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

Sea-surface reflectance factor: replicability of computed values

Open Access Open Access

Abstract

The sea-surface reflectance factor ρ required for the determination of the water- leaving radiance from above-water radiometric measurements is derived from radiative transfer simulations relying on models of the sky-radiance distribution and sea-surface statistics. This work primarily investigates the impact on ρ of various sky-radiance and sea-surface models. A specific replicability analysis, restricted to the 550 nm wavelength, has been performed with the Monte Carlo code for Ocean Color Simulations (so-called MOX) accounting for the measurement geometry recommended in protocols for the validation of satellite ocean color data and commonly applied for operational measurements. Results indicate that the variability of ρ increases with wind speed and reaches the largest values for sun elevations close to the zenith or approaching the horizon. In particular, a variability up to about 2% is observed for wind speeds below 4 ms−1 and sun zenith angles larger than 20°. Finally, the benchmark of the ρ values from this study with those formally determined with the Hydrolight code and widely utilized by the ocean color community, exhibits systematic differences. The source of these differences is discussed and the implications for field measurements are addressed.

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

1. Introduction and background

Ocean color remote sensing requires in situ reference values of the spectral water-leaving radiance ${L_w}$ for the development of bio-optical algorithms, the validation of radiometric data products, or the indirect calibration (i.e., system vicarious calibration) of space sensors [13]. The in situ measurements necessary to compute ${L_w}$ can be acquired with in- or above-water methods. In-water methods determine ${L_w}$ from the regression of the upwelling radiance measured at different depths in the water column [46]. Above-water methods determine ${L_w}$ from sea-viewing measurements of the total radiance ${L_t}$ and estimates of the radiance reflected by the sea surface ${L_r}$ according to

$${L_w}({\theta ,\phi ,\lambda } )= {L_t}({\theta ,\phi ,\lambda } )- {L_r}({\theta ,\phi ,\lambda } )$$
where $\theta $ indicates the zenith angle, and $\phi $ the relative azimuth between the measurement and sun planes (for brevity of notation the dependence on wavelength $\lambda $ is omitted in the following equations).

The reflected radiance ${L_r}$ in any direction $(\theta ,\phi )$ is the result of the reflection by the sea surface of the sky-radiance ${L_{sky}}$ originating from the upper hemisphere including both the direct and the diffuse components (leading to sun- and sky-glint, respectively). By indicating with $r({{\theta^ \ast },{\phi^ \ast } \to \theta ,\phi } )$ the sea-surface radiance reflectance function, i.e., the fraction of incident radiance reflected by the sea surface from direction $({\theta ^ \ast },{\phi ^ \ast })$ to direction $(\theta ,\phi ),\,{L_r}$ is given by

$${L_r}({\theta ,\phi } )= \int_0^{2\pi } {\int_0^{\frac{\pi }{2}} {r({{\theta^ \ast },{\phi^ \ast } \to \theta ,\phi } ){L_{sky}}({{\theta^ \ast },{\phi^ \ast }} )\sin {\theta ^ \ast }d{\theta ^ \ast }} } d{\phi ^ \ast }.$$
Current protocols [7,8] to determine ${L_w}({\theta ,\phi } )$ from in situ data rely on: 1) direct measurements of ${L_t}({\theta ,\phi } )$ with the sensor looking in the sea-viewing direction $(\theta ,\phi )$, and 2) measurements of the sky radiance ${L_i}({\theta^{\prime},\phi } )$ in the direction $({\theta^{\prime},\phi } )$ to estimate ${L_r}({\theta ,\phi } )$ as
$${L_r}({\theta ,\phi } )= \rho {L_i}({\theta^{\prime},\phi } )$$
where
$$\rho = \frac{{\int_0^{2\pi } {\int_0^{\frac{\pi }{2}} {r({{\theta^ \ast },{\phi^ \ast } \to \theta ,\phi } ){L_{sky}}({{\theta^ \ast },{\phi^ \ast }} )\sin {\theta ^ \ast }d{\theta ^ \ast }} } d{\phi ^ \ast }}}{{{L_i}({\theta^{\prime},\phi } )}}$$
is the sea-surface reflectance factor. Note that the sea-surface radiance reflectance function r is an inherent optical property because it only depends on the Fresnel reflectance and the sea-surface wave statistics. Conversely, the sea-surface reflectance factor $\rho$ is an apparent optical property because of its dependence on the sky-radiance distribution.

The theoretical values of the reflected radiance Lr and of the incident radiance Li, needed to compute the sea-surface reflectance factor through Eq. (2) and Eq. (4), are derived from sea-surface and the sky-radiance models [911]. It is emphasized that Monte Carlo (MC) ray-tracing simulations permit to compute $\rho $ accounting for multiple reflections at the sea surface [1217].

Investigating the replicability of $\rho $ as a function of various sky-radiance and sea-surface models is the primary objective of this study. The Monte Carlo code for Ocean Color Simulations (so-called MOX) is utilized [18,19] restricting the analysis to clear sky conditions and accounting for the spatially averaged variability of sea-surface statistics.

The considered measurement geometry is that recommended by protocols for the validation of satellite ocean color data [7] and applied by the Ocean Color component of the Aerosol Robotic Network (AERONET-OC) [20]. As illustrated in Fig. 1, the measurement angles are $\phi = 90^\circ $ with respect to the sun azimuth, $\theta ^{\prime} = 40^\circ $ with respect to the zenith and $\theta = 180^\circ{-} 40^\circ $. Alternative geometries could certainly be considered and might provide relative advantages. For instance, choosing $\phi = 135^\circ $ instead of $\phi = 90^\circ $ might further minimize glint perturbations, but it would concurrently increase measurement perturbations induced by the deployment structure, which increase with the sun zenith.

 figure: Fig. 1.

Fig. 1. (a) Perspective-, (b) side- and (c) top-view of typical above-water measurement geometry. Note that the sea- and sky-viewing angles are defined with respect to the zenith.

Download Full Size | PDF

Look-up-tables of $\rho $ (e.g., [12,15]) allow deriving ${L_\textrm{w}}({\theta ,\phi } )$ from operational measurements of ${L_\textrm{t}}({\theta ,\phi } )$ and ${L_i}({\theta^{\prime},\phi } )$. The determination of ${L_\textrm{w}}({\theta ,\phi } )$ from in situ data presents, however, some caveats. In particular, Lt and Li measurements exhibit dependence on sensor features such as field-of-view and sampling time. Additionally, Li measurements from the direction $({\theta^{\prime},\phi } )$ might not accurately represent the ${L_{sky}}$ contributions leading to ${L_r}$. Consequently, ${L_\textrm{w}}$ determined from field measurements through Eq. (1) and Eq. (3) implies dependence on instrument features, measurement geometry, and obviously, sky-radiance distribution ${L_{sky}}$ and sea-surface radiance reflectance function r.

A core component of this study is the direct comparison between $\rho $ values computed with MOX and those included in the widely used $\rho $-tables generated with Hydrolight [12,21]. This benchmark analysis is performed assuming identical sky-radiance distributions and sea-surface statistics, consistently determined at the 550 nm center-wavelength neglecting polarization. Cross-comparisons based on independent data sources [9,10] are also presented.

It is emphasized that this work strictly focuses on: 1) the quantification of the replicability of ρ-factors, and 2) the assessment of processing methods and results through benchmarks. More general issues of relevance for above-water radiometry, such as the impact of measurement variability on the determination of ${L_t}$ values representing the actual observation conditions, the dependence of the sea-surface reflectance factor on wavelength and polarization [22], the evaluation of alternative methods for the reduction of above-water radiometric data [23,24], or even the provision of practical recommendations on measurements protocols, are all beyond the study objectives.

The structure of the work is as follows. Section 2 describes the simulation framework, the sky-radiance and sea-surface models, and the processing scheme to compute ρ. Section 3 documents the ρ-factor replicability and benchmark results. Findings from benchmark analyses and the impact of the study on the reduction of field measurements for the determination of ${L_w}$ are discussed in Section 4. Summary and remarks in Section 5 complete the work.

2. Methods

2.1 Simulation framework

A discretization of the upper hemisphere into quadrilaterals, called quads, is used in Hydrolight to perform MC ray tracing, collect simulation results and compute $\rho $. To simplify comparisons, MOX adopts the same scheme as Hydrolight. The partition accounts for ${N_\theta } \times {N_\phi }$ intervals in the zenith and azimuth direction, respectively (see Fig. 2). The i-th increment in the zenith direction is $\Delta {\theta _i}$, for $i = 1, \ldots ,{N_\theta }$ with $\mathop \sum \nolimits_i^{{N_\theta }} \Delta {\theta _i} = \pi /2$ (i.e., the $\mathrm{\Delta }{\theta _i}$ values can be different). The reference discretization is ${N_\theta } = 10$ with $\mathrm{\Delta }{\theta _i} = 5^\circ $ if i=1 or i=10, and $\mathrm{\Delta }{\theta _i} = 10^\circ $ otherwise. The j-th increment in the azimuth direction is $\mathrm{\Delta }{\phi _j} = 2\pi /{N_\phi }$ (i.e., the $\mathrm{\Delta }{\phi _j}$ segments are equal) for $N_\phi = 24$ and $\mathrm{\Delta }\phi = 15^\circ $. Grid points have zenith coordinates ${\theta _i} = \mathop \sum \nolimits_{i^{\prime} = 1}^{i - 1} \Delta {\theta _{i^{\prime}}}$, with ${\theta _1} = 0^\circ $ and azimuth coordinates ${\phi _j} = \Delta \phi \cdot ({j - 1/2} )$ for $j = 0, \ldots ,{N_\phi }$. Hydrolight unifies in a single polar cap the 24 slices converging to $\theta = 0^\circ $, while MOX keeps them separated (which is equivalent in terms of $\rho $ computation).

 figure: Fig. 2.

Fig. 2. Partitioning of the upper hemisphere into quads. The thick solid and dashed lines indicate the measurement and the solar plane, respectively. The sensor-quad marked in black represents the virtual radiometer.

Download Full Size | PDF

To define a correspondence between field measurements and the simulation framework, the sun azimuth ${\phi _0}$ is set to $0^\circ $ with the solar plane crossing the hemisphere from $\phi = 0^\circ $ to $\phi = 180^\circ $. The measurement plane defined by the radiometer azimuth is determined by $\phi = 270^\circ $ and $\phi = 90^\circ $ (note that the assumption of ${\phi _0}$ = $0^\circ $ implies the equivalence of the relative azimuth $\phi $ with the azimuth identified by the measurement plane). The quad including $\theta = 40^\circ $ and $\phi = 270^\circ $ represents the virtual radiometer and is denoted as the sensor-quad.

