrogel network are less soluble and adopt collapsed conformation or phase separated microdomains that scatter incident light. As a result, hydrogels appear opaque or translucent. At temperatures below LCST or above UCST, the polymer chains become soluble and swollen and the hydrogels become transparent.
LCST hydrogels at temperatures below τc are transparent and allow incident light to easily pass through thermochromic hydrogels (Fig. 7).23 However, at temperatures above τc, the hydrophobic polymer chains contract and aggregate to resist hydrophilic water molecules, and then phase separation occurs inside the hydrogels to form collapsed polymer clusters that can effectively scatter incident light to impair the visible transmittance of hydrogels. This is fundamental for solar modulation of thermochromic hydrogels.
![]() | ||
Fig. 7 Schematic illustration of thermochromic mechanism of thermochromic hydrogels. Reproduced with permission from Chen et al., Adv. Mater. 35, 2211716 (2023). Copyright 2023 Wiley-VCH GmbH.23 |
Zhou et al. reported the first PNIPAM hydrogel-based smart windows with a τc of around 32 °C.100 The luminous transmittance of hydrogel films decreases with an increase in thickness and shows an optimal Tlum of 70.7% at 52 μm thickness. The PNIPAM hydrogel-based smart window becomes opaque above τc and shows a ΔTsol of 25.5%. It is needed to modulate the intrinsic phase separation behavior of PNIPAM to further improve both the Tlum and ΔTsol for high performance thermochromic smart windows.
One solution is to modulate the scattering behavior of PNIPAM-based hydrogels. Fang et al. synthesized microparticles of a poly(N-isopropylacrylamide)-2-aminoethylmethacrylate hydrochloride (PNIPAM-AEMA) copolymer with large diameters and uniform crosslinks.16 The large particle size (∼1388 nm) and homogeneous internal structures with a very low light scattering contrast provide a high Tlum of 87.2% at 25 °C (Fig. 8a). In contrast, the particles shrink to about 546 nm at 35 °C, which scatter light over a broad spectrum, and thus show a high ΔTsol of 81.3% (Fig. 8b).
![]() | ||
Fig. 8 (a) The phase transition of the PNIPAM-AEMA hydrogel-based smart window triggered by the hand. (b) Thermochromic performances of PNIPAM-AEMA hydrogel and other reported materials. Reproduced with permission from Li et al., Joule 3, 290 (2019). Copyright 2019 Elsevier, Inc.16 |
A homogeneous hydrophilic network, and thus high Tlum, can be further achieved by introducing hydrophilic chains through free radical polymerization at low temperatures to avoid unwanted phase separation induced by reaction heating. Hydrophilic hydroxypropylmethyl cellulose (HPMC) is mixed with N-isopropylacrylamide for free radical polymerization at 0 °C.97 The heat released during the polymerization process could be dispersed in time. No phase separation takes place as that observed during high temperature-induced free radical polymerization of N-isopropylacrylamide. Here, the abundant hydroxyl groups of HPMC could form extensive hydrogen bonds with PNIPAM chains and water. The PNIPAM and HPMC chains form a uniform distribution, leading to a high transparency Tlum of 90.8% at low temperature.
A hydroxypropyl cellulose (HPC) hydrogel is another important LCST hydrogel.101–103 It possesses both hydrophobic epoxy ether groups and hydrophilic hydroxyl groups and has a miscible–immiscible transition at 42 °C.104,105 The random coil chains of HPC transform into collapsed spheres above 42 °C, resulting in an opaque phase that scatters solar radiation. The high LCST of HPC presents solar modulation at temperature much higher than the comfortable room temperature.
Zhang et al. constructed smart windows by using a hydroxypropyl cellulose-polyacrylic acid (HPC-PAA) hydrogel (Fig. 9a). The LCST of the HPC-PAA hydrogel is sensitive to solution pH and can be tuned down to 26.5 °C at pH 2.5.106 The hydrogel has massively interlaced pores (∼13 μm) because of the incorporation of PAA. The enlarged pores with a low scattering contrast increase Tlum to 90.1% at 18 °C. The pore diameter is reduced due to phase separation at 40 °C, which enhances light scattering, and thus the hydrogel shows a ΔTsol of 47.5%. Afterwards, photothermal cesium tungsten oxide (Cs0.32WO3) nanoparticles are further introduced to the HPC-PAA hydrogel to enhance NIR shielding performance (Fig. 9b).107 Cs0.32WO3 can transform the absorbed NIR light into heat through local surface plasmon resonance. Thus, the composite hydrogel reduces the NIR transmittance from 81.7% to 41.6% at 22 °C.
