Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Self-adaptive radiative cooling based on phase change materials

Open Access Open Access

Abstract

With the ability of harvesting the coldness of universe as a thermodynamic resource, radiative cooling technology is important for a broad range of applications such as passive building cooling, refrigeration, and renewable energy harvesting. However, all existing radiative cooling technologies utilize static structures, which lack the ability of self-adaptive tuning based on demand. Here we present the concept of self-adaptive radiative cooling based on phase change materials such as vanadium dioxide. We design a photonic structure that can adaptively turn ‘on’ and ‘off’ radiative cooling, depending the ambient temperature, without any extra energy input for switching. Our results here lead to new functionalities of radiative cooling and can potentially be used in a wide range of applications for the thermal managements of buildings, vehicles and textiles.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The universe, at a temperature of 3 K, represents the ultimate heat sink both in terms of its temperature and its capacity. Moreover, the earth atmosphere is highly transparent in the wavelength range of 8 – 13 μm. This transparency window coincides with the peak of blackbody radiation spectrum at typical ambient temperature near 300 K. Thus, any objects on earth, given sky access, can radiate heat out to the universe and thus lower its temperature. Such a process is known as radiative cooling.

Nighttime cooling effect has been known for centuries [1–3]. Recently, there has been a significant effort in pursuing daytime radiative cooling, by constructing systems that reflect the entire solar spectrum, while generate strong thermal radiation in the wavelength range of 8 – 13 μm [4]. Raman et al have demonstrated a photonic radiative cooler made of a multi-layer photonic structure, with an equilibrium temperature that is 5 °C below the ambient air temperature, in spite of having about 900 W/m2 of sunlight directly impinging upon it [5]. Subsequently, daytime radiative cooling have been considered in various material systems and structures [6–12] and has motivated further studies of radiative cooling in other scenarios [13–18]. At a hot temperature above thermal comfort, for example daytime in summer, radiative cooling technology can be used to efficiently dissipate heat of buildings, vehicles and human body, therefore significantly reduce the energy consumption and improve thermal comfort.

Despite the successes, so far, all the existing radiative cooling systems are static in the sense that the thermal emissivity of the systems is fixed once they are constructed. However, as temperature varies across days and seasons, when the environment temperature is below a critical temperature such that cooling is no longer desired, for example during nighttime in winter, radiative cooling may no longer be desirable and may even increase the energy consumption if heating is required. Therefore, for many applications, it would be highly desirable to develop a self-adaptive radiative cooling system, that can enable cooling when the environment temperature is above the critical temperature, and disable cooling when it’s below.

In this paper, we propose a self-adaptive radiative cooling system. Such a system can automatically turn on and off radiative cooling, depending on temperature, without any extra energy input for switching. Our approach is based on a planar photonic multilayer system incorporating phase change materials such as vanadium dioxide (VO2) and a spectrally-selective filter. We numerically show that such a system can perform radiative cooling, when the system is above the transition temperature of VO2, and automatically turn off radiative cooling when the temperature is below the transition temperature. Our results here explore new functionalities of radiative cooling applications and could be potentially used in a wide range of applications such as vehicles, buildings, textiles and smart thermal regulations.

2. Self-adaptive Radiative Cooling

We start by presenting the general concept of self-adaptive radiative cooling and design considerations. The idea is to design a system that is able to switch on and off radiative cooling, depending on the temperature of the environment. We aim to design a system that provides radiative cooling when the environment is hot, and turns off radiative cooling when the temperature is below a critical temperature [Fig. 1(a)].

 figure: Fig. 1

Fig. 1 (a) Schematic showing the concept of self-adaptive radiative cooling: when temperature is above the critical temperature Tc, radiative cooling is turned on; when temperature is below the critical temperature Tc, radiative cooling is turned off. (b) Schematic of the ideal spectrum for self-adaptive radiative cooling. The spectrum switches between an ‘on’ and an ‘off’ state depending on the ambient temperature.

Download Full Size | PDF

In order to achieve such functionalities, distinct spectral features for ‘on’ and ‘off’ states are required. When the system is in the ‘on’ state, the system needs to fulfil the requirement of daytime radiative cooling, as outlined in [4]. The system needs to maximize emissivity in the atmosphere transparency window in the wavelength range of 8 – 13 μm to efficiently emit heat to the outer space [Fig. 1(b)], while minimize absorption from both the sun light and the thermal radiation from the atmosphere by suppressing the emissivity for the rest of the wavelength range. On the other hand, when the system temperature is below a critical temperature, the system needs to be in the ‘off’ state, where it should exhibit minimum thermal emissivity in the entire thermal wavelength range to turn off radiative cooling process.

