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

Directional radiation for optimal radiative cooling

Open Access Open Access

Abstract

The omnidirectional radiation scheme has been widely applied to thermal emitters for radiative cooling. We quantitatively illustrate that significant net radiative absorption at high zenith angles limits the performance of such isotropic emitters, and demonstrate that simply cutting off components corresponding to high angles can substantially improve the cooling performance of commonly used isotropic emitter designs. We also present an expression for the ideal directional spectral emissivity at conditions below ambient temperature. As our approach can be applied to coolers with arbitrary surfaces, our results may serve as a basic guideline for designing practical systems with various surfaces, such as rooftops or façades of modern buildings with complicated geometries.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The atmosphere of the Earth is relatively transparent to light in particular infrared wavelength ranges [Fig. 1(a)]. Radiative coolers utilize this transparency window for uninhibited outward radiation toward the cold cosmic background, thereby achieving below-ambient cooling at night [14] and even at daytime with strong solar reflection [516]; this is promising for designing zero-energy houses [1721] and novel energy harvesting systems [2224]. To achieve a high cooling power, the emissivity of the cooler has been designed with isotropic but spectrally selective in the 8–13 µm range for below-ambient cooling [5,16,2528] or spectrally broad at wavelengths from 4 µm to tens of microns for above-ambient cooling [6,2932]. Recently, in contrast with these approaches, a few studies have shown that the widely considered 8–13 µm wavelength range is not the optimal spectral range of emissivity at below-ambient temperatures [33]; instead, wavelength- and angle-selective emissivity can yield optimal emitters [34] for energy harvesting [24]. These studies have revealed that designing the emissivity at the expected operation temperature, rather than at the ambient temperature, is crucial for optimal radiative cooling. However, the angular aspect of emissivity for optimal cooling, which is deeply relevant to temperature-based spectral selection, has not been thoroughly examined. Most studies on radiative cooling have assumed, in their design phase, isotropic emissivity without explicit consideration of the actual operating temperature of the cooler, which can be substantially different from the ambient temperature [5,6,8,1012,14,16,2528]. Thus, the current state of research naturally prompts important questions about the impact of directionality on cooling and about the ideal wavelength and angular dependence of emissivity of radiative coolers as a function of their actual temperature as well as the ambient temperature.

 figure: Fig. 1.

Fig. 1. Spectral angular properties of the atmosphere. (a) Atmospheric transmittance. (b) Absorbed spectral radiance from the atmosphere (colored border line) and emitting spectral radiance for a perfect emitter (black solid line). For easy visualization, we convoluted the fine spectral features with a Gaussian smoothing function, g = exp(-λ2/(2σ2)) where σ = 5 nm in (b).

Download Full Size | PDF

In this letter, we explicitly and quantitatively demonstrate that the directionality of emission, in addition to spectral selectivity, is critically important for high-power, below-ambient cooling. Directional radiation by thermal emitters can enhance the net radiative power density, and their temperature can drop further below that of the best omni-directional emitters. Our angle-resolved investigation reveals that radiative heat exchange at high zenith angles causes heating rather than cooling at below-ambient temperatures, even within a commonly considered “transparency” window of 8–13 µm, owing to the dominant atmospheric radiation over the outward radiation from the cooler. The heating power and the angular range over which heating occurs increase as the temperature drops further below the ambient temperature. We show that popular emitters for radiative cooling, with emission bands in the 8–13 µm range or in a broad mid infrared (mid-IR) region, can exhibit significantly better cooling performance if their emission and absorption at high angles are simply blocked. Based on the selective emissivity design rule [24,33,34], we present the ideal spectral directional emissivity that yields the maximal allowed net radiative power at a given temperature and achieves the theoretical lower bound on the steady-state temperature, outperforming all other possible radiative cooling surface designs. We found that the steady-state temperature under a negligible non-radiative heat transfer condition can be much lower than the freezing point of mercury (−38.8 °C) and more than 17 K lower than that of the best isotropic emitter under the same atmospheric conditions. This explains why better cooling performance was observed in some previous experiments with implicitly imposed directionality [9,3537] than that observed for other designs. In addition, our results indicate that directional radiation is practically essential when the surface of the radiative cooler does not point to the zenith direction, such as in complicated rooftops and façades of modern buildings.

2. Theoretical model

In our model, a radiative cooler covers the top surface of the object to be cooled, and the cooler and the object are in the thermal equilibrium at temperature T. The cooler is at the sea level and radiates toward the sky. We first consider horizontally placed coolers with a sky-view angle of 2π steradians, but later generalize to an arbitrary angular orientation. The cooler exchanges radiation with the sun, the outer space, and the atmosphere, and exchanges heat non-radiatively (i.e., via conduction and convection) with the ambient air and other abutting objects. The radiation through the atmosphere is subject to various conditions, such as the season [38], humidity [39], and temperature variation along the altitude [40]. We capture these environmental effects by a collective variable α; the zenith angle of incident sunlight is θsun. We also assume that the temperatures of non-radiatively interacting objects are the same as the ambient air temperature Tamb. Then, the net cooling power density, which can be divided into radiative and non-radiative components, is