![]() | ||
Fig. 9 (a) The transmittance modulation mechanism of HPC-PAA hydrogel. Reproduced with permission from Zhang et al., ACS Appl. Energy Mater. 4, 9783 (2021), Copyright 2021 American Chemical Society.106 (b) The schematic illustration of the phase separation of photothermal nanoparticles-embedded HPC-PAA hydrogel. Reproduced with permission from Zhang et al., Ceram. Int. 48, 37122 (2022). Copyright 2022 Elsevier, Inc.107 |
Combining the temperature-responsive hydrogel mechanism with other mechanisms can endow smart windows with richer modulation abilities. Gao et al. constructed a smart window of HPC-PAA hydrogel doped with H+ and Li+.108 The phase transition temperature is adjusted through changing the hydrogen bond between molecular chains by ion doping. By integrating the ion doped thermochromic HPC-PAA hydrogel layer with the electrochromic WO3-indium tin oxide (ITO) layer, the smart window possesses four reversibly switch modes: transparent state, thermochromic state, electrochromic state, and electrothermal dual response state. Thus, the HPC-PAA hydrogel-based smart window shows a electric-/thermal-dual response.
Chung et al. reported a highly flexible smart window based on PAH hydrogels that allows visible light to transmit at high temperature and blocks mid-infrared radiation at low temperature.109 Such smart windows not only protect personal privacy but also prevent heat loss through blackbody radiation. A hydrophobic monomer (3-(methacryloylamino)propyl)trimethylammonium chloride (MPTC) is introduced to modulate the phase separation behavior of PAH. The PAH-MPTC hydrogel presents a phase transition temperature of 43.4 °C. The particle size (∼1 μm) and homogeneous pore structures at 55 °C provide a transmittance of 82% at 800 nm wavelength. The hydrogel with bubble-like porous structures (∼40 μm) at 20 °C represents macroscopic phase separation, leading to a good mid-infrared modulation ability. By combining a printed electrical heater with the PAH-MPTC hydrogel, this PAH hydrogel-based smart window could be actively controlled. It is necessary to induce the phase separation of PAH without electrical energy input to achieve active control of smart windows.
Salt can be used to trigger PAH hydrogel-based smart windows.22 Since the phase separation of PAH is dominated by ionic bonds and dense entanglement to scatter natural light (Fig. 10), when PAH is immersed in sodium chloride solution [NaCl(aq)], the ionic bonds in PAH are disassociated. Thus, the polymer chain associates are partially ruptured, and the entanglement density decreases to pass through incident light, leading to a Tlum of 92%. Subsequently, the PAH hydrogel automatically recovers its opaqueness within 20 minutes when transferred from NaCl(aq) to H2O. The reversible transparency-shifting endows the PAH hydrogel with a ΔTsol of 88%.
![]() | ||
Fig. 10 Pictures of PAH hydrogel in opaque and transparent states when immersed in H2O and NaCl(aq), respectively, and the proposed dynamic nature of ionic bonds. Reproduced with permission from Guo et al., Mater. Horizons 9, 3039 (2022). Copyright 2022 Royal Society of Chemistry.22 |
Thermochromic properties and characteristics of these thermosensitive hydrogels are shown in Table 2. The performances of hydrogels and related smart windows can be flexibly modulated over a broad range, including adjustable τc (16.0–43.5 °C), strong solar modulation ability (33.4–89.9%), good mechanical strength (89.5 kPa–5.58 MPa tensile strength), and anti-freezing capability (−18 to −100 °C freezing point). Unlike metal oxides, hydrogels are soft materials with excellent processibility and adaptiveness to match specific application demands, even at temperature far below the freezing point.
Materials | Characteristic | T lum (%) | ΔTsol (%) | τ c (°C) | Ref. |
---|---|---|---|---|---|
PNIPAM | Decent visible transmittance | 70.7 | 25.5 | 32.0 | 100 |
PNIPAM/AEMA | Low-cost and scalable fabrication process | 87.2 | 81.3 | 32.0 | 16 |
PNIPAM/HPMC | Good solar modulation capability | 90.8 | 81.5 | 32.0 | 97 |
HPC | First smart window based on HPC hydrogel | 67.4 | 25.7 | 30.0 | 102 |
HPC/PAA | Reduced LCST by modulating pH | 90.1 | 47.5 | 26.5 | 106 |
HPC/PAA | Reduced LCST by modulating pH, and outstanding cycling stability | 91.8 | 73.7 | 25.8 | 107 |
HPC/PAA | Reduced LCST by modulating pH | 90.1 | 80.0 | 35.2 | 119 |
HPC/PAA | Electric-/thermal-dual-responsive | 89.0 | 59.0 | — | 108 |
PAAM/SDS | Adjustable LCST in the range of 16–34 °C | 85.0 | — | 16.0 | 120 |
PAAM/C18/SDS | Adjustable LCST in the range of 20–50 °C | 99.0 | 33.4 | 20.0 | 121 |
PNIPAM/NaCl | Adjustable LCST in the range of 31.7–24.5 °C | 95.1 | 89.9 | 27.2 | 122 |
PNIPAM/DMMA | Adjustable LCST in the range of 32.5–43.5 °C | 91.3 | 88.8 | 32.5 | 23 |