To realize such a self-adaptive emissivity switch, we use phase change material as the switching element. Compare to other thermal emission switching techniques [19] such as electrical carrier injection [20,21] or mechanical switch [22,23], using phase change materials as switching element, where the phase change is thermally induced, does not involve any power input or feedback control, greatly reducing the system complexity [24–30]. As an illustration, we choose VO2 as the phase change material. VO2 has metal-insulator phase transition, exhibiting metallic or insulating state when temperature is above or below the phase transition temperature, respectively. In the infrared wavelength range, optical properties of the two states show distinct features, the insulating state is a low-loss dielectric, whereas the metallic state is a plasmonic metal with a high damping constant as shown in Fig. 2 [31,32]. Such phase transition has been previously employed in demonstrating unique thermal emission properties such as negative differential thermal emittance [24], thermal homeostasis [25] and radiative thermal runaway [26].

 figure: Fig. 2

Fig. 2 The real part (a) and the imaginary part (b) of the dielectric constants for VO2 used in the paper in the wavelength range from visible to mid-infrared.

Download Full Size | PDF

Typically, the phase transition temperature of un-doped VO2 is around 68 °C [31,33]. However, it has been reported that co-doping of W and Sr to pure VO2 can significantly reduce phase transition temperature, even to below room temperature [34–36]. Therefore, for self-adaptive radiative cooling purpose, it is conceivable to use VO2 as the switching material. With proper doping engineering, one can control the phase transition temperature of VO2 to match the critical temperature for switching on and off radiative cooling.

3. Photonic Structure Design

We now introduce the photonic structure design that can fulfil the spectral requirement of self-adaptive radiative cooling. Our system consists of two components, with a switchable radiative cooler at the bottom, and a spectrally-selective filter on the top. The switchable radiative cooler consists of VO2/MgF2/W with VO2 serves as the switching component [Fig. 3(a)]. In the following calculation, the optical constants for VO2 shown in Fig. 2, is determined from [31,32]. And the optical constants for other materials are taken from [37].

 figure: Fig. 3

Fig. 3 Photonic structures for realizing self-adaptive radiative cooling. (a) A schematic for the bottom radiative cooler, which is made of three layer VO2/MgF2/W structure. (b) Solar absorptivity of the radiative cooler. (c) Infrared emissivity of the radiative cooler. (d) A schematic showing the top spectrally-selective filter, which is made of 11 layers of Ge/MgF2. (e) Transmissivity of the filter in the solar wavelength range. (f) Transmissivity of the filter in the thermal wavelength range. (g) Schematic of the combined system. (h) Solar absorptivity of the radiative cooler in presence of the filter on top. (i) Infrared emissivity of the radiative cooler in presence of the filter on top.

Download Full Size | PDF

The absorptivity and emissivity of the radiative cooler in solar and thermal wavelength ranges are shown in Fig. 3(b) and Fig. 3(c), respectively. When VO2 is at metallic state, such a structure acts as a Salisbury screen absorber [38]. The resonant wavelength can be tuned by appropriately selecting the thickness of the MgF2 layer, according to the following equation:

λ4dΜnM
Here, dM and nM are the thickness and the refractive index of the MgF2 layer, respectively. With properly designed thickness of the MgF2 and VO2 layers, the bottom radiative cooler can exhibit strong and broadband infrared emissivity [Fig. 3(c)], with high absorptivity and therefore high emissivity in the wavelength range of 8 to 13 μm. Here, we have taken the dM to be 2.25 μm and the thickness of VO2 layer, dV, to be 10 nm.

On the other hand, at insulating state, such a structure exhibits a strongly suppressed emissivity in the wavelength range of 8 to 13 μm. Therefore, the radiative cooler, by itself, exhibits the desired spectral characteristics in the thermal wavelength range. On the other hand, in the solar wavelength range, due to lossy nature of VO2 at both metallic state and insulating state, and the presence of Fabry-Perot resonances, the structure exhibits significant absorption in both states, which are not desirable for radiative cooling purpose [Fig. 3(b)].

To further improve the performance, we place a spectrally-selective filter on top of the radiative cooler, separated with a large air gap. This spectrally-selective filter is a 11-layer stack structure consisting of Ge/MgF2 [Fig. 3(d)]. This component has two purposes. First, in solar wavelength range, this component serves as a sunshade, preventing the solar irradiation from reaching the bottom radiative cooler, due to the strong absorption of Ge in the solar wavelength range. As is shown in Fig. 3(e), this structure has very low transmission in the solar wavelength range. Second, in the thermal wavelength range, this component serves as a band pass filter with high transmissivity from 8 to 13 μm. Both Ge and MgF2 are low loss materials in infrared. By optimizing the thickness of each layer (the thicknesses of the layers are presented in Table 1), this structure allows highly selective transmission from 8 to 13 μm while strongly suppresses the emissivity of the bottom component for the rest of infrared wavelength range [Fig. 3(f)].