$$\begin{array}{c} {P_{\textrm{net}}}(T,{T_{\textrm{amb}}},\alpha ,{\theta _{\textrm{sun}}}) = {P_{\textrm{rad}}}(T,{T_{\textrm{amb}}},\alpha ,{\theta _{\textrm{sun}}}) - {P_{\textrm{non - rad}}}(T,{T_{\textrm{amb}}})\\ = {P_{\textrm{cooler}}}(T) - {P_{\textrm{sun}}}(T,{T_{\textrm{amb}}},\alpha ,{\theta _{\textrm{sun}}}) - {P_{\textrm{space}}}(T,{T_{\textrm{amb}}},\alpha ) - {P_{\textrm{atm}}}(T,{T_{\textrm{amb}}},\alpha ) - {P_{\textrm{non - rad}}}(T,{T_{\textrm{amb}}}), \end{array}$$
where Prad = PcoolerPsunPspacePatm is the net radiative power density, which accounts for the radiant exitance of the cooler Pcooler as well as the irradiance absorbed by the cooler owing to the sun, the outer space, and the atmosphere, Psun, Pspace, and Patm, respectively. On the other hand, Pnon-rad is the non-radiative power density absorbed from the surroundings. All power densities are in the units of W/m2. Among these, Pspace can be usually ignored because the cosmic background is much cooler than Tamb or T [19]. The rest of the radiative components in Eq. (1) are
$${P_{\textrm{cooler}}}(T) = \int\!\!\!\int {{{\tilde{I}}_{\textrm{BB}}}(\lambda ,T){\varepsilon _\textrm{c}}(\lambda ,\Omega ,T)\cos \theta \,\textrm{d}\Omega \textrm{d}\lambda } ,$$
$${P_{\textrm{sun}}}(T,{T_{\textrm{amb}}},\alpha ,{\theta _{\textrm{sun}}}) = \int\!\!\!\int {{{\tilde{I}}_{\textrm{sun}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}){\varepsilon _\textrm{c}}(\lambda ,\Omega ,T)\cos \theta \,\textrm{d}\Omega \textrm{d}\lambda } ,$$
$$\approx {A_{\textrm{sun}}}\cos {\theta _{\textrm{sun}}}\int {{{\tilde{I}}_{\textrm{sun}}}(\lambda ,{\Omega _{\textrm{sun}}},{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}){\varepsilon _\textrm{c}}(\lambda ,{\Omega _{\textrm{sun}}},T)\,\textrm{d}\lambda }, \qquad (2\textrm{b}^{\prime})$$
$${P_{\textrm{atm}}}(T,{T_{\textrm{amb}}},\alpha ) = \int\!\!\!\int {{{\tilde{I}}_{\textrm{atm}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha ){\varepsilon _\textrm{c}}(\lambda ,\Omega ,T)\cos \theta \,\textrm{d}\Omega \textrm{d}\lambda } ,$$
where $\int {({\cdot} )\,\textrm{d}\Omega } = \int_0^{2\mathrm{\pi }} {\int_0^{\mathrm{\pi }\textrm{/2}} {({\cdot} )\sin \theta \,\textrm{d}\theta \textrm{d}\phi } } $ is a hemispherical integral and ${\tilde{I}_{\textrm{BB}}}(\lambda ,T)$, ${\tilde{I}_{\textrm{sun}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}})$, and ${\tilde{I}_{\textrm{atm}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha )$ are the spectral radiance (radiated power per unit area per unit wavelength per unit solid angle) of an ideal blackbody, the sun, and the atmosphere, respectively. The spectral radiances have the unit of W/(nm·m2·sr). ɛc(λ, Ω, T) is the spectral angular emissivity of the cooler, which is a unitless quantity representing the ratio of the spectral radiance of the cooler to that of a maximally emitting surface modelled by Planck’s law, and is a function of the wavelength, angular direction, and temperature. The spectral angular emissivity is used instead of the absorptivity in Eqs. (2b)–(2c) owing to Kirchhoff’s law. Asun is the solid angle of the sun seen from Earth (∼ 6×10−5 sr). Because the subtended angle of the sun is very small (∼ 0.5°), we assumed that both ${\tilde{I}_{\textrm{sun}}}$ and εc are constant within it when we derived Eq. (2b’). After rearranging, the net radiative power density becomes
$${P_{\textrm{rad}}}(T,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}) = \int {\textrm{d}\Omega \,\cos \theta \int {\textrm{d}\lambda \,{{\tilde{I}}_{\textrm{rad}}}(\lambda ,\Omega ,T,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}){\varepsilon _\textrm{c}}(\lambda ,\Omega ,T)} } ,$$
where ${\tilde{I}_{\textrm{rad}}}(\lambda ,\Omega ,T,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}) = {\tilde{I}_{\textrm{BB}}}(\lambda ,T) - {\tilde{I}_{\textrm{sun}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}) - {\tilde{I}_{\textrm{atm}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha )$ is the net spectral radiance. In addition, we can represent the non-radiative part in Eq. (1) as
$${P_{\textrm{non - rad}}}(T,{T_{\textrm{amb}}}) = {h_\textrm{c}}({T_{\textrm{amb}}} - T),$$
where hc is the effective non-radiative heat transfer coefficient, which depends on various environmental conditions, such as the wind speed [19] and thermal insulators surrounding the cooler and the cooled object [41].