Following the scheme implemented in Hydrolight [21], the sky radiance assumed homogeneous within each quad, is computed with the models presented in Sec. 2.2 for directions $({\theta_i^\ast ,\phi_j^\ast } )$, where $\theta _i^\ast\,{=}\, {\cos ^{ - 1}}({0.5 \cdot ({\cos {\theta_{i - 1}} + \cos {\theta_i}} )} ),\,\,\phi _j^\ast\,{=}\, 0.5 \cdot ({{\phi_j} + {\phi_{j + 1}}} )$, and (${\theta _i},\,{\phi _j}$) indicate the hemisphere grid coordinates. The quad including the sun zenith ${\theta _0}$ and sun azimuth ${\phi _0}$ angles is referred to as the sun-quad. The sea surface below the sky hemisphere permits simulating the light reflection according to the model representing the statistical properties of waves at the air-water interface (Sec. 2.4). Specific schemes adopted by Hydrolight and MOX for MC ray tracing are presented in Sec. 2.5 while detailing methods for the computation of $\rho $.

2.2 Sky-radiance models

The angular distribution of the sky radiance ${L_{sky}}$ is simulated with: 1) the Harrison-Coombes parametric model (HC) [2527] embedded in the Hydrolight code, 2) the analytical model applied by Zibordi and Voss (ZV) providing realistic distributions in clear sky conditions [28], and 3) the highly accurate Finite Element Method radiative transfer numerical code (FEM) matter of several benchmarks and applications [2932]. These methods are briefly introduced in the following sub-sections together with details on their implementation. The irradiance model from Gregg and Carder (GC) [33] is additionally applied to rescale the diffuse sky irradiance in HC and determine the direct radiance component for both HC and ZV. The main quantities necessary for the implementation of the various modeling solutions are summarized in Table 1 specifying their relevance for each scheme (i.e., HC, ZV and FEM).

Tables Icon

Table 1. Main quantities applied for the simulation of sky-radiance distribution.

In view of providing absolute terms of comparison, simulated sky-radiance data are presented in radiance units (i.e., W m-2 nm-1 sr-1). It is however recognized that, in agreement with Mobley [12], relative units could be alternatively considered for the computation of $\rho $ with the sole requirement to respect the ratio between direct and diffuse irradiance components.

2.2.1 Harrison-Coombes model (HC)

The HC model parameterizes the diffuse component of the sky radiance through a non-linear least-squares regression of a set of 300 measurements according to the scheme proposed by Pokrowski [39] and Kittler [40]. These measurements are integrated radiance values in the 300–3000 nm spectral range. The HC wavelength-independent parametric formulation of the diffuse sky-radiance component $L_{diff}^{HC}({\theta ,\phi } )$ for any direction $({\theta ,\phi } )$ of the sky dome and sun zenith ${\theta _0}$ is given by

$$\begin{aligned} L_{diff}^{HC}({\theta ,\phi } ) & = ({1.63 + 53.7{e^{ - 5.49\Psi }} + 2.04\,{{\cos }^2}(\Psi )\cos ({{\theta_0}} )} )\\ & \times ({1 - {e^{ - 0.19\sec (\theta )}}} )({1 - {e^{ - 0.53\sec ({{\theta_0}} )}}} ), \end{aligned}$$
where the scattering angle $\mathrm{\Psi }$ is defined as
$$\cos (\mathrm{\Psi } )= \cos (\theta)\cos ({{\theta_0}} )+ \sin (\theta)\sin ({{\theta_0}} )\cos (\phi )$$
with ${\phi _0} = 0^\circ$ in agreement with the simulation geometry defined in Fig. 2.

The determination of spectral values for the direct and diffuse sky-radiance components follows the scheme used in Hydrolight [21]. Specifically, GC is applied to compute: 1) the direct sun irradiance $E_{dir}^{CG}({\theta _0})$ necessary to determine the direct sun radiance $L_{dir}^{HC}({\theta _0})$ = $E_{dir}^{CG}({\theta _0})/{\Omega _0}$, where ${\Omega _0}$ is the solid angle of the sun-quad, and 2) the diffuse sun irradiance $E_{diff}^{CG}({\theta _0})$ necessary to compute the scaling factor ${S^{HC}}({\theta _0}) = {{E_{diff}^{CG}({\theta _0})} / {E_{diff}^{HC}}}({\theta _0})$, with $E_{diff}^{HC}({\theta _0}) = \int_0^{2\pi } {\int_0^{\frac{\pi }{2}} {L_{diff}^{HC}({\theta ,\phi } )\sin \theta \cos \theta d\theta } } d\phi$. The HC total sky radiance is then determined as $L_{sky}^{HC}({\theta ,\phi } )= L_{diff}^{HC}({\theta ,\phi } )\cdot {S^{HC}}({\theta _0})$ for any $({\theta ,\phi } )$ different from $({{\theta_0},0} )$, and as $L_{sky}^{HC}({{\theta_0},0} )= L_{diff}^{HC}({{\theta_0},0} )\cdot {S^{HC}}({\theta _0}) + L_{dir}^{HC}({{\theta_0}} )$ for $({{\theta_0},0} )$.

2.2.2 Zibordi-Voss model (ZV)

The ZV model for the computation of the diffuse sky radiance relies on the analytical solution of the radiative transfer equation (RTE) proposed by Sobolev for a plane-parallel, homogeneous and cloudless atmosphere, bounded by a Lambertian surface [41].

The expression of the diffuse sky radiance for any direction $({\theta ,\phi } )$ of the sky dome and sun zenith ${\theta _0}$ is given by

$$L_{diff}^{ZV}({\theta ,\phi ,\lambda } )= \frac{{{E_0}(\lambda )\cdot {d^2}}}{\pi }\sigma ({{\tau_d}(\lambda )} )\cos ({{\theta_0}} )\exp \left[ {\frac{{{\tau_a}(\lambda )}}{{\cos ({{\theta_0}} )}}} \right]$$
where ${E_0}(\lambda )$ is the mean solar spectral extra-atmospheric irradiance and ${d^2}$ the sun-earth distance correction factor function for the day of the year D; ${\tau _d}(\lambda )$ indicates the scattering optical depth resulting from the sum of the optical depths of non-absorbing (Rayleigh) molecules ${\tau _R}(\lambda )$ and aerosols ${\tau _A}(\lambda );\,\tau_a(\lambda )$ indicates the sum of the optical depths of the absorbing atmospheric constituents comprising ozone $\tau_O(\lambda)$, water vapor $\tau_W(\lambda)$ and uniformly mixed-gases $\tau_G(\lambda)$, with the condition of weak absorption (for the purpose of this work both $\tau_W(\lambda)$ and $\tau_G(\lambda)$ are nill).

The term $\sigma(\tau_d(\lambda))$ is the scattering transmission of the atmosphere given by

$$\begin{aligned} \sigma ({{\tau_d}(\lambda )} )&= \frac{{({1 - \gamma (\lambda )} )R({\lambda ,\theta } )+ 2\gamma (\lambda )R({\lambda ,{\theta_0}} )}}{{4 + ({3 - {P_1}} )({1 - \gamma (\lambda )} ){\tau _d}(\lambda )}}\\ & - \frac{1}{2}\left( {\exp \left\{ {\frac{{ - {\tau_d}(\lambda )}}{{\cos (\theta )}}} \right\} + \exp \left\{ {\frac{{ - {\tau_d}(\lambda )}}{{\cos ({{\theta_0}} )}}} \right\}} \right)\\ & + [{\beta (\Psi )- ({3 + {P_1}} )\cos (\theta )\cos ({{\theta_0}} )} ]{\sigma _1}({{\tau_d}(\lambda )} )\end{aligned}$$
that includes the spectral albedo of the sea surface $\gamma(\lambda)$, the first term ${P_1}$ in the Legendre polynomial expansion of the global atmospheric phase function $\beta (\mathrm{\Psi } )$ combining molecular and aerosol scattering [42], the single scattering transmission ${\sigma _1}({{\tau_d}(\lambda )} )$ and the auxiliary function $R(\lambda,\theta )$ [28]. The spectral albedo is assumed independent from the sun zenith and set to 0.06. This approximation is expected to be representative of the average simulation conditions considered in this work [43].

The molecular scattering is described by the Rayleigh phase function, while the aerosol scattering phase function is approximated by the two-term Henyey-Greenstein (TTHG) analytical function [44] with asymmetry parameters ${g_1} = 0.713,\,{g_2} = 0.759$, and $a = 0.962$ proposed for maritime aerosols [42].

Recalling that ZV only provides the diffuse component of the spectral sky radiance, the determination of the spectral direct radiance component also benefits from the diffuse and direct irradiances computed with GC. Specifically, assuming equality of the GC and ZV total irradiances, the ZV direct sun irradiance component is determined as $E_{dir}^{ZV}({{\theta_0}} )= E_{dir}^{GC}({{\theta_0}} )+ E_{diff}^{GC}({{\theta_0}} )- E_{diff}^{ZV}({{\theta_0}} )$ where the explicit dependence on $\lambda $ has been omitted for simplicity, and $E_{diff}^{ZV}({\theta _0}) = \int_0^{2\pi } {\int_0^{\frac{\pi }{2}} {L_{diff}^{ZV}({\theta ,\phi } )\sin \theta \cos \theta d\theta } } d\phi$ is the ZV diffuse irradiance for sun zenith angle ${\theta _0}$ implicit in $L_{diff}^{ZV}({\theta ,\phi } )$. The direct radiance is then given by $L_{dir}^{ZV}({{\theta_0}} )$ = $E_{dir}^{ZV}({\theta _0})/{\Omega _0}$, with ${\Omega _0}$ the solid angle of the sun-quad. Computationally, the sky radiance from ZV is given by $L_{sky}^{ZV}({\theta ,\phi } )= L_{diff}^{ZV}({\theta ,\phi } )$ for any $({\theta ,\phi } )$ different from $({{\theta_0},0} )$, and by $L_{sky}^{ZV}({{\theta_0},0} )\,= \, L_{diff}^{ZV}({{\theta_0},0} )+ L_{dir}^{ZV}({{\theta_0}} )$ for $({{\theta_0},0} )$. It is finally noted that $L_{sky}^{ZV}({{\theta_0},0} )$ cannot be computed for ${\theta _0}$ and consequently it can only be approximated by values determined nearby the sun zenith angle. This singularity, however, only affects the determination of the diffuse sky radiance in the sun direction and does not represent a limitation to the present work. In fact, by adopting the same scheme utilized in Hydrolight, the sun is located at the center of the quad, while the diffuse sky radiance is computed for the direction $({\theta_i^\ast ,\phi_j^\ast } )$, as illustrated in Section 2.1.

2.2.3 FEM numerical code

The Finite Element Method radiative transfer code (FEM) [29], applies the finite element deterministic approach to model the propagation of unpolarized monochromatic radiation in the ultraviolet, visible and near-infrared electromagnetic regions within a vertically inhomogeneous atmosphere-ocean system. It fully accounts for multiple scattering, reflection and refraction at the air-water interface, and has been extensively tested and validated through benchmark with other popular codes [2932].