Tables Icon

Table 1. Material composition and thicknesses for the designed spectrally-selective filter.

We now present the performance of the radiative cooler in presence of the top filter [Fig. 3(g)]. The two components are separated with a gap with a width that is much larger than the wavelength. For such a structure, we can treat the optical properties of the combined structure with incoherent calculations [39]. In this treatment, absorptivity for the bottom radiative cooler in presence of the top filter is calculated by an incoherent summation of the contributions from the multiple reflection processes between the filter and the cooler as below:

ϵ(λ,Ω)=(1rC)(tF+tFrCrF+tF(rCrF)2+tF(rCrF)3+)= tF(1rC)/(1rCrF)
Here, tF, rC, and rF are power transmission coefficient for the filter, power reflection coefficient at the top surface of the radiative cooler, and power reflection coefficient at the bottom surface of the filter, respectively. These tF, rC, and rF are functions of wavelength and incident angle.

When the system is above critical temperature and VO2 is in the metallic state, in the presence of the top filter, the radiative cooler structure has a minimum solar absorption in the solar wavelength range [Fig. 3(h)]. Meanwhile, in the infrared wavelength range, we observe a strong selective emissivity from 8 to 13 μm and minimal absorptivity/emissivity for the rest of the thermal wavelength range, approaching the ideal spectrum required for radiative cooling [Fig. 3(i)]. On the other hand, when the system is below critical temperature and VO2 is in the insulating state, minimum emissivity is observed and radiative cooling is turned off. Furthermore, our system also shows robust emissivity against variation of incident angle as shown in Fig. 4(a). In the metallic state, the polarization-averaged emissivity remains high in the wavelength range from 8 to 13 μm even at large incident angles up to around 70 degrees, with only a slight blue shift at large incident angles. The angle and polarization averaged emissivity from 0 to 90 degrees in the wavelength range from 8 to 13 μm is 0.636 in the metallic state. On the contrary, this averaged emissivity drops to 0.054 in the insulating phase, leading to a switching ratio of 11.8 [Fig. 4(b)]. Overall, such a temperature-dependent switching property should enable a self-adaptive radiative cooling system.

 figure: Fig. 4

Fig. 4 (a) The emissivity of the bottom radiative cooler in presence of the top filter, as a function of incident angle and wavelength. The VO2 is in the metallic phase. (b) Angle and polarization averaged absorptivity spectrum of the bottom radiative cooler in presence of the top filter. Red and blue line shows the absorptivity in metallic and insulating VO2 phase, respectively.

Download Full Size | PDF

4. Thermal Performance

We proceed to evaluate the thermal performance of the self-adaptive radiative cooling system. We simulate the temperature of the radiative cooler by solving the heat balance equation:

Qtotal=Qcooler(T) Qatm(Tamb) Qparasitic(T,Tamb)Qsun(T) Qexchange(T,Tfilter)
Here, Qtotal is the net cooling power by the radiative cooler. T, Tamb, and Tfilter are the temperatures of the radiative cooler, ambient, and the filter, respectively. Qcooler stands for the emitting power from the bottom structure expressed by:
Qcooler(T)=AdΩcosθ0dλIBB(T,λ)ϵ(λ,Ω,T)
A is the area of the structure, and dΩ=2π0π2dθsinθ is the angular integral over a hemisphere with θ being the polar angle. IBB(T,λ)=2hc2λ51ehcλkBT1 is the spectral radiance of a blackbody at device temperature T, where h is the Planck’s constant, kB is the Boltzmann constant, c is the speed of light and λ is the wavelength. ϵ(λ,Ω,T) is the temperature-dependent emissivity of the radiative cooler in presence of the top filter. To obtain such emissivity, we assume that the phase transition of VO2 occurs at the temperature of 293K, and the dielectric function of VO2 smoothly interpolates between the two phases in a narrow temperature range around the phase transition temperature [25]. Absorbed power due to incident thermal radiation from atmosphere, Qatm, is
Qatm(Tamb)=AdΩcosθ0dλIBB(Tamb,λ)ϵ(λ,Ω,T)ϵatm(λ,Ω)
The absorptivity of the atmosphere, ϵatm(λ,Ω), is calculated by 1tatm(λ,Ω), where tatm(λ,Ω) is obtained from MODTRAN5 [40]. Parasitic power loss, Qparasitic, due to conduction and convection is
Qparasitic(T,Tamb)=Ah(TambT)
Here, h is the heat transfer coefficient, here we consider h = 8 Wm−2K−1 to mimic a typical natural air convection condition. Incident solar power absorbed by the structure, Qsun, is
Qsun(T)=A0dλIAM1.5(λ)ϵ(λ,θsun,T)
The radiative thermal exchange between the radiative cooler and the top filter, Qexchange, is
Qexchange(T,Tfilter)=AdΩcosθ{0dλIBB(Tfilter,λ)ϵfilter(λ,Ω)ϵcooler(λ,Ω,T)0dλIBB(T,λ)ϵfilter(λ,Ω)ϵcooler(λ,Ω,T)}
Here, ϵfilter(λ,Ω) and ϵcooler(λ,Ω,T) are emissivity of the filter and the radiative cooler, respectively.