To solve Eqs. (2b)–(2c), we approximate ${\tilde{I}_{\textrm{sun}}}$(${\tilde{I}_{\textrm{atm}}}$) as a multiplication of the atmospheric transmittance (atmospheric emissivity) and extraterrestrial solar radiance (radiative spectrum of an ideal blackbody at the ambient temperature). The atmospheric transmittance and its emissivity are related via ɛatm(λ, Ω, Tamb, α) = 1 – tatm(λ, Ω, Tamb, α) [5,42]. The attenuation in tatm can be obtained from the effective path length through atmosphere, such relation can be arranged as tatm(λ, Ω, Tamb, α) = t0AM(θ), where t0(λ, Tamb, α) is the transmittance in the zenith direction (data from MODTRAN 6 [43]) and AM(θ) is angular attenuation coefficient [33]. In this work, we use AM(θ) of the spherical shell model for the atmosphere [33,44], which is more accurate than AM(θ) = 1/cosθ of the flat earth model especially at high zenith angles [19]. The other parameters and their values in our calculations were as follows: the season, latitude, Ωsun, and Tamb are summer, 36.35°N, 12.9° (zenith angle at noon), and 300 K, respectively.

3. Results and discussion

As a motivation for a more detailed analysis, we begin by comparing the spectral radiance of an isotropic blackbody and that of the atmosphere, which correspond to ${\tilde{I}_{\textrm{BB}}}(\lambda ,T)$ and ${\tilde{I}_{\textrm{atm}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha )$ in Eqs. (2a) and (2c), respectively, at several emitter temperatures and zenith angles. The spectral radiance of a blackbody depends on its temperature, but is angle-independent owing to the isotropy assumption. On the other hand, the spectral radiance of the atmosphere is independent of the emitter temperature, but strongly depends on the zenith angle [24,34] because of the effective thickness of the atmosphere and, hence, its transparency are determined by the zenith angle [Fig. 1(a)]. For large zenith angles approaching 90°, it is apparent that the atmospheric transmittance diminishes at all wavelengths, indicating increased absorption. According to reciprocity, this means that the atmospheric emission is intense at these large zenith angles, and it can surpass the emission from a perfect thermal emitter at below-ambient temperatures [Fig. 1(b)]. Shown in Fig. 1(b) are ${\tilde{I}_{\textrm{BB}}}(\lambda ,T)$ at three emitter temperatures (black lines) and ${\tilde{I}_{\textrm{atm}}}(\lambda ,\Omega ,{T_{\textrm{amb}}},\alpha )$ at three zenith angles (colored graphs). When the black line is above the colored graph, the net spectral radiance (${\tilde{I}_{\textrm{BB}}} - {\tilde{I}_{\textrm{atm}}}$) is positive at that specific wavelength, emitter temperature, and zenith angle. If the emitter is at the same temperature (300 K) as Tatm, the net radiance is positive at all wavelengths at all zenith angles, meaning that radiative channels at all wavelengths and angles can contribute to cooling. However, if the emitter temperature drops to 280 K, the net radiation is negative for most wavelengths even within the 8–13 µm transparency window at the zenith angle of 80°, indicating that radiative channels near the zenith angle of 80° or larger actually contribute to heating rather than cooling. If the temperature decreases further to 265 K, a significant part of the transparency window (∼ 60%) contributes to heating even at 60°. These examples show that as the temperature of the emitter decreases, the atmospheric radiance begins to dominate over the radiance of the emitter at progressively smaller zenith angles. The deficiency of isotropic emitters as sub-ambient radiative coolers is evident, especially when the target temperature is well below the ambient temperature.

Building on these initial findings, we quantitatively analyzed the performance of two popular isotropic emitter designs at all zenith angles, over a wide range of cooler temperatures. The two most widely used isotropic emitter designs have unit emissivity values either in the 8–13 µm range (their spectral emissivity is denoted hereafter $\varepsilon _{8 - 13}^{(\textrm{iso})}$) or in the 4–20 µm range or longer wavelengths (denoted $\varepsilon _{\textrm{broad}}^{(\textrm{iso})}$), while the emissivity in the other spectral regions is zero. In this case, we assume that the emitters are positioned flat on the ground, facing directly the zenith direction. Their net radiance in reference to the emitter surface area, instead of the transverse area normal to the direction of radiation, is $\cos (\theta )\,{\tilde{P}_{\textrm{rad}}} = \cos \theta \int {\textrm{d}\lambda \,{{\tilde{I}}_{\textrm{rad}}}{\varepsilon _\textrm{c}}}$ and has the unit W/(m2·sr). At the ambient temperature, the emitters exhibit non-negative net radiance in all upper hemispherical directions; thus, radiative channels at all zenith angles contribute to cooling [Figs. 2(a)–2(b)]. However, at below-ambient temperatures, net radiances become negative at large zenith angles [Figs. 2(c)–2(f)], which originates from two facts: the gradual increase of the atmospheric radiance at higher zenith angles and the gradual decrease of the blackbody radiance at lower temperatures, as illustrated earlier in Fig. 1(b). These are also related to the reasons why radiative cooling is not highly effective in reducing the surface temperature to a habitable level on Venus with its hot and dense atmosphere. Note that the emissivity of the cooler is kept constant regardless of the zenith angle or the temperature. Among the two emitter designs, the broadband emitter suffers more severe heating, and at 280 K, net radiative heating occurs at all angles, as shown in Fig. 2(f). More quantitatively, the azimuthally integrated power densities per unit zenith angle in Figs. 2(g)–2(h) reveal that the emitters encounter stronger radiative heating at high zenith angles when they cool further below the ambient temperature. Owing to this effect, isotropic emitters cannot passively cool down past the dashed lines in Figs. 2(g)–2(h), which indicate the temperatures for the zero net radiative power density (268.5 K and 286.9 K for $\varepsilon _{8 - 13}^{(\textrm{iso})}$ and $\varepsilon _{\textrm{broad}}^{(\textrm{iso})}$, respectively).

 figure: Fig. 2.