The atmosphere-ocean system is divided into ${\rm {\frak L}}$ optically homogeneous horizontal layers. For each layer, the phase function is expanded in Legendre polynomials with a finite number of terms M, and the radiance L is correspondingly represented by a Fourier cosine series

$$L({\tau ;\eta ,\phi } )= \mathop \sum \limits_{m = 0}^M {L^m}({\tau ,\eta } )\textrm{cos}m\phi $$
This allows representing the RTE with M+1 independent equations. The resulting equation for the l-th layer and the m-th Fourier radiance component becomes
$$\eta \frac{{d{L^m}({\tau ,\eta } )}}{{d\tau }} ={-} {L^m}({\tau ,\eta } )+ \frac{{{\omega _{0l}}}}{2}\int\limits_{ - 1}^1 {{L^m}({\tau ,\eta^{\prime}} ){p_l}^m({\tau ,\eta^{\prime} \to \eta } )d\eta ^{\prime}} + \textrm{Q}_l^m({\tau ,\eta } )$$
where τ is the optical depth, $\eta \textrm{ } = \textrm{ cos}\theta ,\,{\omega _{0l}}$ is the single scattering albedo, $p_l^m$ is the m-th Fourier component of the global phase function and $\textrm{Q}_l^m$ represents the source function, whose expression differs between atmosphere and water.

The remaining integrals are further approximated by projecting each radiance harmonic ${L^m}\; $onto a set of basis functions, each different from zero only in a finite interval (hence, the finite element designation of this method). Specifically, the solution of Eq. (10) is sought in the form

$${L^m}({\tau ,\eta } )= \mathop \sum \limits_{j = 1}^{2N} L_j^m(\tau )b_j^m(\eta )$$
with
$$b_j^m(\eta )= Y_j^m(\eta )\left\{ \begin{array}{l} 1 - \frac{{|{\eta - {\eta_j}} |}}{h}\; \; \; \; \; \textrm{if }\; |{\eta - {\eta_j}} |< h\\ 0\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \textrm{elsewhere} \end{array} \right.$$
where $Y_j^m$ are the spherical harmonics, ${\eta _j}\textrm{ }({j = 1,\ldots ,2N} )$ are the nodes of the homogeneous grid in the interval [-1,1], and $h = 1/(N - 1)$ is the grid step.

The resulting system of matricial equations is solved analytically, imposing the necessary boundary conditions, and the solutions are lastly improved by applying an interpolation procedure.

It is highlighted that FEM ensures flux conservation independently of the number of grid points (N = 2 is already sufficient) and of the shape of the phase function (i.e., of the order of polynomials M used to correctly represent it), opposite to alternative numerical methods (e.g., the Discrete Ordinate Method), which require that the number of Gaussian quadrature points N used to solve the RTE integral is at least equal to the number of Legendre polynomials M. Additionally, by selecting basis functions which are finite, FEM generates a piecewise approximate solution that may converge faster than other global approximations (e.g., those adopted in the Spherical Harmonic Method). A number of grid points N = 64 is usually sufficient to accurately model the radiance angular distribution even when characterized by extreme gradients. The approach implemented in FEM is thus particularly suitable for highly asymmetric phase functions, like those characterizing the scattering by aerosols and hydrosols.

For the present work, the atmosphere is divided into 14 plane-parallel layers (i.e., ${\rm {\frak L}}$=14) resolving the vertical distribution of aerosol, ozone, and other gas molecules. Each atmospheric layer contains a variable mixture of ozone (only absorbing), other gas molecules (only scattering), and aerosols (scattering and absorbing). The vertical profile of ozone and Rayleigh molecules are determined in agreement with Guzzi et al. [45], and Marggraf and Griggs [46], respectively. The vertical profile of aerosols is modeled according to Elterman [47] assuming a mean aerosol scale height of 1.2 km. The molecular and aerosols scattering phase functions are defined as in ZV.

2.3 Simulated sky-radiances

Examples of diffuse sky-radiance distributions obtained with HC, ZV and FEM, are displayed in Fig. 3 for different ${\theta _0}$ (i.e., $20^\circ ,\,40^\circ $ and $60^\circ $). These radiance distributions show differences among the various models that become more pronounced with increasing ${\theta _0}$. This is confirmed by the diffuse sky-radiance values shown in Fig. 4 for the solar and measurement planes (solid and dashed lines, respectively), with HC exhibiting the largest deviations with respect to both ZV and FEM, and the latter being regarded as the most accurate approach. These differences are explained by the diverse capability of the applied models to account for forward and multiple scattering, more challenged in the circumsolar region and near the horizon, respectively.

 figure: Fig. 3.

Fig. 3. Sky-radiance distribution Lsky at 550 nm (restricted to the diffuse component) obtained with HC, ZV and FEM displayed from the top to the bottom row, respectively. Panels from left to right correspond to ${\theta _0}$ = 20°, 40° and 60°.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Sky radiance Lsky at 550 nm (restricted to the diffuse component) determined for the solar and measurement planes (solid and dashed lines, respectively) with HC, ZV and FEM for different ${\theta _0}$ (i.e., $20^\circ ,\,40^\circ $ and $60^\circ $). Dots highlight the zenith angles at which the sky-radiance values are computed (see Sec. 2.1). The direct radiance components from the sun-quad at $20^\circ ,\,40^\circ $ and $60^\circ $ (not displayed) are respectively 77.03, 37.32 and 21.16 $\textrm{W }{\textrm{m}^{ - 2}}\textrm{n}{\textrm{m}^{ - 1}}\textrm{s}{\textrm{r}^{ - 1}}$ for both HC and ZV. Corresponding FEM values are 77.13, 37.36 and 21.15 $\textrm{W }{\textrm{m}^{ - 2}}\textrm{n}{\textrm{m}^{ - 1}}\textrm{s}{\textrm{r}^{ - 1}}$.

Download Full Size | PDF

Among the three sky-radiance models applied, FEM ensures highly accurate spectral and angular distributions of the diffuse radiance under any condition. This includes an accurate determination of the sky radiance from both the circumsolar region (mostly affected by a correct treatment of the aerosol forward scattering) and from the horizon (mostly affected by a correct treatment of multiple scattering effects). This further implies accurate estimates of both diffuse and direct irradiances. Conversely, HC exhibits intrinsic limitations due to its wavelength-independent parametrization of the diffuse sky-radiance distribution through Eq. (5). Finally, ZV may provide realistic spectral sky-radiance distributions whose accuracy is expected to be higher when the multiple scattering contribution is lower (i.e., for low values of the aerosol optical depth).

2.4 Sea-surface models

The effect of the sea-surface statistics on $\rho $ replicability is analyzed accounting for: 1) the Cox-Munk sea-surface model (CM) [48], and 2) the Elfouhaily, Chapron, Katsaros and Vandemark sea-surface model (ECKV) [49]. Additional models (e.g., [50,51]) were not considered by restricting the analysis to statistical solutions either already adopted in the determination of the ρ tables currently utilized for the reduction of field data (CM), or characterized by a better capability to represent the sea-surface slope and elevation distributions (ECKV). The values of wind-speed ${v_w}$ considered in the present study vary from 0 to 10 $\textrm{m}{\textrm{s}^{ - 1}}$ with 2 $\textrm{m}{\textrm{s}^{ - 1}}$ increments.

2.4.1 Cox-Munk Sea-surface model (CM)

The CM equations express the slope distribution of surface waves as a Gaussian function with variance components along ($\mathrm{\sigma }_{\textrm{up}}^2$) and across ($\mathrm{\sigma }_{\textrm{cr}}^2$) wind direction given by

$$\mathrm{\sigma }_{\textrm{up}}^2 = 0.00316 \cdot {v_w}$$
$$\mathrm{\sigma }_{\textrm{cr}}^2 = 0.003 + 0.00192 \cdot {v_w}$$
where ${v_w}$ represents the wind velocity in units of $\textrm{m}{\textrm{s}^{ - 1}}$. The sea-surface elevation distribution is not considered. Assuming independence from the wind direction, as in [12], the overall slope variance is
$$\mathrm{\sigma }_{\textrm{slp}}^2 = 0.003 + 0.00512 \cdot {v_w}$$
The tables of $\rho $-factors computed with Hydrolight [12] and extensively applied to derive the water-leaving radiance through Eq. (1) and Eq. (3) are determined with the following formulation of the sea-surface slope variance
$$\mathrm{\sigma }_{\textrm{slp}}^2 = 0.00508 \cdot {v_w}$$
obtained from the work of Cox and Munk excluding the constant background term [52]. Sea surfaces generated with Eq. (15) and (16), hereafter identified as CM1 and CM2, respectively, are both considered in this study. Examples of sea surfaces generated with the CM2 model are presented in panels (a) and (b) of Fig. 5.

 figure: Fig. 5.

Fig. 5. CM2 (top panels) and ECKV (bottom panels) representations of the sea surface. The left and right columns refer to ${v_\textrm{w}}$ of 3 and 7 $\textrm{m}{\textrm{s}^{ - 1}}$, respectively.

Download Full Size | PDF

2.4.2 ECKV sea-surface model

The ECKV sea-surface model is based on the distribution of waves spectral energy. The elevation $z({i,j} )$ is expressed at the grid-points of the sea surface, discretized with ${N_x}$ and ${N_y}$ intervals along the x- and y-axis, respectively, under the assumption of linear wave theory, applying [15]

$$z({i,j} )= \frac{1}{{{N_x}{N_y}}}\mathop \sum \nolimits_{u = 1}^{{N_x}} \mathop \sum \nolimits_{v = 1}^{{N_y}} \hat{z}({u,v} ){e^{i({r \cdot u/{N_x} + s \cdot v/{N_y}} )}}$$
where the harmonic components $\hat{z}({u,v} )$ are indexed by u and v in the wavenumber k domain. The $\hat{z}({u,v} )$ values are set based on the ECKV omnidirectional spectral density ${\cal \textrm{S}}(k )$ [15,49] to express the variance of the sea-surface elevation and slope as a function of wind speed.

Processing steps for generating the sea surface require: 1) computing the two-sided discrete values of the elevation variance, 2) sampling the amplitudes of the harmonic components with normal distributions, and 3) defining the Inverse Fast Fourier Transform Hermitian coefficients to obtain the sea-surface elevation. The spectral density function ${\cal \textrm{S}}(k )$ is iteratively adjusted until converging to the target value of the waves slope distribution [17]. In the present work, the propagation direction of waves is assumed isotropic. Sea-surface representations generated with the ECKV model are illustrated in panels (c) and (d) of Fig. 5.