Similarly, the temperature of the filter, Tfilter, is obtained by calculating net cooling power for the filter, Qtotal,F, as below.

Qtotal,F=Qfilter(Tfilter) Qatm,F(Tamb) Qparasitic,F(Tfilter,Tamb)Qsun,F+ Qexchange(T,Tfilter)
Here, Qfilter is the thermal radiation to the sky by the filter. Qatm,F, Qparasitic,F, and Qsun,F are absorbed thermal radiation, parasitic power loss, and absorbed solar power by the filter, respectively. The formulas for Qfilter, Qatm,F, Qparasitic,F, and Qsun,F are obtained by replacing the emissivity for the radiative cooler with that for the filter in the equations above for the cooler. In the calculation procedure discussed above, the equilibrium temperature for both the filter and the cooler are self-consistently determined by simultaneously solving for Eqs. (3) and (9) with both Qtotal and Qtotal,F set to be zero.

We can also obtain the time-dependent temperatures of the radiative cooler as well as the top filter by solving the differential equations:

CcoolerdTdt= Qtotal(T,Tamb,Tfilter)
CfilterdTfilterdt= Qtotal,F(T, Tamb,Tfilter)
To determine the heat capacitance of the radiative cooler, Ccooler, we assume that the radiative cooler, which consists of VO2 (10 nm) / MgF2 (2.25 μm) / W (5 μm) layers as mentioned in the previous section, is deposited onto a Si substrate with a thickness of 525 μm. The heat capacitance includes the contribution of the Si wafer. For the top filter, the heat capacitance Cfilter is determined from the sum of the heat capacitance of each layer listed in Table 1.

We now calculate the transient temperature response of the self-adaptive radiative cooling system under various ambient temperature conditions. For each ambient temperature, we assume that at the initial time both the cooler and the filter has a temperature that is the same as the ambient temperature. We then integrate Eqs. (10) and (11) to obtain the temperature of the filter and the cooler as a function of time, as shown in Fig. 5(a).

 figure: Fig. 5

Fig. 5 Transient thermal performance. (a) Time-dependent temperature evolution of the radiative cooler under different ambient temperature. The initial temperature is assumed to be the same as ambient temperature. The whole system is exposed to the sky at time = 0. (b) Time-dependent Qtotal, Qcooler, Qexchange, of the radiative cooler, and the temperature of the filter Tfilter, at Tamb = 313 K. (c) Time-dependent Qtotal, Qcooler, Qexchange, of the radiative cooler, and the temperature of the filter Tfilter, at Tamb = 298 K. (d). Time-dependent Qtotal, Qcooler, Qexchange, of the radiative cooler, and temperature of the filter, Tfilter at Tamb = 283K.

Download Full Size | PDF

When the ambient temperature is significantly above the critical temperature, for example 313 K [black curve in Fig. 5(a)], the system is in the ‘on’ state. As the time evolves, the radiative cooler temperature is reduced and eventually reaches an equilibrium temperature that is about 10 K below the ambient temperature, due to the effect of radiative cooling. At the equilibrium, the net outgoing power from the cooler approaches zero. [Red line, Fig. 5(b)]. The radiative cooling power of the radiative cooler, Qcooler, maintains over 100 W/m2 in this case [blue curve in Fig. 5(b)]. Although the heat exchange between the filter and the cooler, Qexchange, increases to ~18 W/m2 as the Tfilter rises to around 360 K due to solar absorption [Dashed line, Fig. 5(b)], the Qexchange [Green line, Fig. 5(b)] is not significant in comparison with the Qcooler in the cooling process.

When the ambient temperature is slightly above critical temperature [green curve in Fig. 5(a)], for example 298 K, the system is initially in the ‘on’ state. The cooler exhibits significant cooling power [Blue curve in Fig. 5(c)]. As time evolves, the radiative cooler temperature is reduced until it reaches the critical temperature, at which point the system is switched to the ‘off’ state, with a significant sudden reduction of the cooling power as shown by the blue curve in Fig. 5(c). After the phase transition, the radiative cooler temperature is maintained near critical temperature.