PNIPAM/Ethanol | Enhanced solar modulation ability and solar shielding | 89.9 | 71.8 | 28.0 | 123 |
PNIPAM/HA | Enhanced visible transmittance and solar modulation ability | 85.9 | 68.6 | 32.0 | 114 |
W-VO2/PNIPAM/HPC | Enhanced solar modulation ability | 87.2 | 65.7 | 29 | 124 |
PNIPAM particles/water | Revolutionarily liquid smart window for enhancing optical properties | 90 | 68.1 | 32.5 | 21 |
PNIPAM/PAA chain/microparticle hybrids suspension | Liquid smart window for enhancing solar modulation ability | 91.5 | 85.8 | 26.7 | 112 |
SPU/binary ionic liquid | Enhanced mechanical strength by hydrogen bonds | 95.1 | 83.0 | — | 115 |
PNIPAM/ACMO | Enhanced mechanical strength by 3D printing technology | 85.8 | 79.3 | — | 99 |
PNIPAM/PDMAA/EG | Excellent anti-freezing ability and adjustable LCST | 89.3 | 80.7 | ∼30.0 | 125 |
PNIPAM/glycerol | Good anti-freezing ability and adjustable LCST | 90.0 | 60.8 | ∼26.0 | 118 |
PNIPAM/HPC | Good anti-freezing ability | 80.7 | 64.5 | ∼29.0 | 95 |
PNIPAM microgel | High luminous transmittance and fast response time | 91.2 | 81.8 | — | 126 |
HPC/KCl/polyaniline | Thermo/electrochromic dual response | 73.0 | 57.5 | 30.0 | 127 |
PNIPAM/upconverting nanoparticles | Converting infrared light into visible light | 82.8 | 79.8 | 40.0 | 128 |
PNIPAM/Si/Al | Printable hydrogel | 80.1 | 72.7 | 35.0 | 129 |
Inorganic salts are introduced into hydrogel networks to regulate phase transition temperature of thermochromic hydrogels. A thermochromic hydrogel-based smart window with good optical properties (Tlum of 99.1% and a ΔTsol of 33.4%) is developed by copolymerizing hydrophilic acrylamide (AAM), hydrophobic stearyl methacrylate (C18), and sodium dodecyl sulfate (SDS) micelles.121 SDS micelles serve as surfactants to modulate the self-assembled particle size of the hydrophobic C18 phase that is covalently linked to PAAM chains, which influence the light transmittance and the phase transition temperature. The metal ions interact with the SDS headgroup (–SO4−) and reduce the electrostatic repulsion, leading to a lower critical micelle concentration (CMC) of SDS. The salt concentration and cation radius also affect the CMC of SDS. Thus, the thermochromic hydrogel presents a tunable phase transition temperature from 20 °C to 50 °C. Similarly, Tian et al. developed a PNIPAM hydrogel-based smart window with optical properties (Tlum of 92.7% and a ΔTsol of 93.6%).122 Adjustable phase transition temperature within 24.5–31.7 °C is achieved by changing the sodium chloride (NaCl) concentration. Na+ interacts with the polymer surface and weakens the hydrogen bonds between amide groups of polymer chains and water. Thus, the phase separation of PNIPAM is more susceptible to the salt concentration, resulting in a tunable τc.
Ding et al. constructed a PNIPAM hydrogel-based smart window through introducing ethanol to enhance solar modulation to 71.8%.123 Ethanol decreases the size of scattering centers and increases the scattering center number. Meanwhile, it weakens interactions between water and polymer chains and discharges more water from the PNIPAM network during the phase transition, resulting in a high refractive index difference between PNIPAM and water. Thus, the hydrogel has highlight scattering behavior and presents an enhanced solar modulation ability.
W-doped VO2 nanoparticles are introduced into PNIPAM/HPC hydrogels to enhance ΔTsol.124 W-doped VO2 nanoparticles are attached to the surface and interior of PNIPAM/HPC hydrogels. As a result, the hydrogel presents a thicker pore wall and higher porous density, which scatters light above its τc, leading to a higher solar modulation ability of 65.7%.
A liquid hydrogel-based smart window is recently developed by copolymerizing N-isopropylacrylamide (NIPAM) and acrylic acid (AA).112 The liquid hydrogel includes linear PNIPAM-PAA chains and Cu3Cit2@PNIPAM-PAA core–shell microgel particles. Above τc, the chains are dissolved and aggregate to form new compact microgel core–shell particles, which undergo a phase transition and shrink. The optical contrast between microgel particles and water results in strong light scattering performance, leading to an enhanced solar modulation ability. The synergetic effect of polymer chains-microgel particle conversion and core–shell particle shrinkage strengthens ΔTsol to 85.5%.
Dai et al. developed a physically crosslinked ionogel by photopolymerizing hydroxyethyl acrylate (HEA) and butyl acrylate (BA) within ionic liquid [BMIm]+[BF4]− (Fig. 11a).117 HEA and BA show excellent solubility in [BMIm]+[BF4]−. The ionogel with three-dimensional porous structures has high-density molecular entanglements and cross-linking. The maximum fracture strain increases with the [BMIm]+[BF4]− content. Thus, the ionogel presents a tensile strength of 89.5 kPa at 670% fracture strain.