2.5 Computation of the sea-surface reflectance factor

The partition of the upper hemisphere into quads presented in Sec. 2.1 is here used to define the values of a radiance reflectance matrix ${\cal {R}}$ towards the sensor-quad (i.e., the quad at $\theta = 40^\circ$ and $\phi = 270^\circ $—see Fig. 2), so that

$${L_r} = \mathop \sum \nolimits_{{i^\ast }}^{{N_\theta }} \mathop \sum \nolimits_{{j^\ast }}^{{N_\phi }} {\cal {R}}({{i^\ast },{j^\ast }} ){L_{sky}}({{i^\ast },{j^\ast }} ).$$
The $\rho $ value is computed as
$$\rho = \frac{{{L_r}}}{{{L_i}}},$$
where ${L_i}$ corresponds to the ${L_{sky}}$ value of the quad at $\theta = 40^\circ $ and $\phi = 90^\circ $. The different MC schemes utilized in Hydrolight and MOX to derive ${\cal {R}}({{i^\ast },{j^\ast }} )$ are presented in Sec. 2.5.1 and 2.5.2, respectively. All simulations rely on the refractive index of the water ${n_w} = 1.340$.

2.5.1 Hydrolight scheme

Hydrolight adopts forward ray tracing from the source-quad to the sea surface. Rays are initiated with unitary statistical weight. One quad at a time is selected as a source until the entire hemisphere is considered. The sea surface is represented by a hexagonal domain divided into triangular facets, called triads, tilted according to the Cox-Munk waves slope distribution (CM2, Eq. (16); see also Fig. 6). The Snell and Fresnel equations are applied to update the weight and direction of rays hitting the sea surface. To enhance the computing efficiency (an essential feature when the algorithm was conceived [53]), rays are forced to converge to the center of the hexagon as illustrated in Fig. 6(b). Rays initiated from a quad close to the horizon might be reflected by a triad located between the starting and the target points (the latter corresponding to the central triad) without reaching the hexagon center. This occlusion results in a phenomenon usually termed as “shadowing effect”, which nonetheless affects only a small number of rays.

 figure: Fig. 6.

Fig. 6. Illustration of the Hydrolight ray-tracing scheme. (a) Rays initialization from a random position of each emitting quad. (b) Perspective view, where dashed and solid lines illustrate incident rays and their reflection towards the sensor-quad, respectively. Rays reflected out of the sensor-quad are not shown. Each ray corresponds to a different hexagon realization (only one surface is however displayed).

Download Full Size | PDF

For each quad, the accumulated weight of the rays reflected toward the sensor-quad is divided by the number of emitted rays to determine the ratio $\Phi ({{i^\ast },{j^\ast }} )$ between reflected and incoming radiant fluxes. The process is repeated for several hexagon realizations by resampling the inclination of all triads to comprehensively represent the overall statistical properties of the sea surface. The computed average flux ratios $\bar{\mathrm{\Phi }}({{i^\ast },{j^\ast }} )$ are then transformed into radiance ratios (and hence elements of the reflectance matrix ${\cal {R}}$) by accounting for the geometry and solid angles of the emitting and receiving quads (for details, see [21]). Note that surface reflectance and transmission properties are inherent optical properties of the sea surface, parameterized as a function of the wind speed. Their values are pre-computed with the ray-tracing scheme formerly detailed and are not generated for each Hydrolight simulation.

2.5.2 MOX scheme

The MOX code relies on backward MC starting the rays from the sensor-quad, see Fig. 7(a). The sea surface is defined over a square domain of side $L = 50$ m discretized in $N \approx 500$ intervals along the x- and y-axis [17]. The spatial resolution and coordinates along the x-axis are ${\Delta _x} = L/{N_x}$ and $x(i )= i \cdot {\Delta _x}$, with $i = 0, \ldots ,{N_x}$. Equivalent quantities for the y-axis are ${\Delta _y} = L/{N_y}$ and $y(j )= j \cdot {\Delta _y}$, for $j = 0, \ldots ,{N_y}$. Each $({{\Delta _x},{\Delta _y}} )$ element is split into two triads whose vertices are elevated based on the CM1, CM2 or ECKV sea-surface model (Sec. 2.4).

 figure: Fig. 7.

Fig. 7. As in Fig. 6, but for the MOX backward ray-tracing scheme. The solid and dashed lines denote rays emitted from the sensor-quad and reflected by the sea surface, respectively. Rays that exit the domain from one side continue their path at the opposite side based on the periodicity of the sea surface.

Download Full Size | PDF

Rays are initiated at random horizontal positions from the same height above the sea surface—see Fig. 7(b). Those exiting the domain from one side continue their paths at the opposite side based on the periodicity imposed at the sea surface boundaries. The weight of the reflected ray is determined with the Fresnel equation. Upon the last reflection, the ray weight is accumulated at the hemisphere quads following the same scheme described for Hydrolight. Namely, the result of the weight accumulation, scaled by the number of emitted rays, represents the radiant flux ratio for the considered sea-surface model and wind speed. The MC simulations are repeated by resampling ten times the triads slope to compute the mean radiant flux ratio $\bar{\mathrm{\Phi }}({{i^\ast },{j^\ast }} )$ to better represent the sea-surface statistical properties (see Fig. 8). Considering that ray-tracing is performed in backward mode, ${\cal {R}}({{i^\ast },{j^\ast }} )= \bar{\mathrm{\Phi }}({{i^\ast },{j^\ast }} )$ as a result of the reciprocity principle [54].

 figure: Fig. 8.

Fig. 8. Slope variance as a function of the wind speed for each sea-surface model addressed in the study (i.e., CM1, CM2 and ECKV labeled with a circle, square and triangle, respectively). The solid lines connecting the symbols represent the expected theoretical values. The mean and the standard deviation of the slope variance resulting from multiple sea-surface realizations are displayed as dashed lines and shaded regions, respectively. Overlapping lines confirm the satisfactory agreement between expected and computed trends.

Download Full Size | PDF

The ray-tracing executed in this work to analyze the $\rho $-factor replicability considers 3 different sea-surface models and 6 wind-speed values. As anticipated, each sea surface, made of approximatively $5 \cdot {10^5}$ triads, is resampled 10 times. Individual MC simulations account for ${10^7}$ rays, and for each it is necessary to find the target triad along the ray trajectory, and then determine the intersection point. The overall possible number of ray-triad intersections is then approximately ${10^{15}}$. MOX simulations were carried out using the Navigator supercomputer of the University of Coimbra, Portugal. The system features 164 identical computing nodes equipped with 2 twelve-core Intel Xeon E5-2697v2 processors at 2.70 GHz, 96 GB RAM and InfiniBand interconnect. Multiple simulation cases were executed simultaneously using 24 CPU cores in parallel. This allowed running all simulation cases in about 15 minutes. Examples of radiance reflectance matrix obtained with the MOX code are presented in Fig. 9.

 figure: Fig. 9.

Fig. 9. Examples of radiance reflectance matrix based on the CM1, CM2 and the ECKV sea-surface statistics in the top, middle and bottom row panels, respectively. Column panels from left to right correspond to vw = 2, 6 and 10 ms-1, respectively. The solar plane crossing the hemisphere from $\phi = 0^\circ $ to $\phi = 180^\circ $and the radiometer orientation from $\phi = 270^\circ $ to $\phi = 90^\circ $are indicated by the dashed line and the arrow, respectively.

Download Full Size | PDF

3. Results

3.1 Replicability of the sea-surface reflectance factor

The mean and standard deviation of $\rho $ values are analyzed in Fig. 10 for different combinations of sky-radiance (HC, ZV, and FEM) and sea-surface models (CM1, CM2 and ECKV). The radiance reflected at the sea surface accounts cumulatively for the contribution originating from the sun-quad (S) and that from all other quads (D). The total $\rho $ value (where $\rho = {\rho _D} + {\rho _S}$) is then examined in terms of ${\rho _S}$ and ${\rho _D}$, respectively. The two components of $\rho $ are modeled through Eq. (18) and (19) accounting for the entry of the radiance reflectance matrix corresponding to the sun-quad.

 figure: Fig. 10.

Fig. 10. Variability of $\rho $ values as a function of wind speed, as resulting from the application of various models for sky-radiance distribution at 550 nm (first row of panels) and sea-surface statistics (middle row of panels), and the combined effects of sky-radiance and sea surface (bottom row of panels). Column panels from left to right report results for ${\theta _0}$ = 20°, 40° and 60°, respectively. The mean and the standard deviation of $\rho $ values are represented with a trend line and a shaded region, respectively.

Download Full Size | PDF

The capability of separating $\rho $ into its direct- and diffuse-related components, and additionally discussing the dependence of the two with wind speed, is of interest for above-water applications. In this respect, it is noted that the sun-quad has dimensions larger than the sun disc. As such, the glint contribution of the sky radiance originating from the sun-quad with respect to the case where the exact size of the sun disc is considered, translates into a larger number of rays characterized by a lower weight. In terms of operational above-water radiometry, the impact of this approximation is indeed lessened by the application of measurement geometries minimizing the sun-glint contribution. In fact, the contribution of ${\rho _S}$ is the most difficult to be statistically represented in field measurements and, for this reason, its perturbing effects are minimized through the adoption of suitable viewing geometries.

The HC and CM2 are the sky-radiance and sea-surface reference models applied in Hydrolight. These models are here considered as baselines to investigate the replicability of $\rho $ values. Namely, panels in the top row of Fig. 10 refer to the same sea-surface model (CM2) to detail the variability of $\rho $ induced by different sky-radiance models. Conversely, panels in the middle row rely on the same sky-radiance distribution (HC) to illustrate the effects due to different sea-surface representations. These results hence permit to document the specific sensitivity of the $\rho $ factor to the selection of the sky-radiance model (top-row panels) or of the sea-surface model (middle-row panels). Panels in the bottom row finally show the variability obtained when considering all sky-radiance and sea-surface models. Panels in the left, central and right columns refer to ${\theta _0} = 20^\circ ,40^\circ \textrm{ and }60^\circ$, respectively (results are presented adopting different scales to better visualize trends).

The solid lines and the shaded regions in Fig. 10 represent the mean and the standard deviation of different factors. The scrutiny of results indicates that the mean value of $\rho $ is driven by two distinct trends of ${\rho _D}$ and ${\rho _S}$ as a function of ${v_w}$. An increase approximately proportional to ${v_w}$ characterizes ${\rho _D}$ for the considered wind-speed interval. This trend displays minor variations among different sky-radiance and sea-surface models. Instead, at low ${\theta _0}$ and for ${v_w}$ above 4 ms-1, ${\rho _\textrm{S}}$ grows rapidly with ${v_w}$ showing an exponential-like tendency as a result of increasing contributions to Lr of the reflected sun-quad radiance. This effect reduces with large ${\theta _0}$ leading to less pronounced sun glint perturbations.