When the ambient temperature is below the critical temperature, for example at 283 K [pink curve in Fig. 5(a)], the system is in the ‘off’ state without radiative cooling. The radiative cooler has a temperature that is almost unchanged as a function of time. In this case, the cooling power Qcooler remains low around 12 W/m2 during the entire time period considered [blue curve in Fig. 5(d)], due to the lower emissivity for the radiative cooler, as compared to the case considered above when the ambient temperature is higher than the phase transition temperature. Qexchange also shows lower value. Therefore, the total outgoing power, Qtotal, is less than 7 W/m2 at the initial time and is further reduced to zero as time evolves. Such a small total outgoing power is consistent with the near complete absence of temperature variation as a function of time for the radiative cooler.

Finally, we simulate the radiative cooler temperature as a function of time under an outdoor condition. We solve Eqs. (3) and (9), using as inputs the ambient temperature [41] and solar illumination data [42] of July 19, 2017, representing typical summer climate conditions at Stanford, California. Figure 6 shows simulated radiative cooler temperature and radiative cooling power across a 24-hour period. Starting from 12:00 AM, the radiative cooler temperature is below critical temperature of 293 K. In this case, no cooling is required. The system is in the ‘off’ state with minimum cooling power and maintain its temperature near ambient temperature. As ambient temperature and device temperature increase in the morning, starting from around 9:00 AM, the radiative cooler temperature goes above the critical temperature and the system is switched to ‘on’ state, radiative cooling effect is clearly observed as the radiative cooler temperature deviates below the ambient temperature and reach a maximum difference of about 9 K around 3:00 PM when the ambient temperature is about 304 K. We also observe significant cooling power during this period when the radiative cooling is turned on. At 9:00 PM, when ambient temperature falls below the phase transition temperature, radiative cooling is again turned off. Such results clearly show the self-adaptive radiative cooling performance of our system. As a comparison, we also simulate the temperature performance of a static radiative cooler, assume its emissivity is the same as the ‘on’ state of the self-adaptive radiative cooler case shown in Fig. 3(i), but without the switching ability. As is shown in black dashed curve in Fig. 6, this static radiative cooler performs radiative cooling for the entire 24 hours, even when it’s cold at night and cooling is less desired.

 figure: Fig. 6

Fig. 6 Thermal performance of the self-adaptive radiative cooler (black curve) over a 24h cycle with ambient temperature variation (red curve). Radiative cooling power of the radiative cooler is also plotted (blue curve). As a comparison, the thermal performance of a static radiative cooler with the same emissivity as the ‘on’ state of the self-adaptive radiative cooler is also plotted (black dashed curve).

Download Full Size | PDF

5. Concluding Remarks

We have presented a self-adaptive radiative cooling system based on a photonic design incorporating phase change material VO2. This system can automatically turn on radiative cooling when the ambient temperature is above critical temperature and turn off radiative cooling when the ambient temperature is below. As a proof of concept demonstration, here we use simple planar stack structures. The concept here can be applied with other materials and structure systems for large scale implementations and can be generalized towards other radiative cooling applications, for example thermal textiles. Our results here explore new functionalities of radiative cooling applications and could be potentially used in a wide range of applications such as vehicles, buildings and textiles for energy consumption reduction and improved thermal comfort.

Funding, acknowledgments, and disclosures

This work is supported in part by the National Science Foundation (Grant No. CMMI-1562204), and by the Global Climate and Energy Project (GCEP) at Stanford University.

References and links

1. S. Catalanotti, V. Cuomo, G. Piro, D. Ruggi, V. Silvestrini, and G. Troise, “The radiative cooling of selective surfaces,” Sol. Energy 17(2), 83–89 (1975). [CrossRef]  

2. C. G. Granqvist and A. Hjortsberg, “Radiative cooling to low temperatures: General considerations and application to selectively emitting SiO films,” J. Appl. Phys. 52(6), 4205–4220 (1981). [CrossRef]  

3. A. R. Gentle and G. B. Smith, “Radiative Heat Pumping from the Earth Using Surface Phonon Resonant Nanoparticles,” Nano Lett. 10(2), 373–379 (2010). [CrossRef]   [PubMed]  

4. E. Rephaeli, A. Raman, and S. Fan, “Ultrabroadband photonic structures to achieve high-performance daytime radiative cooling,” Nano Lett. 13(4), 1457–1461 (2013). [CrossRef]   [PubMed]  