Fig. 2. Net radiance dependence on temperature as (a, b) 300 K, (c, d) 290 K, and (e, f) 280 K, for horizontally placed (a, c, e) isotropic 8–13 µm emitter and (b, d, f) isotropic broadband emitter. Net radiance per unit zenith angle for (g) isotropic 8–13 µm emitter and (h) isotropic broadband emitter. Dashed lines in (g, h) indicate steady-state temperature with zero radiative power density.

Download Full Size | PDF

Now, we demonstrate that by suppressing radiative channels at high zenith angles, emitters can be designed for reaching temperatures below the temperature limit of isotropic emitters. We focus on directional emitters that possess the same spectral emissivities as isotropic emitters within a cut-off angle θc and zero emissivity above θc [Figs. 3(a)–3(b)]. In other words, they are angle-limited versions of isotropic emitters. Then, the steady-state temperature can be obtained by solving Eq. (1) for Pnet = 0 depending on θc and hc [Figs. 3(c)–3(d)]. The optimal θc for the lowest possible steady-state temperature can be found numerically and is shown as red triangular points in Figs. 3(c)–3(d), for different hc values. The optimal θc is smaller for broadband emitters than for selective emitters. For example, at hc = 3 W/(m2K), θc is 72.3° and 61.4° for the 8–13 µm emitters ($\varepsilon _{\textrm{8 - 13}}^{(\textrm{dir})}$) and broadband emitters ($\varepsilon _{\textrm{broad}}^{(\textrm{dir})}$), respectively. As hc decreases, the optimal θc becomes even smaller, and the difference between the steady-state temperature of optimal directional emitters and isotropic emitters increases. For instance, at hc = 0.4 W/(m2K), optimal directional emitters $\varepsilon _{\textrm{8 - 13}}^{(\textrm{dir})}$ can freeze water at the steady-state temperature of 271.8 K, whereas isotropic emitters with $\varepsilon _{\textrm{8 - 13}}^{(\textrm{iso})}$ cannot.

 figure: Fig. 3.

Fig. 3. Directional radiation for limited high angular emissivity. (a) Schematic of directional radiation. (b) Angularly limited emissivities of 8–13 µm emitter and broadband mid-IR emitter. (c, d) Steady-state temperature for various cutoff angles and non-radiative heat transfer coefficients.

Download Full Size | PDF

Now, we derive the ideal spectral directional emissivity that maximizes Prad at a given emitter temperature by implementing a selective emissivity design rule that we developed in Ref. [33] and generalized here to include the angle dependence. In Eq. (3), the emissivity (εc) affects Prad as a non-negative multiplicative factor to the net spectral radiance (${\tilde{I}_{\textrm{rad}}}$) at each (λ,ω) combination. Thus, Prad can be maximized by assigning the largest allowed value of emissivity (i.e., unity) to (λ,ω) combinations with ${\tilde{I}_{\textrm{rad}}} > 0$ and the smallest allowed emissivity (i.e., zero) to (λ,ω) combinations with ${\tilde{I}_{\textrm{rad}}} < 0$. Such a design scheme can be expressed simply as

$$\varepsilon _{\textrm{ideal}}^{(\textrm{dir})}(\lambda ,\Omega ,T,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}) = \frac{1}{2}[1 + {\mathop{\rm sgn}} ({\tilde{I}_{\textrm{rad}}})],$$
where ${\mathop{\rm sgn}} (x) = \mathop \{ \nolimits_{ - 1\quad \textrm{if}\,x < 0}^{1\quad \textrm{if}\,x > 0} $ is the sign function, whose value at x = 0 is undefined but does not affect Prad as long as it is finite. As a main result of our work, Fig. 4(a) shows that $\varepsilon _{\textrm{ideal}}^{(\textrm{dir})}$ is highly sensitive to temperature and becomes more directional and spectrally selective as temperature decreases. It also indicates that radiative heating at high zenith angles becomes progressively more significant as the emitter is cooled to lower temperatures, and that it undermines the overall cooling performance if not blocked.

 figure: Fig. 4.

Fig. 4. Radiative cooling for horizontally placed emitters. (a) Ideal spectral directional emissivity depending on the cooler’s temperature, where the colored region presents unit emissivity. (b) Radiative (black, blue, and red) and non-radiative (green) power densities as a function of the cooler’s temperature. (c) The steady-state temperature and temperature difference between the isotropic emitter and the optimal directional emitter, as a function of the non-radiative heat transfer coefficient.

Download Full Size | PDF

The ideal spectral directional emissivity derived above imposes an upper bound on Prad at a given temperature, which can never be exceeded by any other emitting surface in reciprocal systems [Fig. 4(b)]. Even when some additional constraints are imposed on the spectral directional emissivity, Eq. (3) can still be used for deriving an optimal solution under given constraints. In those cases, however, the resulting Prad will be lower than what can be achieved by $\varepsilon _{\textrm{ideal}}^{(\textrm{dir})}$ without such constraints. For example, one can impose the isotropy constraint (i.e., the emissivity is not a function of direction), and the optimal spectral emissivity (hereinafter referred to as $\varepsilon _{\textrm{ideal}}^{(\textrm{iso})}$) has an expression similar to Eq. (5), but with ${\tilde{I}_{\textrm{rad}}}$ replaced by its directionally integrated value. This spectral emissivity places an upper bound on Prad for all isotropic designs, but this bound is lower than that imposed by $\varepsilon _{\textrm{ideal}}^{(\textrm{dir})}$. Instead of constraints on directional properties, one can also consider constraints on spectral features and a “single-band” constraint is such an example, in which non-zero emissivity only occurs over a single, continuous band of wavelengths. Two well-known cases are the 8–13 µm selective emitter and the broadband emitter that we have already reviewed above, and optimal directional emitters with such constraints ($\varepsilon _{\textrm{8 - 13}}^{(\textrm{dir})}$ and $\varepsilon _{\textrm{broad}}^{(\textrm{dir})}$) are in fact angle-limited emitters that we derived in Fig. 3; these outperform the corresponding isotropic single-band emitters ($\varepsilon _{\textrm{8 - 13}}^{(\textrm{iso})}$ and $\varepsilon _{\textrm{broad}}^{(\textrm{iso})}$, respectively) in terms of Prad. These examples reveal that the directional selectivity and spectral selectivity of emissivity are crucial for obtaining the highest possible Prad.