As displayed in Fig. 10(a), both ${\rho _D}$ and ${\rho _S}$ show dependence on the selected sky-radiance model at high sun elevations (e.g., ${\theta _0} = 20^\circ $). With reference to ${\rho _S}$, its variability is mainly ascribed to the different sky-radiance values in the circumsolar region, where the HC ones are significantly larger than those from ZV and FEM (see Fig. 4). For the same sun elevations, Fig. 10(d) shows that ${\rho _\textrm{D}}$ is almost unaffected by the choice of the sea-surface model. The variability of ${\rho _\textrm{S}}$ appears instead similar to that documented in Fig. 10(a), although the mean value of ${\rho _S}$ is larger in Fig. 10(d) than in Fig. 10(a) for ${v_w}$ = 10 $\textrm{m}{\textrm{s}^{ - 1}}$. This confirms the relevance of the sea-surface slope variance in defining the fraction of direct light reflected into the sensor-quad.

At low sun elevations (e.g., ${\theta _0} = 60^\circ $), the variability of ${\rho _D}$ due to different sky-radiance models is larger than that due to different sea-surface models—c.f., Fig. 10(c) and Fig. 10(f). This results from the different capabilities of the various methods to represent the atmospheric multiple scattering, which exhibits more pronounced effects with sun zenith (see Fig. 4).

The above considerations on the different impacts of sky-radiance distributions and modeled sea surfaces are confirmed by the analysis of Fig. 10(g-i) showing their combined effects. Variations of $\rho $ are similar to those induced by alternative sea-surface models with low ${\theta _0}$ (c.f., Fig. 10(g) and Fig. 10(d)), and by different sky-radiance models with high ${\theta _0}$ (c.f., Fig. 10(i) and Fig. 10(c)).

Table 2 primarily aims at reporting the values of $\rho $ and providing additional elements to better understand peculiarities embedded in Fig. 10. Note that any application relying on the $\rho $ values included in this table should consider that simulations have been performed at a specific wavelength (i.e., 550 nm). Data in the top, middle and bottom data-block refer to the impact of different sky-radiance models, sea-surface models, and their combined effects, respectively. Row- and column-wise data contain statistical figures for the considered wind-speed ${v_w}$ and sun-zenith ${\theta _0}$ values, respectively. For selected ${v_w}$ and ${\theta _0}$, data sub-sets report the mean $\mu $, the standard deviation $\sigma $ and the coefficient of variation CV in percent ($100 \cdot \sigma /\mu $) of $\rho $, in the first, second and third entry from top to bottom.

Tables Icon

Table 2. Statistical figures summarizing the variability of ρ at 550 nm. For each selected ${\theta _0}$ and ${v_w}$, data sub-sets report the mean $\mu $, the standard deviation $\sigma $ and the coefficient of variation CV in percent ($100 \cdot \sigma /\mu $) in the first, second and third entry from top to bottom.

Data indicate that CVs characterizing $\rho $ due to the application of different sky-radiance models are generally below 2% for ${v_w}$ up to 4 $\textrm{m}{\textrm{s}^{ - 1}}$. Conversely, CVs vary between 3% and 6% when ${v_w}$ reaches 10 $\textrm{m}{\textrm{s}^{ - 1}}$. At low wind speed, the sky-radiance contributions mainly originate from regions of the sky dome where the models mostly agree. An increased wind speed, instead, conveys sky-radiance contributions from a larger number of quads into the sensor field of view. These contributions may include those from the circumsolar and the horizon regions where the accuracy in the simulation of forward and multiple scattering, respectively, is more challenged (see Fig. 4).

When considering the diverse models applied to represent the sea surface, CVs of approximately $1.7$% characterize the values determined with ${v_w}$ = 0 $\textrm{m}{\textrm{s}^{ - 1}}$. This non-null value is explained by the fact that, in absence of wind, the CM2 and ECKV models assume a completely flat sea surface, while CM1 considers the sea surface to be always perturbed by some smooth wave, even in the absence of local wind (as in reality, see Fig. 8).

Figure 8 is also relevant to explain another detail that emerges from the analysis of Table 2. Still referring to the case of different sea-surface models, statistical figures indicate a general CV increase with ${v_w}$. However, an opposite tendency can be seen varying ${v_w}$ from 4 to 6 $\textrm{m}{\textrm{s}^{ - 1}}$. For instance, when ${\theta _0} = 10^\circ $, CV varies from 7.4% for ${v_w}$ = 4 $\textrm{m}{\textrm{s}^{ - 1}}$ to 6.5% for ${v_w}$ = 6 $\textrm{m}{\textrm{s}^{ - 1}}$. This can be explained with the help of Fig. 8. The variance of the sea-surface slope distribution foreseen by CM1 and CM2 linearly varies with wind speed (the former is higher, and a constant bias separates them). The variance predicted by the ECKV model, instead, is non-linear with wind speed. Specifically, the ECKV value is above that of CM1 with ${v_w}$ = 4 $\textrm{m}{\textrm{s}^{ - 1}}$, while it is in between those of CM1 and CM2 with ${v_w}$ = 6 $\textrm{m}{\textrm{s}^{ - 1}}$. The slope variance hence varies more in the former than in the latter case, which also affects the $\rho $ replicability.

Finally, considering the combined effects of the various sky-radiance and sea-surface models on the replicability of $\rho $, results further confirm that cases with higher wind speed, and sun elevation close to the zenith or to the horizon, are those where the sky-radiance and the sea-surface modeling is more influential.

3.2 Benchmark results

The values of $\rho $ determined with Hydrolight based on the HC sky-radiance and CM2 sea-surface models [12] are largely applied by the scientific community and currently recommended into the validation protocols [7] to derive ${L_w}$ from measurements of ${L_i}$ and ${L_t}$. A comparison between these values and equivalent ones obtained with MOX using the HC sky-radiance and the CM2 sea-surface models is presented in Fig. 11.

 figure: Fig. 11.

Fig. 11. Values of $\rho $ determined with (a) Hydrolight and (b) MOX adopting the same sky-radiance and sea-surface statistics, with solar zenith angle varying from $10^\circ $ to $80^\circ $. (c) Direct comparison between Hydrolight and MOX derived $\rho $. (d) Percent difference of $\rho $ values determined with MOX with respect to Hydrolight.

Download Full Size | PDF

Results show that the values of $\rho $ computed with MOX are always appreciably higher than those from Hydrolight. This difference also suggests an impact on the experimental determination of the water-leaving radiance through above-water radiometry, which relies on the application of $\rho $. The source of the observed difference is addressed in the next discussion section, benefitting of simplified simulation conditions and additional benchmark cases.

4. Discussion

4.1 Analysis of benchmark results and explanation of differences

In view of addressing differences between Hydrolight and MOX results, additional $\rho $ simulations were performed for an isotropic sky-radiance distribution, still modeling the sea-surface statistics with CM2 for measurement geometries defined by $\theta = 140^\circ $ and additionally $\theta = 180^\circ $. This simulation set permits to consider, as an additional and fully independent term of comparison, the $\rho $ values determined by Austin [10] for an isotropic sky and the Cox-Munk sea-surface statistics applying the ergodic-cap scheme proposed by Gordon [9]. The ergodic-cap approach is based on the construction of a three-dimensional surface with a slope distribution equivalent to that of the surface waves. The center and the boundary of this surface are associated with the lower and the larger waves slope, hence the term cap. Note that the cap is stretched to account for the slope probability distribution of waves, and a correction term is applied to resolve its projection in the viewing direction when computing $\rho $.

For notation compactness, the sea-surface reflectance factors are here indicated as $\rho _{{v_w}}^C(\theta )$, where: 1) the symbol C is H for Hydrolight, M for MOX, and A for Austin, 2) the values of the zenith viewing angle $\theta $ are $180^\circ $ and $140^\circ $, and 3) the wind-speed values ${v_w}$ are $4$ and 10 $\textrm{m}{\textrm{s}^{ - 1}}$. As an example, $\rho _4^H(180) = 0.0201$ denotes the sea-surface reflectance factor computed with Hydrolight assuming an isotropic radiance for ${v_w}$=4 $\textrm{m}{\textrm{s}^{ - 1}}$and nadir view.

Table 3 confirms that the values of $\rho $ computed with Hydrolight are lower than those obtained with MOX also in the presence of isotropic sky-radiance distribution, e.g., $\rho _4^H(140) = 0.0251$ and $\rho _4^M(140) = 0.0267$. The latter MOX values show a significant agreement with those from Austin. It is further highlighted that the Hydrolight results $\rho _4^H(180) = 0.0201$ and $\rho _{10}^H(180) = 0.0194$ are below the minimum expected values of $\rho = 0.0211$ for a flat Fresnel surface in nadir view. The latter finding, which does not appear to have a physical explanation, suggests investigating ray-tracing details.

Tables Icon

Table 3. Comparison of ρ values obtained from Hydrolight, MOX and by Austin for an isotropic sky-radiance distribution and the CM2 sea-surface statistics.

This is undertaken with the help of Fig. 12, where: 1) the sky-radiance is assumed isotropic, 2) the sea surface is one-dimensional and for simplicity chosen to coincide with the measurement plane (see Fig. 2), 3) multiple reflections are not accounted for, and finally 4) waves occluding effects are neglected (the last assumption is supported by the viewing direction of the virtual radiometer [9,11]).

 figure: Fig. 12.

Fig. 12. Schematic of ray tracing in the measurement plane.

Download Full Size | PDF

The discretized sea surface shown in Fig. 12 is composed by a set of segments with coordinates $({x_j},{z_j})$ with $j = 0, \ldots ,N$. These segments are determined by sampling angles ${\alpha _j}$ from the waves slope distribution, and by defining ${z_0} = 0$ and ${z_j} = {z_{j - 1}} + \Delta {z_j}$ with a fixed interval $\Delta x$ and elevation $\Delta {z_j} = \Delta x\tan {\alpha _j}$. The length of the j-th segment is ${A_j} = \Delta x\sec {\alpha _j}$. The projection of the j-th segment onto the plane perpendicular to the observation direction ${\mathbf v}$, which is the segment seen by the virtual radiometer, has length ${w_j}$

$${w_j} = {A_j}\cos {\beta _j} = \Delta x\frac{{\cos {\beta _j}}}{{\cos {\alpha _j}}}$$
where ${\beta _j}$ is the angle between the surface normal ${{\mathbf n}_j}$ and $- {\mathbf v}$.

With $W = \sum\limits_{j = 1}^N {{w_j}} $ and recalling the assumption of isotropic sky-radiance, the value of ${L_r}$ is computed as

$${L_r} = {L_i}\frac{1}{W}\sum\limits_{j = 1}^N {R_j^F} {w_j}$$
and hence the sea-surface reflectance factor is
$$\rho = \frac{1}{W}\sum\limits_{j = 1}^N {R_j^F} {w_j}$$
where $R_j^F$is the Fresnel reflectance corresponding to the angle ${\beta _j}$. Note that the normalized weight ${{{w_j}} / W}$ is the probability to observe the j-th sea-surface facet [11,55].