5. A. P. Raman, M. A. Anoma, L. Zhu, E. Rephaeli, and S. Fan, “Passive radiative cooling below ambient air temperature under direct sunlight,” Nature 515(7528), 540–544 (2014). [CrossRef]   [PubMed]  

6. S. Fan and A. Raman, “Metamaterials for radiative sky cooling,” Natl. Sci. Rev. 5(2), 10–13 (2018). [CrossRef]  

7. A. R. Gentle and G. B. Smith, “A Subambient Open Roof Surface under the Mid-Summer Sun,” Adv Sci (Weinh) 2(9), 1500119 (2015). [CrossRef]   [PubMed]  

8. M. M. Hossain, B. Jia, and M. Gu, “A Metamaterial Emitter for Highly Efficient Radiative Cooling,” Adv. Opt. Mater. 3(8), 1047–1051 (2015). [CrossRef]  

9. Z. Chen, L. Zhu, A. Raman, and S. Fan, “Radiative cooling to deep sub-freezing temperatures through a 24-h day-night cycle,” Nat. Commun. 7, 13729 (2016). [CrossRef]   [PubMed]  

10. J. Kou, Z. Jurado, Z. Chen, S. Fan, and A. J. Minnich, “Daytime Radiative Cooling Using Near-Black Infrared Emitters,” ACS Photonics 4(3), 626–630 (2017). [CrossRef]  

11. N. N. Shi, C. C. Tsai, F. Camino, G. D. Bernard, N. Yu, and R. Wehner, “Keeping cool: Enhanced optical reflection and radiative heat dissipation in Saharan silver ants,” Science (80-.). 349, 298– 301 (2015).

12. Y. Zhai, Y. Ma, S. N. David, D. Zhao, R. Lou, G. Tan, R. Yang, and X. Yin, “Scalable-manufactured randomized glass-polymer hybrid metamaterial for daytime radiative cooling,” Science (80-.). 355, 1062– 1066 (2017).

13. J. K. Tong, X. Huang, S. V. Boriskina, J. Loomis, Y. Xu, and G. Chen, “Infrared-Transparent Visible-Opaque Fabrics for Wearable Personal Thermal Management,” ACS Photonics 2(6), 769–778 (2015). [CrossRef]  

14. P. C. Hsu, A. Y. Song, P. B. Catrysse, C. Liu, Y. Peng, J. Xie, S. Fan, and Y. Cui, “Radiative human body cooling by nanoporous polyethylene textile,” Science (80-.). 353, 1019– 1023 (2016).

15. L. Zhu, A. Raman, K. X. Wang, M. A. Anoma, and S. Fan, “Radiative cooling of solar cells,” Optica 1(1), 32 (2014). [CrossRef]  

16. T. S. Safi and J. N. Munday, “Improving photovoltaic performance through radiative cooling in both terrestrial and extraterrestrial environments,” Opt. Express 23(19), A1120–A1128 (2015). [CrossRef]   [PubMed]  

17. S.-H. Wu and M. L. Povinelli, “Solar heating of GaAs nanowire solar cells,” Opt. Express 23(24), A1363–A1372 (2015). [CrossRef]   [PubMed]  

18. W. Li, Y. Shi, K. Chen, L. Zhu, and S. Fan, “A Comprehensive Photonic Approach for Solar Cell Cooling,” ACS Photonics 4(4), 774–782 (2017). [CrossRef]  

19. W. Li and S. Fan, “Nanophotonic Control of Thermal Radiation for Energy Applications,” Opt. Express 26(12), 15995–16021 (2018). [CrossRef]  

20. T. Inoue, M. De Zoysa, T. Asano, and S. Noda, “Realization of dynamic thermal emission control,” Nat. Mater. 13(10), 928–931 (2014). [CrossRef]   [PubMed]  

21. V. W. Brar, M. C. Sherrott, M. S. Jang, S. Kim, L. Kim, M. Choi, L. A. Sweatlock, and H. A. Atwater, “Electronic modulation of infrared radiation in graphene plasmonic resonators,” Nat. Commun. 6(1), 7032 (2015). [CrossRef]   [PubMed]  

22. X. Liu and W. J. Padilla, “Dynamic Manipulation of Infrared Radiation with MEMS Metamaterials,” Adv. Opt. Mater. 1(8), 559–562 (2013). [CrossRef]  

23. X. Liu and W. J. Padilla, “Reconfigurable room temperature metamaterial infrared emitter,” Optica 4(4), 430 (2017). [CrossRef]  

24. M. A. Kats, R. Blanchard, S. Zhang, P. Genevet, C. Ko, S. Ramanathan, and F. Capasso, “Vanadium dioxide as a natural disordered metamaterial: Perfect thermal emission and large broadband negative differential thermal emittance,” Phys. Rev. X 3, 041004 (2014).