![]() | ||
Fig. 11 (a) Schematic diagram of the ionogel network structure. Reproduced with permission from Dai et al., ACS Appl. Polym. Mater. 5, 3398 (2023). Copyright 2023 American Chemical Society.117 (b) Schematics of the PNIPAM-PAA/PAAM hydrogel with a crack under tension. Reproduced with permission from Zheng et al., Compos. Commun. 42, 101684 (2023). Copyright 2023 Elsevier, Inc.138 (c) Impact resistance testing of ordinary glass, photochromic glass, armored glass, and smart glass. Reproduced with permission from Chen et al., Adv. Funct. Mater. 33, 2214382 (2023). Copyright 2023 Wiley-VCH GmbH.116 |
Zheng et al. reported a double-network hydrogel by incorporating PNIPAM-PAA microgels into a PAAM matrix for thermochromic smart windows, forming interpenetrating networks.138 PNIPAM-PAA microgels prevent crack propagation, and the PAAM matrix serves as a sacrificial network to dissipate energy (Fig. 11b). As a result, the hydrogel with only 2 wt% microgels shows a fracture strength of around 70 kPa and breaking elongation of 420%.
An alcogel with supramolecular configuration is constructed to improve mechanical properties of smart windows.116 The alcogel with 68.7 wt% ethanol content has host-cellulose chains wrapping guest-polyacrylamide (PAAM) chains through extensive hydrogen bonds. The alcogel shows compact supramolecular structures between cellulose and PAAM and ethanol-induced curling of PAAM chains. It shows a tensile strength of 13.6 MPa and a higher impact resistance of 42.8 kJ m−2 than ordinary glass (4.1 kJ m−2) (Fig. 11c).
Li et al. fabricated a poly(propylene glycol) (PPG)-based ionogel by synthesizing hydroxyl-terminated PPG, isophorone diisocyanate (IPDI), and adipic acid dihydrazide (ADH) for thermochromic smart windows.115 The hydrazide groups of ADH react with isocyanate groups to produce acylsemicarbazide (ASCZ) moieties. Meanwhile, multiple hydrogen bonds form between the ASCZ moieties, which increase the crosslinking density of the polymer network. Thus, the ionogel presents a high tensile strength of 5.58 MPa and a tensile strain of 3210%.
Wang et al. prepared an anti-freezing smart window with a Tlum of 89.3% and ΔTsol of 80.7% by using thermochromic poly(N-isopropylacrylamide)-poly(N,N-dimethylacrylamide)/ethylene glycol hydrogel (PNIPAM-PDMAA/EG) that is resistant to low temperatures.125 The introduction of EG is conducive to the formation more hydrogen bonds between polymer chains and water, which effectively hampers the freezing of the hydrogel. The hydrogel does not freeze at temperatures down to −100 °C. Similarly, an anti-freezing smart window with a Tlum of 90% and ΔTsol of 60.8% is fabricated based on the noncovalent crosslinked PNIPAM hydrogel.118 The PNIPAM hydrogel with a freezing point of around −18 °C is synthesized in a glycerol–water (GW) binary solvent system. The hydrogen bonds between glycerol and water are more stable than that between water and water, thereby resulting in excellent low-temperature resistance.
Feng et al. developed an anti-freezing hydrogel for smart windows with a Tlum of 80.7% and ΔTsol of 64.5% by simply entangling PNIPAM and HPC together without adding other organic solvents.95 The HPC fibers are liable to form hydrogen bonds with water, efficiently restricting water crystallization. As a result, the PNIPAM/HPC hydrogel has a good low-temperature resistant ability at −13 °C.
Dai et al. developed a thermochromic hydrogel with a reversible three-stage transition of opaque-transparent-translucent by introducing SDS and NaCl into the cross-linking network of PNIPAM-PAAM.140 The solubility of SDS increases as the temperature increases and shows a clear transition point (Krafft point, Tk). At the temperature below Tk, SDS molecules separate from the solvent and precipitate inside the hydrogel network, making the hydrogel opaque. At the temperature between Tk and τc, SDS micelles are formed and the hydrogel becomes transparent, which results in a Tlum of 80.3%. At temperature above τc, PNIPAM chains aggregate and form phase separated microdomains to scatter incident light. Thus, the hydrogel becomes opaque to inhibit indoor heat dissipation through the window to the outside at low temperature and demonstrates a ΔTsol of 72.9%. Meanwhile, it becomes translucent to scatter sunlight at high temperature, showing a ΔTsol of 42.7%. A model styrofoam house with a hydrogel smart window is used for the solar energy shielding test. The model house with normal glass is set as a control. The indoor temperature of the model house with a smart window is reduced by 3.6 °C after 30 minutes of infrared irradiation (55 °C), and the interior temperature of the model house with normal glass is 1.3 °C lower than that of the model house with a smart window after being placed in a cold closet for 40 minutes.