The ray initialization performed in MOX from a random point of the domain top permits to statistically consider the relative weight of each triad according to its projection in the ray direction (see Fig. 7). In fact, the “visible” area (i.e., projection) of each facet in the sensor direction varies according to the facet orientation. This feature is also considered by Austin with the ergodic-cap approach utilizing a projection correction, which supports the agreement between $\rho $ determined with these simulation schemes. Conversely, the Hydrolight scheme implies that about the same number of rays (i.e., excluding the minority that is affected by the shadowing effect) hits the central triad for repeated hexagon realizations, as illustrated in Fig. 6. When computing the mean of the radiant flux ratio (see Sec. 2.5.1), the former assumption hampers the capability to account for the relative weight of the reflecting triad at the center of each hexagon realization due to its projection towards the sensor-quad. In other words, directing the rays towards the central triad does not allow to properly account that the “visible” area of the reflecting facet varies among sea surface realizations, with clear implications on the computation of the average contribution.

Additional MC tests have confirmed the above findings by showing that: 1) backward and forward Monte Carlo simulations converge to the values of $\rho $ that are close to those computed with MOX, and 2) the forward scheme performed with the simple arithmetic mean instead of a weighted mean as in Eq. (22), leads to smaller values of $\rho $. In this case, the underestimate is analogous to that observed for Hydrolight—e.g., $\rho = 0.0250$ vs. $\rho _H^4(140) = 0.0251$ and $\rho = 0.0192$ vs. $\rho _H^{10}(180) = 0.0194$. Differences in the ray-tracing schemes are hence identified as the key factor explaining the lower values of $\rho $ computed by Hydrolight with respect to MOX.

4.2 Relevance of the statistical representativity of in situ measurements

The statistical representativity of ${L_t}$ measurements is a fundamental aspect in above-water radiometry. Remarking that ${L_t} = {L_r} + \textrm{ }{L_w}$, and recalling that ${L_r}$ applied for the determination of $\rho $ is quantified numerically, the actual ${L_r}$ characterizing the in situ measurements may exhibit differences from the theoretical value of ${L_r}$. The working hypothesis that underlies the theoretical determination of $\rho $, which should be also featured by in situ measurements, is that the reflective properties of the sea surface are averaged over time, space and propagation direction of waves. Individual ${L_t}$ measurements may instead vary over time and as a function of the sensor features (e.g., field-of-view and sampling time). Consequently, each measurement can only capture part of the variability induced by a wavy surface over time and space.

This aspect is illustrated through Fig. 13. Dots in the right side of the sky hemisphere exhibiting HC sky-radiance distribution with ${\theta _0} = 20^\circ $, represent ten independent rays trajectories randomly sampled in backward mode from a virtual sensor without any restriction applied to sky- and sun-glint contributions. Their dispersion is explained by the slope distribution of waves here generated with CM2 for ${v_w}$ = 4 $\textrm{m}{\textrm{s}^{ - 1}}$. Assuming that each trajectory corresponds to a different measurement performed by the sensor with infinitesimal integration time and field-of-view, the mean ${L_r}$ values of the two simulated data records are 0.0020 and 0.0022 $\textrm{W}\; {\textrm{m}^{ - 2}}\; \textrm{n}{\textrm{m}^{ - 1}}\; \textrm{s}{\textrm{r}^{ - 1}}$, thus indicating a $10$% difference. The former result firmly suggests the need to rely on the averaging of a statistically significant number of ${L_t}$ measurements to enhance consistency between simulated and actual ${L_r}$.

 figure: Fig. 13.

Fig. 13. Example of repeated determinations of ${L_r}$ through a limited set of virtual measurements. The sky radiance and the sea surface are represented by the HC (${\theta _0} = 20^\circ$) and the CM2 (${v_w} = 4\textrm{m}{\textrm{s}^{ - 1}}$) models, respectively. The ${L_r}$ values corresponding to panel (a) and (b) are 0.0020 and 0.0022 $\textrm{W }{\textrm{m}^{ - 2}}\textrm{n}{\textrm{m}^{ - 1}}\textrm{s}{\textrm{r}^{ - 1}}$, respectively.

Download Full Size | PDF

Figure 13 illustrates an idealized case that only accounts for the variability of ${L_r}$ due to the slope variance of surface waves. Differences in repeated ${L_t}$ measurements, however, also depend on the sky-radiance distribution (that can vary as a function of several parameters including wavelength). Still, acknowledging the impossibility to account for any specific parameter that may impact the spatio-temporal averaging of surface effects in ${L_t}$ such as the sensor sampling time and field-of-view or its height above the sea level, the former considerations indicate the need for $\rho $ factors relying on those sky-radiance (also accounting for spectral dependence) and sea-surface models exhibiting the highest accuracy.

4.3 Further considerations on operational above-water measurements

In addition to the statistical representativity of ${L_t}$ measurements addressed in the former section, uncertainties and biases affecting $\rho $ unavoidably impact the determination of ${L_w}$. Such an impact can be quantified in relative terms following Gergely and Zibordi [56], but excluding the effects of the uncertainties affecting ${L_t}$ and ${L_i}$, as

$$ \frac{{u({L_w})}}{{{L_w}}} = \frac{{{L_i}}}{{{L_w}}}u(\rho ),$$
where $u({L_w})$ and $u(\rho )$ indicate the absolute uncertainties of ${L_w}$ and $\rho$, respectively.

Specifically, by assuming 1) ${v_w}$ = 4 $\textrm{m}{\textrm{s}^{ - 1}}$, 2) ${\theta _0}$ larger than $20^\circ $, and 3) setting $u(\rho )$ equal to the standard deviation of $\rho $ given in Table 2, the impact on ${L_w}$ (i.e., $100\textrm{ } \cdot \textrm{ }u({L_w})/{L_w}$) at 550 nm due to the adoption of diverse sky-radiance and sea-surface models in the determination of ρ, is lower than 0.5% for most water types and clear sky conditions.

More importantly, in agreement with Eq. (1), any underestimate (overestimate) of $\rho $ leads to an overestimate (underestimate) of ${L_w}$. Regarding the data shown in Fig. 11, a negative bias of approximately 0.002 affecting $\rho $ at 550 nm (comparable to that shown by the values of $\rho $ determined with MOX and Hydrolight for ${v_w} = \textrm{ }4\,\textrm{m}{\textrm{s}^{ - 1}}$ and ${\theta _0}$ larger than $20^\circ $), induces an overestimate in ${L_w}$ approaching 1.5%, (see Eq.s (1) and (3)).

The above considerations have relevance for the current processing solutions applied to AERONET-OC data [57,20], which rely on a modified version of Eq. (1) and the application of $\rho $ determined using Hydrolight [12]. Anticipating that the AERONET-OC data largely refer to sea-state conditions driven by low-to-mild wind speeds generally not exceeding ${v_w} = 5$ $\textrm{m}{\textrm{s}^{ - 1}}$ as a result of strict quality assurance and quality control schemes, the value of ${L_t}$ applied for the determination of ${L_w}$ in the AERONET-OC processing is the mean of a percentage of the individual sea-radiance data exhibiting the lowest values in a measurement sequence. This solution, only supported by experimental evidence provided by the comparison of ${L_w}$ data determined from above-water and in-water methods [7,58], was tentatively justified by the need to exclude those sea-viewing measurements exhibiting excessively large glint contamination and expected to lead to an overestimate of ${L_t}$. The $\rho $ values displayed in Fig. 11 give novel elements to explain the better performance of Eq. (1) when relying on the averaging of relative minima from a series of sea-viewing measurements opposite to the averaging of all the measurements. Given the underestimate of the $\rho $ values applied in data processing, the related impact on the determination of ${L_w}$ is heuristically compensated by an underestimate of ${L_t}$ as resulting from the averaging of a fraction of the sea-viewing data exhibiting the lowest values in each measurement sequence.

5. Summary and remarks

Results from the study, restricted to a consolidated measurement geometry [7], indicate that the variability of $\rho $ derived from the application of alternative methods to determine the sky-radiance distribution and the sea-surface statistics is generally below 2% when considering wind speeds up to 4 ms−1 and sun zenith angles larger than approximately $20^\circ $. The variability of $\rho $ increases with wind speed, especially when the sun is close to the zenith or the horizon. This dependence has been interpreted acknowledging that, when the variance of the sea-surface slope distribution increases due to wind speed, the sky region that contributes to the radiance reflected towards the sensor-quad widens to include areas where sky-radiance distributions can show significant differences (e.g., the circumsolar and horizon regions) both in intensity and spectrally.

The variability of $\rho $ resulting from the replicability exercise can exceed 5% for wind speeds higher than 4 $\textrm{m}{\textrm{s}^{ - 1}}$ and sun elevations close to the zenith (${\theta _0} = 20^\circ $ or less). This result indicates that at low sun zenith angles, the determination of ${L_w}$ through above-water radiometric methods is challenged by the capability to properly represent the relative contribution of the direct and diffuse sky-radiance in the computation of $\rho $.

The ρ replicability has been evaluated at 550 nm to specifically account for reference values as benchmark [12], whilst analyses at different wavelengths are beyond the scope of this work. The study has hence shown that the values of $\rho $ computed with Hydrolight are lower than those obtained with MOX regardless of the adoption of the same sky-radiance and sea-surface models. The reason for this difference has been identified in the ray-tracing schemes employed in the two radiative transfer codes. The solution applied in Hydrolight to improve simulation efficiency by fixing the point towards which rays are emitted, presents some caveats when combining results from different sea-surface realizations. Specifically, the relative weight of the reflecting sea-surface facets due to their projections towards the sensor-quad is not taken into consideration in the determination of $\rho $. Conversely, the MOX code implicitly accounts for this effect in statistical terms as a result of the randomness of the starting point from the top of the simulation domain for rays traced in backward mode. This allows improving the accuracy of the computed $\rho $ although additional processing resources are demanded.

The study also offered discussion elements on the operational determination of the water-leaving radiance with above-water systems. One aspect is that the uncertainties affecting ${L_w}$ naturally depend on the actual statistical representativity of the replicated sea-viewing measurements applied for the determination of ${L_t}$, besides the accuracy of $\rho $.

Results from the study also support the rationale for the AERONET-OC processing applying ${L_t}$ values determined from the averaging of relative minima, in alternative to the averaging, of a series of sea-viewing measurements. This solution is shown contributing to compensate for the impact of the underestimated $\rho $ values from Hydrolight simulations.

Funding

Joint Research Center of European Commission.