In addition to Prad at a given temperature, another important indicator of the cooler performance is the steady-state temperature of the cooler under given atmospheric conditions and for a given solar zenith angle. In contrast to Prad, the steady-state temperature strongly depends on how well the cooled object is insulated to reduce conductive and convective heat exchange with surroundings (as phenomenologically captured by hc in our model). With perfect insulation (i.e., hc = 0), the steady-state temperature of an object radiatively cooled with $\varepsilon _{\textrm{ideal}}^{(\textrm{dir})}$ can be as low as 226.5 K at mid-day in summer. If it is placed horizontally on the ground at latitude 36.35°N, this temperature serves as the lower bound on the steady-state temperature for any spectral directional emissivity design. Below this temperature, $\varepsilon _{\textrm{ideal}}^{(\textrm{dir})}$ in Eq. (5) is zero because ${\tilde{I}_{\textrm{rad}}}$ is negative for any (λ,Ω) combination, and any thermal emitter with non-zero emissivity is radiatively heated owing to a negative Prad, as illustrated in Fig. 4(b). Similarly, the steady-state temperature of $\varepsilon _{\textrm{ideal}}^{(\textrm{iso})}$ is also bounded from below, but the isotropy constraint places the bound at a higher level of 243.6 K, as reported in Ref. [33].

In general, the steady-state condition, Pnet = 0, can be solved graphically by finding a crossing point between Prad and Pnon-rad (green solid lines) in Fig. 4(b). Under negligible non-radiative heat transfer conditions (hc = 0), the directional emitters with $\varepsilon _{\textrm{ideal}}^{(\textrm{dir})}$, $\varepsilon _{\textrm{8 - 13}}^{(\textrm{dir})}$, and $\varepsilon _{\textrm{broad}}^{(\textrm{dir})}$ exhibit steady-state temperatures lower than those of their isotropic counterparts, by 17.1 K, 18.5 K, and 36.9 K, respectively. In the presence of a non-radiative effect (hc ≠ 0), the emitters are stabilized at higher temperatures, but the directional selectivity helps to attain lower temperatures. This also means that, when cooling an object down to a specific target temperature, a directional emitter requires less heat insulation than the corresponding isotropic emitter. For example, coolers with $\varepsilon _{\textrm{8 - 13}}^{(\textrm{dir})}$ can reach the freezing point of water when hc ≤ 0.49 W/(m2K), in contrast to isotropic coolers $\varepsilon _{\textrm{8 - 13}}^{(\textrm{iso})}$ that require hc ≤ 0.29 W/(m2K), which means that the insulation materials for isotropic coolers should be at least ∼70% thicker than those for directional coolers.

In actual applications of radiative cooling, the surfaces of target objects (such as buildings and vehicles) typically are not fully parallel to the ground: building façades and slanted rooftops of houses are such examples. In these cases, in which the surface normal direction of the emitter attached to the surface is not aligned with the zenith, directional selectivity becomes even more important and the effect can be analyzed by modifying Eq. (3) as

$${P_{\textrm{rad}}}(T,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}},{\theta _{\textrm{slope}}}) = \int\!\!\!\int {\tilde{I}^{\prime}_{\textrm{rad}}(\lambda ,\Omega ,T,{T_{\textrm{amb}}},\alpha ,{\Omega _{\textrm{sun}}}){\varepsilon _\textrm{c}}(\lambda ,\Omega ,T)\cos \theta^{\prime }\,\textrm{d}\Omega \textrm{d}\lambda } ,$$
where θslope is the zenith angle corresponding to the surface normal direction of the slanted emitter (“slope angle”), θ’ is the angle of radiation measured from the surface normal, and $\tilde{I}^{\prime}_{\textrm{rad}}$ is the net radiance that is the same as ${\tilde{I}_{\textrm{rad}}}$ for acute zenith angles (θ < π/2), but for obtuse angles (θ > π/2), it entails the effect of reflection and emission from the ground based on its effective albedo (R) as $\tilde{I}^{\prime}_{\textrm{rad}} = {\tilde{I}_{\textrm{BB}}} - R(\tilde{I}_{\textrm{sun}}^{(\textrm{iso})} + \tilde{I}_{\textrm{atm}}^{(\textrm{iso})}) - (1 - R){\tilde{I}_{\textrm{ground}}}$. Here, we assumed diffuse reflection for sunlight and atmospheric emission, and used their average spectral radiances, $\tilde{I}_{\textrm{sun}}^{(\textrm{iso})}$ and $\tilde{I}_{\textrm{atm}}^{(\textrm{iso})}$, in the calculation. ${\tilde{I}_{\textrm{ground}}}$ is the radiance of a blackbody at the same temperature as the ground. In the following calculations, we assume θslope = 60° and R = 0, but the method and qualitative conclusion remain valid for other choices of θslope = 60° and R = 0. Even when slanted, the favorable direction for emitting thermal radiation is still near the zenith direction. As expected, directional emitters present larger Prad values compared to those presented by isotropic emitters, as shown in Fig. 5(a), and their difference is larger compared with that in the case of horizontal emitters in Fig. 4(a). This also makes the difference between the steady-state temperatures of directional and isotropic emitters more pronounced than in the case of horizontal emitters [Fig. 5(b)]. Figure 5(c) shows that slanted emitters with isotropic emissivity absorb thermal radiation strongly at obtuse and large acute zenith angles and, for optimal radiative cooling, the emissivity should be designed to be maximal in the off-normal directions pointing toward the zenith. Another notable point is that, in the titled case, the broadband directional emitter approaches cooling performances of the 8–13 µm isotropic emitter, which indicates that directionality design can be as important as spectral selectivity design. To realize the directional radiation of off-normal emitters, we may use asymmetric structures, including gratings and metasurfaces that have been actively studied [45,46].