Recently, our group developed thermochromic hydrogel smart windows for all climate applications to modulate solar radiation and heat radiation.23 The smart windows are equipped with LCST type thermoresponsive hydrogels with different τc. Transparent thermochromic hydrogels are prepared by copolymerizing hydrophilic N,N-dimethylacrylamide (DMAA) and NIPAM monomers at room temperature with UV initiation. The introduction of a hydrophilic DMAA monomer enhances hydrophilicity and then slightly suppresses phase separation at high temperatures. As a result, the LCST or τc of hydrogels gradually increases from 32.5 °C to 43.5 °C with increasing DMAA content in the network. Here, UV-initiated free radical polymerization at low temperature is used to avoid potential phase separation caused by reaction heat. The obtained hydrogels are highly transparent, with a luminous transmittance of 91.3%. In August, in Guangzhou, as an example, after 1 h of sunlight exposure at 14:00 p.m., the hydrogel smart window becomes opaque (Fig. 12a).
![]() | ||
Fig. 12 Smart windows for all-climate applications. (a) Photos of a hydrogel smart window in the sun in Guangzhou. (b) Schematic illustration of a two-compartment simulation test setup. (c) Temperature distribution in the setup in the sun in Guangzhou, left: control and right: test group. (d) Field testing of hydrogel smart windows in Beijing, Xi'an, Shanghai, Fuzhou, Guangzhou and Dalian. Monthly temperature regulation and energy saving of smart windows in (e) Guangzhou and (f) Dalian, and temperature differences between test and control rooms, where T0, T1, and T2 are temperatures in the shadow of the test group and the control group. (g) Thermal infrared images of hydrogel smart windows with different grid patterns. Reproduced with permission from Chen et al., Adv. Mater. 35, 2211716 (2023). Copyright 2023 Wiley-VCH GmbH.23 |
Two-compartment thermal insulation boxes are assembled to test the solar regulation ability of the smart windows in various climates. The test room is equipped with a thermochromic hydrogel smart window, and the control room is equipped with an ordinary double layer glass window (Fig. 12b). Here, the smart window comprises a thermochromic hydrogel sandwiched between two glasses. Since the refractive constant of the hydrogel is close to that of glass, the transmittance of smart windows is augmented from 79.4% to 85.5%. At high temperatures, the smart windows become opaque and prevent local heating. Therefore, the temperature of the test room is more uniform than that of the control room (Fig. 12c). The temperature increase of the test room (40.6–44.0 °C) is less than that of the control room (45.8–51.6 °C).
The smart windows show outstanding solar and heat modulation in different climates. Simulation tests are carried out in six cities (Beijing, Dalian, Xi'an, Shanghai, Fuzhou and Guangzhou) with different geographical locations and climates from December 2021 to August 2022 (Fig. 12d). In Guangzhou, a tropical southern city, the temperature in direct sunlight areas in the summer can reach 50–60 °C. The hydrogel smart window with a τc of 40.3 °C becomes opaque and scatters sunlight, showing a ΔTsol of 88.8%. As a result, the air temperature in the test room (T1) is lower than that in the control room (T2). The indoor temperature is reduced by 4 °C in March, and an energy saving of 5.14 kJ m−3 is achieved (Fig. 12e). In December, in Dalian, the northern city, the ambient temperature in winter is below the freezing point and central heating temperature is used. As the room temperature is raised to 35 °C as a proof-of-concept test, the transparent hydrogel smart window with a τc of 33 °C becomes opaque and prevents indoor heat dissipation through the window to the outside. The temperature outside the test room (T1) is 4.7 °C lower than that of the control room (T2), and an energy saving of 6.05 kJ m−3 is achieved (Fig. 12f). These results demonstrate the capability of such smart windows to keep rooms “warm in winter and cool in summer”.
On the other hand, the energy saving performance of smart windows is investigated by theoretical calculations based on theoretical models.141 For example, Long et al. reported a thermochromic smart window using a HPC entangled PNIPAM hydrogel with a fiber structure.95 At 20 °C, the fibers are thin and elongated to allow light to pass through, leading to a Tlum of 80.7%. In contrast, the fibers aggregate to scatter light at 40 °C, and thus the smart window shows a ΔTsol of 64.5%. The energy consumption is calculated by using an actual-size house model (8 m in length, 6 m in width, and 2.7 m in height) with four windows (3 m in width and 2 m in height) and the weather conditions in Singapore. The calculation results show an energy saving of ∼6 MJ m−2 in July. Moreover, the same group estimated the energy saving of another thermochromic smart window with a Tlum of 90% and ΔTsol of 68.1% based on PNIPAM particle dispersion in water.21 Such a smart window combines solar heat absorption by water with large specific heat capacity (4.2 kJ kg−1 K−1) and the solar modulation of PNIPAM particles at 32 °C, which results in an energy saving of ∼7 MJ m−2 in July. These results indicate that thermochromic smart windows are promising for building energy saving.