25. S.-H. Wu, M. Chen, M. T. Barako, V. Jankovic, P. W. C. Hon, L. A. Sweatlock, and M. L. Povinelli, “Thermal homeostasis using microstructured phase-change materials,” Optica 4(11), 1390 (2017). [CrossRef]  

26. D. M. Bierman, A. Lenert, M. A. Kats, Y. Zhou, S. Zhang, M. De La Ossa, S. Ramanathan, F. Capasso, and E. N. Wang, “Radiative thermal runaway due to negative differential thermal emission across a solid-solid phase transition,” http://arxiv.org/abs/1801.00376 (2017).

27. Y. Qu, Q. Li, K. Du, L. Cai, J. Lu, and M. Qiu, “Dynamic Thermal Emission Control Based on Ultrathin Plasmonic Metamaterials Including Phase-Changing Material GST,” Laser Photonics Rev. 11(5), 1700091 (2017). [CrossRef]  

28. K.-K. Du, Q. Li, Y.-B. Lyu, J.-C. Ding, Y. Lu, Z.-Y. Cheng, and M. Qiu, “Control over emissivity of zero-static-power thermal emitters based on phase-changing material GST,” Light Sci. Appl. 6(1), e16194 (2017). [CrossRef]  

29. K. Du, L. Cai, H. Luo, Y. Lu, J. Tian, Y. Qu, P. Ghosh, Y. Lyu, Z. Cheng, M. Qiu, and Q. Li, “Wavelength-tunable mid-infrared thermal emitters with a non-volatile phase changing material,” Nanoscale 10(9), 4415–4420 (2018). [CrossRef]   [PubMed]  

30. L. Cai, K. Du, Y. Qu, H. Luo, M. Pan, M. Qiu, and Q. Li, “Nonvolatile tunable silicon-carbide-based midinfrared thermal emitter enabled by phase-changing materials,” Opt. Lett. 43(6), 1295–1298 (2018). [CrossRef]   [PubMed]  

31. H. W. Verleur, A. S. Barker, and C. N. Berglund, “Optical Properties of VO2 between 0.25 and 5 eV,” Phys. Rev. 172(3), 788–798 (1968). [CrossRef]  

32. J. B. Kana Kana, G. Vignaud, A. Gibaud, and M. Maaza, “Thermally driven sign switch of static dielectric constant of VO2thin film,” Opt. Mater. 54, 165–169 (2016). [CrossRef]  

33. F. J. Morin, “Oxides Which Show a Metal-to-Insulator Transition at the Neel Temperature,” Phys. Rev. Lett. 3(1), 34–36 (1959). [CrossRef]  

34. G. V. Jorgenson and J. C. Lee, “Doped vanadium oxide for optical switching films,” Sol. Energy Mater. 14(3-5), 205–214 (1986). [CrossRef]  

35. W. Burkhardt, T. Christmann, S. Franke, W. Kriegseis, D. Meister, B. K. Meyer, W. Niessner, D. Schalch, and A. Scharmann, “Tungsten and fluorine co-doping of VO2films,” Thin Solid Films 402(1-2), 226–231 (2002). [CrossRef]  

36. M. K. Dietrich, F. Kuhl, A. Polity, and P. J. Klar, “Optimizing thermochromic VO2 by co-doping with W and Sr for smart window applications,” Appl. Phys. Lett. 110(14), 141907 (2017). [CrossRef]  

37. E. Palik Handbook of Optical Constants of Solids (1998).

38. J.-Y. Jung, J. Y. Park, S. Han, A. S. Weling, and D. P. Neikirk, “Wavelength-selective infrared Salisbury screen absorber,” Appl. Opt. 53(11), 2431–2436 (2014). [CrossRef]   [PubMed]  

39. M. C. J. Large, D. R. McKenzie, and M. I. Large, “Incoherent reflection processes: a discrete approach,” Opt. Commun. 128(4-6), 307–314 (1996). [CrossRef]  

40. A. Berk, G. P. Anderson, P. K. Acharya, L. S. Bernstein, L. Muratov, J. Lee, M. Fox, S. M. Adler-Golden, J. H. Chetwynd Jr, M. L. Hoke, R. B. Lockwood, J. A. Gardner, T. W. Cooley, C. C. Borel, P. E. Lewis, and E. P. Shettle, “MODTRAN5: 2006 update,” Proc. SPIE 6233, 62331F (2006).

41. “Weather Underground,” https://www.wunderground.com/personal-weather-station/dashboard?ID=KCASTANF2#history/s20170719/e20170719/mdaily.