 figure: Fig. 5.

Fig. 5. Cooling performance for slanted emitters. (a) Radiative (black, blue, and red) and non-radiative (green) power densities as a function of the cooler’s temperature. (b) The steady-state temperature and temperature difference between the isotropic emitter and the optimal directional emitter as a function of the non-radiative heat transfer coefficient. (c) Net radiance dependence on temperature.

Download Full Size | PDF

4. Conclusion

In conclusion, we investigated the directional aspect of emissivity for improving the net radiative power density as well as the steady-state temperatures of radiative coolers. We revealed that previous isotropic emitters can be cooled further by simply restricting the maximal angle of radiation while maintaining their spectral properties. We also presented the ideal spectral directional emissivity for general environmental conditions and derived the ultimate upper bound on the net radiative power density and the lower bound on the steady-state temperature. In particular, we showed that the steady-state temperature can be in principle extended down to 226.5 K, which is 17.1 K lower than that of the ideal isotropic emitter. The results were generalized to slanted emitters, making them useful for the optimal design of directional emissivity in practical situations with complex surface morphologies.

Funding

National Research Foundation of Korea (2018M3D1A1058998, 2019M3A6B3031046).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. B. Orel, M. K. Gunde, and A. Krainer, “Radiative cooling efficiency of white pigmented paints,” Sol. Energy 50(6), 477–482 (1993). [CrossRef]  

2. 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]  

3. M. Dong, N. Chen, X. Zhao, S. Fan, and Z. Chen, “Nighttime radiative cooling in hot and humid climates,” Opt. Express 27(22), 31587 (2019). [CrossRef]  

4. B. Zhao, M. Hu, X. Ao, N. Chen, Q. Xuan, D. Jiao, and G. Pei, “Performance analysis of a hybrid system combining photovoltaic and nighttime radiative cooling,” Appl. Energy 252, 113432 (2019). [CrossRef]  

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]  

6. 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 355(6329), 1062–1066 (2017). [CrossRef]  

7. S. Y. Heo, G. J. Lee, D. H. Kim, Y. J. Kim, S. Ishii, M. S. Kim, T. J. Seok, B. J. Lee, H. Lee, and Y. M. Song, “A Janus emitter for passive heat release from enclosures,” Sci. Adv. 6(36), eabb1906 (2020). [CrossRef]  

8. B. Xiang, R. Zhang, Y. Luo, S. Zhang, L. Xu, H. Min, S. Tang, and X. Meng, “Nano Energy 3D porous polymer film with designed pore architecture and auto-deposited SiO 2 for highly efficient passive radiative cooling,” Nano Energy 81, 105600 (2021). [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(1), 13729 (2016). [CrossRef]  

10. J. long 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. B. Bhatia, A. Leroy, Y. Shen, L. Zhao, M. Gianello, D. Li, T. Gu, J. Hu, M. Soljačić, and E. N. Wang, “Passive directional sub-ambient daytime radiative cooling,” Nat. Commun. 9(1), 5001 (2018). [CrossRef]  

12. A. Leroy, B. Bhatia, C. C. Kelsall, A. Castillejo-Cuberos, M. H. Di Capua, L. Zhao, L. Zhang, A. M. Guzman, and E. N. Wang, “High-performance subambient radiative cooling enabled by optically selective and thermally insulating polyethylene aerogel,” Sci. Adv. 5(10), eaat9480 (2019). [CrossRef]  

13. X. Ao, M. Hu, B. Zhao, N. Chen, G. Pei, and C. Zou, “Preliminary experimental study of a specular and a diffuse surface for daytime radiative cooling,” Sol. Energy Mater. Sol. Cells 191, 290–296 (2019). [CrossRef]  

14. T. Li, Y. Zhai, S. He, W. Gan, Z. Wei, M. Heidarinejad, D. Dalgo, R. Mi, X. Zhao, J. Song, J. Dai, C. Chen, A. Aili, A. Vellore, A. Martini, R. Yang, J. Srebric, X. Yin, and L. Hu, “A radiative cooling structural material,” Science 364(6442), 760–763 (2019). [CrossRef]  

15. X. Xue, M. Qiu, Y. Li, Q. M. Zhang, S. Li, Z. Yang, C. Feng, W. Zhang, J. G. Dai, D. Lei, W. Jin, L. Xu, T. Zhang, J. Qin, H. Wang, and S. Fan, “Creating an Eco-Friendly Building Coating with Smart Subambient Radiative Cooling,” Adv. Mater. 32(42), 1906751 (2020). [CrossRef]  