Acknowledgments

The authors would like to acknowledge constructive comments from Curt Mobley on an early draft of the paper. Computer simulations presented in this paper were partly executed on the Navigator supercomputer (University of Coimbra, Portugal).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. B. A. Franz, S. W. Bailey, P. J. Werdell, and C. R. McClain, “Sensor-independent approach to the vicarious calibration of satellite ocean color radiometry,” Appl. Opt. 46(22), 5068–5082 (2007). [CrossRef]  

2. G. Zibordi, C. J. Donlon, and A. C. Parr, Optical Radiometry for Ocean Climate Measurements, Experimental Methods in the Physical Sciences (Elsevier Science, 2014), Vol. 47.

3. P. J. Werdell and C. R. McClain, “Satellite Remote Sensing: Ocean Color,” in Encyclopedia of Ocean Sciences (Third Edition), J. K. Cochran, H. J. Bokuniewicz, and P. L. Yager, eds., 3rd Edition (Academic Press, 2019), pp. 443–455. [CrossRef]  

4. G. Zibordi, D. D’Alimonte, and J.-F. Berthon, “An evaluation of depth resolution requirements for optical profiling in coastal waters,” J. Atmos. Ocean. Tech. 21(7), 1059–1073 (2004). [CrossRef]  

5. D. D’Alimonte, E. B. Shybanov, G. Zibordi, and T. Kajiyama, “Regression of in-water radiometric profile data,” Opt. Express 21(23), 27707–27733 (2013). [CrossRef]  

6. D. D’Alimonte, G. Zibordi, and T. Kajiyama, “Effects of integration time on in-water radiometric profiles,” Opt. Express 26(5), 5908–5939 (2018). [CrossRef]  

7. G. Zibordi, K. J. Voss, B. C. Johnson, and J. L. Mueller, IOCCG Protocol Series (2019). Protocols for Satellite Ocean Colour Data Validation: In Situ Optical Radiometry (IOCCG, 2019).

8. K. G. Ruddick, K. Voss, E. Boss, A. Castagna, R. Frouin, A. Gilerson, M. Hieronymi, B. C. Johnson, J. Kuusk, Z. Lee, M. Ondrusek, V. Vabson, and R. Vendt, “A review of protocols for Fiducial Reference Measurements of water-leaving radiance for validation of satellite remote-sensing data over water,” Remote Sensing 11, (2019).

9. J. I. Gordon, “Directional radiance (luminance) of the sea surface,” in (Naval Ship Systems Command Department of the Navy Washington, D. C. 20360, 1969).

10. R. W. Austin, “Inherent Spectral Radiance Signatures of the Ocean Surface,” in Ocean Color Analysis (Scripps Institution of Oceanography, 1974).

11. V. Kisselev and B. Bulgarelli, “Reflection of light from a rough water surface in numerical methods for solving the radiative transfer equation,” J. Quant. Spectrosc. Radiat. Transfer 85(3-4), 419–435 (2004). [CrossRef]  

12. C. D. Mobley, “Estimation of the remote-sensing reflectance from above-surface measurements,” Appl. Opt. 38(36), 7442–7455 (1999). [CrossRef]  

13. Z. Lee, Y.-H. Ahn, C. Mobley, and R. Arnone, “Removal of surface-reflected light for the measurement of remote-sensing reflectance from an above-surface platform,” Opt. Express 18(25), 26313–26324 (2010). [CrossRef]  

14. T. Harmel, A. Gilerson, A. Tonizzo, J. Chowdhary, A. Weidemann, R. Arnone, and S. Ahmed, “Polarization impacts on the water-leaving radiance retrieval from above-water radiometric measurements,” Appl. Opt. 51(35), 8324–8340 (2012). [CrossRef]  

15. C. D. Mobley, “Polarized reflectance and transmittance properties of windblown sea surfaces,” Appl. Opt. 54(15), 4828–4849 (2015). [CrossRef]  

16. M. Hieronymi, “Polarized reflectance and transmittance distribution functions of the ocean surface,” Opt. Express 24(14), A1045–A1068 (2016). [CrossRef]  

17. D. D’Alimonte and T. Kajiyama, “Effects of light polarization and waves slope statistics on the sea-surface reflectance,” Opt. Express 24(8), 7922–7942 (2016). [CrossRef]  

18. D. D’Alimonte, G. Zibordi, T. Kajiyama, and J. C. Cunha, “Monte Carlo code for high spatial resolution ocean color simulations,” Appl. Opt. 49(26), 4936–4950 (2010). [CrossRef]  

19. T. Kajiyama, D. D’Alimonte, and J. C. Cunha, “A high-performance computing framework for Monte Carlo ocean color simulations,” Concurrency Computat.: Pract. Exper. 29(4), e3860 (2017). [CrossRef]  

20. G. Zibordi, B. N. Holben, M. Talone, D. D’Alimonte, I. Slutsker, D. M. Giles, and M. G. Sorokin, “Advances in the Ocean Color Component of the Aerosol Robotic Network (AERONET-OC),” J. Atmos. Ocean. Tech. 38(4), 725–746 (2021). [CrossRef]  

21. C. D. Mobley, Light and Water. Radiative Transfer in Natural Waters (Academic Press, 1994).

22. X. Zhang, S. He, A. Shabani, P.-W. Zhai, and K. Du, “Spectral sea surface reflectance of skylight,” Opt. Express 25(4), A1 (2017). [CrossRef]  

23. P. M. M. Groetsch, R. Foster, and A. Gilerson, “Exploring the limits for sky and sun glint correction of hyperspectral above-surface reflectance observations,” Appl. Opt. 59(9), 2942 (2020). [CrossRef]  

24. J. Pitarch, M. Talone, G. Zibordi, and P. Groetsch, “Determination of the remote-sensing reflectance from above-water measurements with the “3C model”: a further assessment,” Opt. Express 28(11), 15885 (2020). [CrossRef]  

25. A. W. Harrison and C. A. Coombes, “An opaque cloud cover model of sky short wavelength radiance,” Sol. Energy 41(4), 387–392 (1988). [CrossRef]  

26. A. W. Harrison and C. A. Coombes, “Angular distribution of clear sky short wavelength radiance,” Sol. Energy 40(1), 57–63 (1988). [CrossRef]  

27. A. W. Harrison and C. A. Coombes, “Performance validation of the Gueymard sky radiance model,” Atmos.-Ocean 27(3), 565–576 (1989). [CrossRef]  

28. G. Zibordi and K. J. Voss, “Geometrical and spectral distribution of sky radiance—comparison between simulations and field measurements,” Remote Sens. Environ. 27(3), 343–358 (1989). [CrossRef]  

29. B. Bulgarelli, V. B. Kisselev, and L. Roberti, “Radiative transfer in the atmosphere-ocean system: the finite-element method,” Applied Optics 38, (1999).

30. B. Bulgarelli, G. Zibordi, and J.-F. Berthon, “Measured and modeled radiometric quantities in coastal waters: toward a closure,” Appl. Opt. 42(27), 5365–5381 (2003). [CrossRef]  

31. B. Bulgarelli and J. P. Doyle, “Comparison between numerical models for radiative transfer simulation in the atmosphere-ocean system,” J. Quant. Spectrosc. Radiat. Transfer 86(3), 315–334 (2004). [CrossRef]  

32. B. Bulgarelli, V. Kiselev, and G. Zibordi, “Simulation and analysis of adjacency effects in coastal waters: a case study,” Appl. Opt. 53(8), 1523–1545 (2014). [CrossRef]  

33. W. W. Gregg and K. L. Carder, “A simple spectral solar irradiance model for cloudless maritime atmospheres,” Linmol. Oceanogr. 35(8), 1657–1675 (1990). [CrossRef]  

34. G. Thuillier, M. Hersé, D. Labs, T. Foujols, W. Peetermans, D. Gillotay, P. C. Simon, and H. Mandel, “The solar spectral irradiance from 200 to 2400 nm as measured by the Solspec spectrometer from the Atlas and Eureca missions,” Sol. Phys. 214(1), 1–22 (2003). [CrossRef]  

35. C. Fröhlich and G. E. Shaw, “New determination of Rayleigh scattering in the terrestrial atmosphere,” Appl. Opt. 19(11), 1773–1775 (1980). [CrossRef]  

36. A. T. Young, “Revised depolarization corrections for atmospheric extinction,” Appl. Opt. 19(20), 3427–3428 (1980). [CrossRef]  

37. A. Ångström, “Techniques of determinig the turbidity of the atmosphere 1,” Tellus 13(2), 214–223 (1961). [CrossRef]  

38. E. Vigroux, “Contribution à l’étude expérimentale de l’absorption de l’ozone,” Ann. Phys. 12(8), 709–762 (1953). [CrossRef]  

39. G. I. Pokrowski, “Uber eisen scheinbaren Mie-effect und seine mogliche rolle in der atmosphairenoptik,” Z. Phys 53(1-2), 67–71 (1929). [CrossRef]  

40. R. Kittler, “Standardization of outdoor conditions for the calculation of daylight factor with clear skies,” in Proc. C.I.E (Bouwcentrum International, 1965), pp. 273–285.

41. V. V. Sobolev, Light Scattering in Planetary Atmospheres, International Series of Monographs in Natural Philosophy (Pergamon, 1975), Vol. 76.

42. H. R. Gordon, D. K. Clark, J. W. Brown, O. B. Brown, R. H. Evans, and W. W. Broenkow, “Phytoplankton pigment concentrations in the Middle Atlantic Bight: comparison of ship determinations and CZCS estimates,” Appl. Opt. 22(1), 20–36 (1983). [CrossRef]  

43. Z. Jin, T. P. Charlock, W. L. Smith, and K. Rutledge, “A parameterization of ocean surface albedo,” Geophys. Res. Lett. 31, (2004).

44. L. C. Henyey and J. L. Greenstein, “Diffuse radiation in the Galaxy,” Astrophys. J. 93, 70 (1941). [CrossRef]  

45. R. Guzzi, R. Rizzi, and G. Zibordi, “Atmospheric correction of data measured by a flying platform over the sea: elements of a model and its experimental validation,” Appl. Opt. 26(15), 3043 (1987). [CrossRef]  

46. W. A. Marggraf and M. Griggs, “Aircraft measurements and calculations of the total downward flux of solar radiation as a function of altitude,” J. Atmos. Sci. 26(3), 469–477 (1969). [CrossRef]  

47. L. Elterman, UV, Visible, and IR Attenuation for Altitudes to 50 Km (Air Force Cambridge Research Laboratories, 1968).