42. “Altitude and Azimuth of the Sun,” http://aa.usno.navy.mil/cgi-bin/aa_altazw.pl?form=1&body=10&year=2017&month=7&day=19&intv_mag=1&state=DC&place.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 (a) Schematic showing the concept of self-adaptive radiative cooling: when temperature is above the critical temperature Tc, radiative cooling is turned on; when temperature is below the critical temperature Tc, radiative cooling is turned off. (b) Schematic of the ideal spectrum for self-adaptive radiative cooling. The spectrum switches between an ‘on’ and an ‘off’ state depending on the ambient temperature.
Fig. 2
Fig. 2 The real part (a) and the imaginary part (b) of the dielectric constants for VO2 used in the paper in the wavelength range from visible to mid-infrared.
Fig. 3
Fig. 3 Photonic structures for realizing self-adaptive radiative cooling. (a) A schematic for the bottom radiative cooler, which is made of three layer VO2/MgF2/W structure. (b) Solar absorptivity of the radiative cooler. (c) Infrared emissivity of the radiative cooler. (d) A schematic showing the top spectrally-selective filter, which is made of 11 layers of Ge/MgF2. (e) Transmissivity of the filter in the solar wavelength range. (f) Transmissivity of the filter in the thermal wavelength range. (g) Schematic of the combined system. (h) Solar absorptivity of the radiative cooler in presence of the filter on top. (i) Infrared emissivity of the radiative cooler in presence of the filter on top.
Fig. 4
Fig. 4 (a) The emissivity of the bottom radiative cooler in presence of the top filter, as a function of incident angle and wavelength. The VO2 is in the metallic phase. (b) Angle and polarization averaged absorptivity spectrum of the bottom radiative cooler in presence of the top filter. Red and blue line shows the absorptivity in metallic and insulating VO2 phase, respectively.
Fig. 5
Fig. 5 Transient thermal performance. (a) Time-dependent temperature evolution of the radiative cooler under different ambient temperature. The initial temperature is assumed to be the same as ambient temperature. The whole system is exposed to the sky at time = 0. (b) Time-dependent Qtotal, Qcooler, Qexchange, of the radiative cooler, and the temperature of the filter T filter , at Tamb = 313 K. (c) Time-dependent Qtotal, Qcooler, Qexchange, of the radiative cooler, and the temperature of the filter T filter , at Tamb = 298 K. (d). Time-dependent Qtotal, Qcooler, Qexchange, of the radiative cooler, and temperature of the filter, T filter at Tamb = 283K.
Fig. 6
Fig. 6 Thermal performance of the self-adaptive radiative cooler (black curve) over a 24h cycle with ambient temperature variation (red curve). Radiative cooling power of the radiative cooler is also plotted (blue curve). As a comparison, the thermal performance of a static radiative cooler with the same emissivity as the ‘on’ state of the self-adaptive radiative cooler is also plotted (black dashed curve).

Tables (1)

Tables Icon

Table 1 Material composition and thicknesses for the designed spectrally-selective filter.

Equations (11)

Equations on this page are rendered with MathJax. Learn more.

λ4 d Μ n M
ϵ( λ,Ω )=( 1 r C )( t F + t F r C r F + t F ( r C r F ) 2 + t F ( r C r F ) 3 + )=  t F ( 1 r C )/( 1 r C r F )
Q total = Q cooler ( T )  Q atm ( T amb )  Q parasitic ( T, T amb ) Q sun ( T )  Q exchange ( T, T filter )
Q cooler ( T )=AdΩcosθ 0 dλ I BB ( T,λ )ϵ( λ,Ω,T )
Q atm ( T amb )=AdΩcosθ 0 dλ I BB ( T amb ,λ )ϵ( λ,Ω,T ) ϵ atm ( λ,Ω )
Q parasitic ( T, T amb )=Ah( T amb T )
Q sun ( T )=A 0 dλ I AM1.5 ( λ )ϵ( λ, θ sun ,T )
Q exchange ( T, T filter )= AdΩcosθ{ 0 dλ I BB ( T filter ,λ ) ϵ filter ( λ,Ω ) ϵ cooler ( λ,Ω,T ) 0 dλ I BB ( T,λ ) ϵ filter ( λ,Ω ) ϵ cooler ( λ,Ω,T ) }
Q total,F = Q filter ( T filter )  Q atm,F ( T amb )  Q parasitic,F ( T filter , T amb ) Q sun,F +  Q exchange ( T, T filter )
C cooler dT dt =  Q total ( T, T amb , T filter )
C filter d T filter dt =  Q total,F ( T,  T amb , T filter )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.