16. D. Chae, S. Son, Y. Liu, H. Lim, and H. Lee, “High-Performance Daytime Radiative Cooler and Near-Ideal Selective Emitter Enabled by Transparent Sapphire Substrate,” Adv. Sci. 7(19), 2001577 (2020). [CrossRef]  

17. X. Lu, P. Xu, H. Wang, T. Yang, and J. Hou, “Cooling potential and applications prospects of passive radiative cooling in buildings: The current state-of-the-art,” Renewable Sustainable Energy Rev. 65, 1079–1097 (2016). [CrossRef]  

18. M. Zeyghami, D. Y. Goswami, and E. Stefanakos, “A review of clear sky radiative cooling developments and applications in renewable power systems and passive building cooling,” Sol. Energy Mater. Sol. Cells 178, 115–128 (2018). [CrossRef]  

19. D. Zhao, A. Aili, Y. Zhai, S. Xu, G. Tan, X. Yin, and R. Yang, “Radiative sky cooling: Fundamental principles, materials, and applications,” Appl. Phys. Rev. 6(2), 021306 (2019). [CrossRef]  

20. B. Zhao, M. Hu, X. Ao, N. Chen, Q. Xuan, Y. Su, and G. Pei, “A novel strategy for a building-integrated diurnal photovoltaic and all-day radiative cooling system,” Energy 183, 892–900 (2019). [CrossRef]  

21. J. Mandal, Y. Yang, N. Yu, and A. P. Raman, “Paints as a Scalable and Effective Radiative Cooling Technology for Buildings,” Joule 4(7), 1350–1356 (2020). [CrossRef]  

22. A. P. Raman, W. Li, and S. Fan, “Generating Light from Darkness,” Joule 3(11), 2679–2686 (2019). [CrossRef]  

23. W. Li, S. Buddhiraju, and S. Fan, “Thermodynamic limits for simultaneous energy harvesting from the hot sun and cold outer space,” Light: Sci. Appl. 9(1), 68 (2020). [CrossRef]  

24. L. Fan, W. Li, W. Jin, M. Orenstein, and S. Fan, “Maximal nighttime electrical power generation via optimal radiative cooling,” Opt. Express 28(17), 25460–25470 (2020). [CrossRef]  

25. Z. Huang and X. Ruan, “Nanoparticle embedded double-layer coating for daytime radiative cooling,” Int. J. Heat Mass Transfer 104, 890–896 (2017). [CrossRef]  

26. C. Zou, G. Ren, M. M. Hossain, S. Nirantar, W. Withayachumnankul, T. Ahmed, M. Bhaskaran, S. Sriram, M. Gu, and C. Fumeaux, “Metal-Loaded Dielectric Resonator Metasurfaces for Radiative Cooling,” Adv. Opt. Mater. 5, 1700460 (2017). [CrossRef]  

27. M. Ono, K. Chen, W. Li, and S. Fan, “Self-adaptive radiative cooling based on phase change materials,” Opt. Express 26(18), A777 (2018). [CrossRef]  

28. Y. Li, L. Li, L. Guo, and B. An, “Systematical analysis of ideal absorptivity for passive radiative cooling,” Opt. Mater. Express 10(8), 1767–1777 (2020). [CrossRef]  

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

30. P. Hsu, C. Liu, A. Y. Song, Z. Zhang, Y. Peng, J. Xie, K. Liu, C. Wu, P. B. Catrysse, L. Cai, S. Zhai, A. Majumdar, S. Fan, and Y. Cui, “A dual-mode textile for human body radiative heating and cooling,” Sci. Adv. 3(11), e1700895 (2017). [CrossRef]  

31. W. Li, Y. Shi, Z. Chen, and S. Fan, “Photonic thermal management of coloured objects,” Nat. Commun. 9(1), 1–8 (2018). [CrossRef]  

32. S. Son, S. Jeon, D. Chae, S. Y. Lee, Y. Liu, H. Lim, S. J. Oh, and H. Lee, “Colored emitters with silica-embedded perovskite nanocrystals for efficient daytime radiative cooling,” Nano Energy 79, 105461 (2021). [CrossRef]  

33. S. Jeon and J. Shin, “Ideal spectral emissivity for radiative cooling of earthbound objects,” Sci. Rep. 10(1), 1–7 (2020). [CrossRef]  

34. P. Wray, M. Su, and H. Atwater, “Design of efficient radiative emission and daytime cooling structures with Si3N4 and SiO2 nanoparticle laminate films,” Opt. Express 28(24), 35784–35794 (2020). [CrossRef]  

35. L. Zhou, H. Song, J. Liang, M. Singer, M. Zhou, E. Stegenburgs, N. Zhang, C. Xu, T. Ng, Z. Yu, B. Ooi, and Q. Gan, “A polydimethylsiloxane-coated metal structure for all-day radiative cooling,” Nat. Sustain. 2(8), 718–724 (2019). [CrossRef]  

36. Y. Qu, Y. Qu, Y. Qu, M. Pan, M. Qiu, and M. Qiu, “Directional and spectral control of thermal emission and its application in radiative cooling and infrared light sources,” Phys. Rev. Appl. 13(6), 064052 (2020). [CrossRef]  

37. A. R. Gentle and G. B. Smith, “Angular selectivity: impact on optimised coatings for night sky radiative cooling,” Proc. SPIE 7404, 74040J (2009). [CrossRef]  