Such thermochromic hydrogel smart windows can be processed into well-designed structures to balance visibility and solar regulation. Gridded thermochromic hydrogel smart windows are prepared by 3D printing. Upon heating, the hydrogel grids become opaque, while the rest of the domains remain transparent. By adjusting the mesh size and density, both good visibility and solar modulation ability (59.6%) are achieved (Fig. 12g).
![]() | ||
Fig. 13 (a) The temperature-dependent crystal structure change of ionic liquids. Reproduced with permission from Zhu et al., ACS Appl. Mater. Interfaces 8, 29742 (2016), Copyright 2016 American Chemical Society.147 (b) Schematic illustration of the change process of perovskite materials: with increasing temperature, the perovskite crystals nucleate and grow. Reproduced with permission from Ke et al., Adv. Funct. Mater. 28, 1800113 (2018). Copyright 2018 Wiley-VCH GmbH.12 (c) Schematic diagram of the structures in a hybrid perovskite compound. Reproduced with permission from Teri et al., J. Mater. Chem. C 11, 8903 (2023). Copyright 2023 Royal Society of Chemistry.160 (d) Schematic illustration of the phase transition between the aligned phase and focal conic phase. Reproduced with permission from Shen et al., Laser Photonics Rev. 17, 2200207 (2023). Copyright 2023 Wiley-VCH GmbH.167 |
Ionic liquids can be used to prepare thermochromic ionogels for use in smart windows. For example, Wu et al. reported a thermochromic ionic liquid-based gel electrolyte with a Tlum of 90.2% and ΔTsol of 82.3% for thermos-/electro-dual-responsive smart windows.148 With the addition of an ionic liquid, hydrogen bonds form between the polymer network and ionic liquid, which result in better swelling and movement of polymer segments, leading to an adjustable τc (47–55 °C).
Combining thermochromic ionic liquid-based gels with other mechanisms can impart multiple adjustabilities to smart windows. A thermochromic ionic liquid 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([EMIM]TFSI) microemulsion-based gel is prepared by polymerizing an acrylamide monomer in an ionic liquid-based microemulsion.149 Tween 20 is distributed at the oil–water interface to stabilize the droplets in the ionic liquid microemulsion. At 25 °C, the droplets are stable and difficult to precipitate. Thus, the thermochromic gel shows a Tlum of 99.3%. When temperature is above 40 °C, the microemulsion droplets aggregate, which highly scatter the incident light, leading to a high ΔTsol of 92.6%. Besides, [EMIM]TFSI endows the thermochromic gel with ionic conductivity. Thus, the gel is also utilized as an electrolyte layer for electrochromic smart windows. Integrating the conductive gel with a WO3 electrochromic layer achieves electric-/thermal-dual modulation of smart windows. Similarly, Deng et al. constructed a thermochromic ionic liquid 1-butyl-3-methylimidazolium tetrafluoroborate (BMIMBF4)-based gel electrolyte by polymerizing N-isopropylacrylamide.150 At temperature below LCST, the PNIPAM chains are dissolved in BMIMBF4 and water. When temperature is above LCST, the polymer chains aggregate into globules to scatter incident light, leading to a good thermochromic performance. Moreover, BMIMBF4 endows the gel with ionic conductivity up to 15.4 mS cm−1. Combining the thermochromic PNIPAM layer with an electrochromic TiO2 layer, an electric-/thermal-dual smart window is developed.
The thermochromic mechanism of perovskites is based on its abnormal crystallization phenomenon that its solubility gradually decreases with an increase in temperature. At low temperature, the solubility of thermochromic perovskites is high, and the hydrated perovskite is colorless to allow sunlight to pass through. As the temperature increases, the solubility of the perovskite decreases, and the colorless perovskite hydrate phase transforms into a colored photovoltaic perovskite phase to block sunlight (Fig. 13b).12,159
Thermochromic perovskites show a red shift in absorbance with increasing temperature. Teri et al. developed a copper-based hybrid perovskite (MBA)2CuCl4 (MBA is 4-methoxybenzylamine) with a reversible phase transition temperature of 430.2 K.160 The dielectric states are different because the organic aminium cations transition between partial order and disorder states (Fig. 13c). The coordinate geometry of Cu2+ is relevant with the position of the absorption maximum for the d–d transition. The structure of Cu-based perovskite changes from tetrahedral coordination to an octahedron with increasing temperature, which causes a shift in the electronic absorption band. Thus, the red shift of hydride perovskite occurs in the absorbance spectrum, resulting in a color change from yellow to brown.
One challenge of perovskite-based smart windows is how to reduce the high critical transition temperature to around ambient temperature. Liu et al. fabricated a novel smart window based on the thermochromic perovskite H-MAPbI3−Clx with a ΔTsol of 23.7%.161 The hydration and dehydration processes of thermochromic perovskite H-MAPbI3−xClx during the phase transition demonstrate that the ambient humidity is important to the τc. When the relative humidity decreases to 50%, the τc is around 40 °C, which is lower than that of 80% relative humidity (51.4 °C). The τc can be reduced from 51.4 °C to 29.4 °C with a decrease in relative humidity (from 80% to 25%).