48. C. Cox and W. Munk, “Measurement of the roughness of the sea surface from photographs of the sun’s glitter,” J. Opt. Soc. Am. 44(11), 838–850 (1954). [CrossRef]  

49. T. Elfouhaily, B. Chapron, K. Katsaros, and D. Vandemark, “A unified directional spectrum for long and short wind-driven waves,” J. Geophys. Res. 102(C7), 15781–15796 (1997). [CrossRef]  

50. W. J. Pierson and L. Moskowitz, “A proposed spectral form for fully developed wind seas based on the similarity theory of S. A. Kitaigorodskii,” J. Geophys. Res. 69(24), 5181–5190 (1964). [CrossRef]  

51. F. M. Bréon and N. Henriot, “Spaceborne observations of ocean glint reflectance and modeling of wave slope distributions,” J. Geophys. Res. 111(C6), C06005 (2006). [CrossRef]  

52. R. W. Preisendorfer and C. D. Mobley, “Albedos and glitter patterns of a wind-roughened sea surface,” J. Phys. Oceanogr. 16(7), 1293–1316 (1986). [CrossRef]  

53. R. W. Preisendorfer and C. D. Mobley, Unpolarized Irradiance Reflectances and Glitter Patterns of Random Capillary Waves on Lakes and Seas, by Monte Carlo Simulation. NOAA Technical Memorandum ERL PMEL-63 (National Oceanic and Atmospheric Administration, Pacific Marine Environmental Laboratory, Seattle, Washington, 1985).

54. K. M. Case, “Transfer problems and the reciprocity principle,” Rev. Mod. Phys. 29(4), 651–663 (1957). [CrossRef]  

55. H. R. Gordon, Physical Principles of Ocean Color Remote Sensing (University of Miami, 2019). [CrossRef]  

56. M. Gergely and G. Zibordi, “Assessment of AERONET-OC Lwn uncertainties,” Metrologia 51(1), 40–47 (2014). [CrossRef]  

57. G. Zibordi, B. Holben, I. Slutsker, D. Giles, D. D’Alimonte, F. Mélin, J.-F. Berthon, D. Vandemark, H. Feng, G. Schuster, B. E. Fabbri, S. Kaitala, and J. Seppälä, “AERONET-OC: A network for the validation of ccean color primary radiometric products,” J. Atmos. Oceanic Tech. 26(8), 1634–1651 (2009). [CrossRef]  

58. G. Zibordi, “Experimental evaluation of theoretical sea surface reflectance factors relevant to above-water radiometry,” Opt. Express 24(6), A446–A459 (2016). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (13)

Fig. 1.
Fig. 1. (a) Perspective-, (b) side- and (c) top-view of typical above-water measurement geometry. Note that the sea- and sky-viewing angles are defined with respect to the zenith.
Fig. 2.
Fig. 2. Partitioning of the upper hemisphere into quads. The thick solid and dashed lines indicate the measurement and the solar plane, respectively. The sensor-quad marked in black represents the virtual radiometer.
Fig. 3.
Fig. 3. Sky-radiance distribution Lsky at 550 nm (restricted to the diffuse component) obtained with HC, ZV and FEM displayed from the top to the bottom row, respectively. Panels from left to right correspond to ${\theta _0}$  = 20°, 40° and 60°.
Fig. 4.
Fig. 4. Sky radiance Lsky at 550 nm (restricted to the diffuse component) determined for the solar and measurement planes (solid and dashed lines, respectively) with HC, ZV and FEM for different ${\theta _0}$ (i.e., $20^\circ ,\,40^\circ $ and $60^\circ $ ). Dots highlight the zenith angles at which the sky-radiance values are computed (see Sec. 2.1). The direct radiance components from the sun-quad at $20^\circ ,\,40^\circ $ and $60^\circ $ (not displayed) are respectively 77.03, 37.32 and 21.16 $\textrm{W }{\textrm{m}^{ - 2}}\textrm{n}{\textrm{m}^{ - 1}}\textrm{s}{\textrm{r}^{ - 1}}$ for both HC and ZV. Corresponding FEM values are 77.13, 37.36 and 21.15 $\textrm{W }{\textrm{m}^{ - 2}}\textrm{n}{\textrm{m}^{ - 1}}\textrm{s}{\textrm{r}^{ - 1}}$ .
Fig. 5.
Fig. 5. CM2 (top panels) and ECKV (bottom panels) representations of the sea surface. The left and right columns refer to ${v_\textrm{w}}$ of 3 and 7 $\textrm{m}{\textrm{s}^{ - 1}}$ , respectively.
Fig. 6.
Fig. 6. Illustration of the Hydrolight ray-tracing scheme. (a) Rays initialization from a random position of each emitting quad. (b) Perspective view, where dashed and solid lines illustrate incident rays and their reflection towards the sensor-quad, respectively. Rays reflected out of the sensor-quad are not shown. Each ray corresponds to a different hexagon realization (only one surface is however displayed).
Fig. 7.
Fig. 7. As in Fig. 6, but for the MOX backward ray-tracing scheme. The solid and dashed lines denote rays emitted from the sensor-quad and reflected by the sea surface, respectively. Rays that exit the domain from one side continue their path at the opposite side based on the periodicity of the sea surface.
Fig. 8.
Fig. 8. Slope variance as a function of the wind speed for each sea-surface model addressed in the study (i.e., CM1, CM2 and ECKV labeled with a circle, square and triangle, respectively). The solid lines connecting the symbols represent the expected theoretical values. The mean and the standard deviation of the slope variance resulting from multiple sea-surface realizations are displayed as dashed lines and shaded regions, respectively. Overlapping lines confirm the satisfactory agreement between expected and computed trends.
Fig. 9.
Fig. 9. Examples of radiance reflectance matrix based on the CM1, CM2 and the ECKV sea-surface statistics in the top, middle and bottom row panels, respectively. Column panels from left to right correspond to vw = 2, 6 and 10 ms-1, respectively. The solar plane crossing the hemisphere from $\phi = 0^\circ $ to $\phi = 180^\circ $ and the radiometer orientation from $\phi = 270^\circ $ to $\phi = 90^\circ $ are indicated by the dashed line and the arrow, respectively.
Fig. 10.
Fig. 10. Variability of $\rho $ values as a function of wind speed, as resulting from the application of various models for sky-radiance distribution at 550 nm (first row of panels) and sea-surface statistics (middle row of panels), and the combined effects of sky-radiance and sea surface (bottom row of panels). Column panels from left to right report results for ${\theta _0}$  = 20°, 40° and 60°, respectively. The mean and the standard deviation of $\rho $ values are represented with a trend line and a shaded region, respectively.
Fig. 11.
Fig. 11. Values of $\rho $ determined with (a) Hydrolight and (b) MOX adopting the same sky-radiance and sea-surface statistics, with solar zenith angle varying from $10^\circ $ to $80^\circ $ . (c) Direct comparison between Hydrolight and MOX derived $\rho $ . (d) Percent difference of $\rho $ values determined with MOX with respect to Hydrolight.
Fig. 12.
Fig. 12. Schematic of ray tracing in the measurement plane.
Fig. 13.
Fig. 13. Example of repeated determinations of ${L_r}$ through a limited set of virtual measurements. The sky radiance and the sea surface are represented by the HC ( ${\theta _0} = 20^\circ$ ) and the CM2 ( ${v_w} = 4\textrm{m}{\textrm{s}^{ - 1}}$ ) models, respectively. The ${L_r}$ values corresponding to panel (a) and (b) are 0.0020 and 0.0022 $\textrm{W }{\textrm{m}^{ - 2}}\textrm{n}{\textrm{m}^{ - 1}}\textrm{s}{\textrm{r}^{ - 1}}$ , respectively.

Tables (3)

Tables Icon

Table 1. Main quantities applied for the simulation of sky-radiance distribution.

Tables Icon

Table 2. Statistical figures summarizing the variability of ρ at 550 nm. For each selected θ 0 and v w , data sub-sets report the mean μ , the standard deviation σ and the coefficient of variation CV in percent ( 100 σ / μ ) in the first, second and third entry from top to bottom.

Tables Icon

Table 3. Comparison of ρ values obtained from Hydrolight, MOX and by Austin for an isotropic sky-radiance distribution and the CM2 sea-surface statistics.

Equations (23)

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

L w ( θ , ϕ , λ ) = L t ( θ , ϕ , λ ) L r ( θ , ϕ , λ )
L r ( θ , ϕ ) = 0 2 π 0 π 2 r ( θ , ϕ θ , ϕ ) L s k y ( θ , ϕ ) sin θ d θ d ϕ .
L r ( θ , ϕ ) = ρ L i ( θ , ϕ )
ρ = 0 2 π 0 π 2 r ( θ , ϕ θ , ϕ ) L s k y ( θ , ϕ ) sin θ d θ d ϕ L i ( θ , ϕ )
L d i f f H C ( θ , ϕ ) = ( 1.63 + 53.7 e 5.49 Ψ + 2.04 cos 2 ( Ψ ) cos ( θ 0 ) ) × ( 1 e 0.19 sec ( θ ) ) ( 1 e 0.53 sec ( θ 0 ) ) ,
cos ( Ψ ) = cos ( θ ) cos ( θ 0 ) + sin ( θ ) sin ( θ 0 ) cos ( ϕ )
L d i f f Z V ( θ , ϕ , λ ) = E 0 ( λ ) d 2 π σ ( τ d ( λ ) ) cos ( θ 0 ) exp [ τ a ( λ ) cos ( θ 0 ) ]
σ ( τ d ( λ ) ) = ( 1 γ ( λ ) ) R ( λ , θ ) + 2 γ ( λ ) R ( λ , θ 0 ) 4 + ( 3 P 1 ) ( 1 γ ( λ ) ) τ d ( λ ) 1 2 ( exp { τ d ( λ ) cos ( θ ) } + exp { τ d ( λ ) cos ( θ 0 ) } ) + [ β ( Ψ ) ( 3 + P 1 ) cos ( θ ) cos ( θ 0 ) ] σ 1 ( τ d ( λ ) )
L ( τ ; η , ϕ ) = m = 0 M L m ( τ , η ) cos m ϕ
η d L m ( τ , η ) d τ = L m ( τ , η ) + ω 0 l 2 1 1 L m ( τ , η ) p l m ( τ , η η ) d η + Q l m ( τ , η )
L m ( τ , η ) = j = 1 2 N L j m ( τ ) b j m ( η )
b j m ( η ) = Y j m ( η ) { 1 | η η j | h if  | η η j | < h 0 elsewhere
σ up 2 = 0.00316 v w
σ cr 2 = 0.003 + 0.00192 v w
σ slp 2 = 0.003 + 0.00512 v w
σ slp 2 = 0.00508 v w
z ( i , j ) = 1 N x N y u = 1 N x v = 1 N y z ^ ( u , v ) e i ( r u / N x + s v / N y )
L r = i N θ j N ϕ R ( i , j ) L s k y ( i , j ) .
ρ = L r L i ,
w j = A j cos β j = Δ x cos β j cos α j
L r = L i 1 W j = 1 N R j F w j
ρ = 1 W j = 1 N R j F w j
u ( L w ) L w = L i L w u ( ρ ) ,
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.