38. Y. Iijima and M. Shinoda, “The influence of seasonally varying atmospheric characteristics on the intensity of nocturnal cooling in a high mountain hollow,” J. Appl. Meteorol. 41(7), 734–743 (2002). [CrossRef]  

39. A. W. Harrison, “Effect of atmospheric humidity on radiation cooling,” Sol. Energy 26(3), 243–247 (1981). [CrossRef]  

40. X. Berger, J. Bathiebo, F. Kieno, and C. N. Awanou, “Clear sky radiation as a function of altitude,” Renewable Energy 2(2), 139–157 (1992). [CrossRef]  

41. W. A. B. John and A. Duffie, Solar Engineering of Thermal Processes, 4th ed. (Wiley, 2013).

42. 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]  

43. A. Berk, P. Conforti, R. Kennett, T. Perkins, F. Hawes, and J. van den Bosch, “MODTRAN radiative transfer code,” Proc. SPIE 9088, 90880H (2014). [CrossRef]  

44. F. Kasten and A. T. Young, “Revised optical air mass tables and approximation formula,” Appl. Opt. 28(22), 4735–4738 (1989). [CrossRef]  

45. X. Wang, A. Díaz-Rubio, V. S. Asadchy, G. Ptitcyn, A. A. Generalov, J. Ala-Laurinaho, and S. A. Tretyakov, “Extreme Asymmetry in Metasurfaces via Evanescent Fields Engineering: Angular-Asymmetric Absorption,” Phys. Rev. Lett. 121(25), 256802 (2018). [CrossRef]  

46. M. Haghtalab, M. Tamagnone, A. Y. Zhu, S. Safavi-Naeini, and F. Capasso, “Ultrahigh Angular Selectivity of Disorder-Engineered Metasurfaces,” ACS Photonics 7(4), 991–1000 (2020). [CrossRef]  

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 (5)

Fig. 1.
Fig. 1. Spectral angular properties of the atmosphere. (a) Atmospheric transmittance. (b) Absorbed spectral radiance from the atmosphere (colored border line) and emitting spectral radiance for a perfect emitter (black solid line). For easy visualization, we convoluted the fine spectral features with a Gaussian smoothing function, g = exp(-λ2/(2σ2)) where σ = 5 nm in (b).
Fig. 2.
Fig. 2. Net radiance dependence on temperature as (a, b) 300 K, (c, d) 290 K, and (e, f) 280 K, for horizontally placed (a, c, e) isotropic 8–13 µm emitter and (b, d, f) isotropic broadband emitter. Net radiance per unit zenith angle for (g) isotropic 8–13 µm emitter and (h) isotropic broadband emitter. Dashed lines in (g, h) indicate steady-state temperature with zero radiative power density.
Fig. 3.
Fig. 3. Directional radiation for limited high angular emissivity. (a) Schematic of directional radiation. (b) Angularly limited emissivities of 8–13 µm emitter and broadband mid-IR emitter. (c, d) Steady-state temperature for various cutoff angles and non-radiative heat transfer coefficients.
Fig. 4.
Fig. 4. Radiative cooling for horizontally placed emitters. (a) Ideal spectral directional emissivity depending on the cooler’s temperature, where the colored region presents unit emissivity. (b) Radiative (black, blue, and red) and non-radiative (green) power densities as a function of the cooler’s temperature. (c) The steady-state temperature and temperature difference between the isotropic emitter and the optimal directional emitter, as a function of the non-radiative heat transfer coefficient.
Fig. 5.
Fig. 5. Cooling performance for slanted emitters. (a) Radiative (black, blue, and red) and non-radiative (green) power densities as a function of the cooler’s temperature. (b) The steady-state temperature and temperature difference between the isotropic emitter and the optimal directional emitter as a function of the non-radiative heat transfer coefficient. (c) Net radiance dependence on temperature.

Equations (9)

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

P net ( T , T amb , α , θ sun ) = P rad ( T , T amb , α , θ sun ) P non - rad ( T , T amb ) = P cooler ( T ) P sun ( T , T amb , α , θ sun ) P space ( T , T amb , α ) P atm ( T , T amb , α ) P non - rad ( T , T amb ) ,
P cooler ( T ) = I ~ BB ( λ , T ) ε c ( λ , Ω , T ) cos θ d Ω d λ ,
P sun ( T , T amb , α , θ sun ) = I ~ sun ( λ , Ω , T amb , α , Ω sun ) ε c ( λ , Ω , T ) cos θ d Ω d λ ,
A sun cos θ sun I ~ sun ( λ , Ω sun , T amb , α , Ω sun ) ε c ( λ , Ω sun , T ) d λ , ( 2 b )
P atm ( T , T amb , α ) = I ~ atm ( λ , Ω , T amb , α ) ε c ( λ , Ω , T ) cos θ d Ω d λ ,
P rad ( T , T amb , α , Ω sun ) = d Ω cos θ d λ I ~ rad ( λ , Ω , T , T amb , α , Ω sun ) ε c ( λ , Ω , T ) ,
P non - rad ( T , T amb ) = h c ( T amb T ) ,
ε ideal ( dir ) ( λ , Ω , T , T amb , α , Ω sun ) = 1 2 [ 1 + sgn ( I ~ rad ) ] ,
P rad ( T , T amb , α , Ω sun , θ slope ) = I ~ rad ( λ , Ω , T , T amb , α , Ω sun ) ε c ( λ , Ω , T ) cos θ d Ω d λ ,
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.