Developing alternative strategies to trigger the thermochromic effect of perovskite-based smart windows at ambient temperatures is necessary for applications. Liu et al. proposed a NIR activated perovskite-based smart window at room temperature.154 The smart window is fabricated by integrating cesium-doped tungsten trioxide (CWO) with T-perovskite H-MAPbI3−xClx. CWO has high visible transmittance of 82% and strong NIR absorption of 90%. As a result, the perovskite-based smart window shows a Tlum of 65.7%. At 23.5 °C, the strong NIR absorption of the CWO layer produces thermal energy and heats the window to the phase transition temperature of thermochromic perovskite. Then the thermochromic T-perovskite dissociates water from the MAPbI3−xClx layer. Thus, the smart window changes from a transparent state to a colored state and shows a ΔTsol of 17.5%.
Liquid crystal-based thermochromic smart windows have excellent solar modulation capability because their strong control in both visible and near-infrared regions. For example, Liang et al. developed a thermochromic liquid crystal-based film with a Tlum of 67%.172 The composite film is composed of liquid crystal, poly(vinylpyrrolidone) (PVP), and tungsten bronze (CsxWO3) nanorods. The phase of liquid crystal changes from a smectic state to a chiral nematic state with an increase in temperature, which provides a change from transparency to opaqueness. CsxWO3 has strong and broadband LSPR absorption of NIR irradiation. Thus, the liquid crystal-based smart window shows an ultrahigh near-infrared irradiation shielding of 95% from 800 nm to 2500 nm.
Combining the thermochromic mechanism of liquid crystals with other mechanisms can give the smart windows multiple adjustment capabilities. Meng et al. constructed a photothermal dual-driven liquid crystal-based smart window.176 When the ambient temperature is 26 °C, the liquid crystal presents a smectic phase, allowing incident light to pass through, resulting in a Tlum of 70%. When the ambient temperature is above 41 °C, the liquid crystal changes from the smectic phase to a cholesteric phase. Then the refractive index of the liquid crystal is changed in the thickness direction. Therefore, the liquid crystal-based smart window scatters light, leading to 50% reduction in transmittance. Moreover, a photothermal molecule, isobutyl-substituted diimmonium borate (IDI), is doped into the liquid crystal-based system. IDI has strong absorption in the NIR range, and the temperature of the liquid crystal-based system with IDI increases from 24 °C to about 31.7 °C under sunlight for 160 seconds. IDI can convert light irradiation to heat and accelerate the phase change of the liquid crystal, leading to a photothermal dual-driven smart window. Similarly, an electrical/thermal dual-control film with 57.8% Tlum and 34.6% ΔTsol is prepared by combining liquid crystal/polyacrylate with W-doped VO2.177 The composite materials are sandwiched between two plastic conductive substrates. At low environmental temperature and no voltage, the composite materials present a homeotropically aligned liquid-crystalline polymer network. Both the liquid crystal and polymer possess similar refractive indices so that the film appears transparent. When an electric field is applied, the liquid crystal domains are aligned parallel across the direction of the electric field, yielding an anisotropic distribution of refractive index values. The difference between the liquid crystal and polymer results in light scattering. Meanwhile, the thermochromic W-doped VO2 exhibits passively thermochromic behavior during the phase transition at 43.2 °C. Thus, the liquid crystal-based composite materials are promising for smart windows that can be activated by temperature changes and/or using an electric field.
![]() | ||
Fig. 14 (a) Schematic for multi-directional heat conduction in the phase change composite. Reproduced with permission from Zhang et al., Adv. Funct. Mater. 32, 2109255 (2022). Copyright 2022 Wiley-VCH GmbH.179 (b) Schematic illustration of the proposed power-generating smart window architecture and working process. Reproduced with permission from Zhang et al., Adv. Energy Mater. 11, 2101213 (2021). Copyright 2021 Wiley-VCH GmbH.28 |
Although the solar thermal-electric conversion ability has been improved, it is still rare to design thermal-electric conversion smart windows due to high operation temperature and low visible transmittance. Zhang et al. constructed a Cs0.33WO3/resin/glass smart window with a high luminous transmittance of 88% and outstanding solar absorption (Fig. 14b).28 The absorbed near-infrared light and ultraviolet light can be effectively converted into heat. The thermoelectric system attached to the edges of the film collects the heat and transforms it into electricity. The smart window shows an output voltage of around 4 V. Besides, under a sunlight intensity of 100 mW cm−2, the open-circuit voltage and maximum power density reach up to 3.8 V and 0.4 mW cm−2 respectively. This work not only realizes efficient energy saving and power generation in the field of smart windows, but also provides significant insights and promising pathways for building energy saving.
![]() | ||
Fig. 15 (a) Illustration of the photovoltaic smart window work on cold and warm days. Reproduced with permission from Meng et al., Nano Energy 91, 106632 (2022). Copyright 2022 Elsevier, Inc.190 (b) The schematic diagram of the energy saving and energy storage integrated smart window in different environments. Reproduced with permission from Niu et al., Adv. Sci. 9, 2105184 (2022). Copyright 2022 Wiley-VCH GmbH.191 |
Thermochromic hydrogels have also been widely combined with solar cells to achieve both energy saving and energy generation. For example, Niu et al. constructed a multi-layer louver smart window containing a thermochromic host–guest hydrogel (HGT hydrogel) and Si-based solar cell (Fig. 15b).191 The HGT hydrogel is prepared by introducing HPC microparticles into the transparent PAM-PAA hydrogel matrix. Solar light can pass through the smart window at low temperatures, but the incident light is blocked due to the microstructure change of the HGT hydrogel at high temperatures. Therefore, the louver smart window has an ultrahigh luminous transmittance of around 90% and a strong solar modulation capability of about 54%. The unique louver structure of solar cell endows the smart window with an outstanding energy generation capability of 18.2%. Besides, the louver smart window exhibits a lower temperature of 13.3 °C than the normal one after 1 h radiation of simulated sunlight, showing good energy-saving performance. Similarly, Meng et al. developed a photovoltaic smart window by integrating a thermochromic HPC hydrogel with a perovskite solar cell.192 Because of the structural transition of the HPC hydrogel, the photovoltaic smart window has a luminous transmittance of 27.3% at 20 °C and 10.4% at above 40 °C, as well as a solar modulation ability of 15.7%. Indium tin oxide is used as the electrode of the solar cell. The smart window presents a high peak conversion efficiency of 17.5%. An excellent combination of solar modulation and power generation is achieved in this smart window, contributing to sustainable building energy saving.
Despite many excellent properties of thermochromic smart windows, there are still many challenges and unmet demands that stimulate further studies in this field. More novel thermoresponsive materials should be explored to satisfy the expansive application range of smart windows. Besides, the performances of current thermochromic materials are unsatisfactory and need to be improved. For example, the undesired color, biotoxicity and instability of VO2 severely restrict its development. And it is difficult to achieve a perfect balance of τc, ΔTsol and Tlum in thermochromic VO2. Thermochromic hydrogels have an outstanding solar modulation ability in smart windows based on their reversible hydrophilic/hydrophobic phase transition. However, the problems of solvent volatilization, freezing and mechanical weakness of hydrogels cannot be ignored. Then the high cost and intricate fabrication process of ionic liquids, perovskites and liquid crystals also need to be solved. Additionally, structural modification that elevates thermochromic performances have been successfully realized in VO2 and hydrogel systems, and different thermosensitive materials typically possess different modulation ranges of solar light. Thus, more efforts should be devoted to structure modification and composite application of thermochromic materials.
Thermoelectricity conversion based on the Seebeck effect for thermochromic smart windows is rarely investigated. More investigations are needed to develop new materials and mechanisms to promote the development of building energy generation. For solar cell integration, the low light transmittance and undesired color state should be addressed. Solar cells in smart windows have a lower power conversion efficiency than traditional perovskite cells due to the limitation of the phase transition of thermochromic materials; hence more fundamental research on the photovoltaic properties of solar cells should be conducted to enhance the electricity generation of thermochromic smart windows.
There are still some challenges with thermochromic smart windows to be addressed to meet commercial use. Common thermochromic smart windows are fabricated by sandwiching thermosensitive materials between two pieces of normal glass. The additional expenditure for thermosensitive materials and encapsulation is huge. Besides, the thermochromic properties of thermosensitive materials weaken with time, and therefore smart windows need to be regularly maintained and repaired. The long-term service life of thermochromic smart windows is important in practical applications of building. To this end, thermosensitive materials should have stable structures and optical properties. Traditional thermochromic materials show slow response times due to the tardy phase transition process caused by outdoor temperature changes. Thus, photo-thermal nanoparticles or additional applied voltage should be introduced to the thermochromic intelligent system to accelerate the response process. To meet the practical requirements of active control or privacy protection, thermal response smart windows should be combined with other stimuli to realize dual-response or even multi-response smart windows. Moreover, the promising trend is to integrate other functions into smart windows, such as self-cleaning, water harvesting, energy storage, super hydrophobicity, anti-drying, anti-freezing, etc. On the one hand, it is necessary to optimize the solar modulation of smart windows while maintaining good luminous transmittance. On the other hand, combining smart windows with solar cells to collect solar radiation and convert it into electricity will further enhance building energy saving.
In the future, smart windows will become a meaningful platform. Artificial intelligence and 5G society emerge one after another with increasing sophistication, so that smart windows may be more intelligent with other amazing features to serve people, for example, danger warning, sound insulation, intelligent ventilation, film projection, weather forecasts, and so on. We expect that more research communities will devote efforts together to promote the development of smart windows because the smart window system is a highly interdisciplinary technique. Throughout future developments, thermochromic smart windows will face tremendous tasks and challenges, but in parallel, they will also offer the exciting prospect of building energy saving.
Footnote |
† These two authors contributed equally. |
This journal is © The Royal Society of Chemistry 2024 |