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

Scattering and absorption effects on asymptotic light fields in seawater

Open Access Open Access

Abstract

Asymptotic theory is based on the principle that the shape of the light field with depth gradually transforms from being dependent on the incident surface light field to being constant azimuthally and dependent only on the absorption and scattering properties of the water column. Properties such as the average cosine of the oceanic light field in the asymptotic regime, μ¯, are thus strictly inherent optical properties (IOPs). Because of the close link between asymptotic light fields and IOPs, radiative transfer approximations (RTAs) for the asymptotic regime have been adapted for use in algorithms describing surface remote sensing reflectance RRS ( = Lu/Ed) in terms of the IOPs. For such algorithms to have utility, the asymptotic average cosine needs to be parameterized in terms of IOPs useful for ocean color remote sensing. With this motivation, μ¯is approximated as a function of the ratio of total backscattering to total absorption, bb/a. An additional variable in assessments is the fractional water content of pure seawater in total backscattering, ηbb. A full range of representative phase functions for natural particle fields is included in the analysis using the Fournier-Forand analytical approximation. Analytical expressions for multi-order polynomial fits are provided for μ¯ as a function of bb/a for each ηbb assessed, for ηbb ≤ 0.49, and for the entire data set. The full range of phase function shapes were included in each fit. Percent absolute errors were a modest 3.4% for the fit for the entire data set. Additionally, a key assumption by Zaneveld that the attenuation coefficient for upwelling nadir radiance KLu in surface waters should be approximately equivalent to the attenuation coefficient in the asymptotic regime K was evaluated. Results provide justification and relationships for the targeted application of asymptotic parameters in ocean color RTAs for the surface ocean.

© 2017 Optical Society of America

1. Introduction

The average cosine μ¯ of the underwater light field is a single parameter describing the angular distribution of light at a particular depth, defined as:

μ¯(z)=02π0πL(θ,φ,z)cosθsinθdθdφ02π0πL(θ,φ,z)sinθdθdφ=EdEuEod+Eou  ,
where θ and ϕ are zenith and azimuth directions, z is depth (m), L is spectral radiance (W sr−1 m−2 nm−1), and E is spectral irradiance (W m−2 nm−1) with subscripts d, u, od, and ou representing plane downwelling, plane upwelling, scalar downwelling, and scalar upwelling, respectively (for optical property definitions, see Mobley [1]). If a light field is perfectly isotropic, μ¯= 0; if it is aligned entirely in the vertical downward direction, μ¯= 1.

The average cosine of the oceanic light field in the asymptotic regime, μ¯, is an inherent optical property (IOP), i.e., it is not dependent on the shape of the incident light field but on the absorption and scattering properties of the water column [2–6]. Asymptotic theory is based on the principle that the shape of the light field with depth gradually transforms from being dependent on the incident surface light field to being constant and azimuthally symmetric (so L is only a function of θ). In the asymptotic regime, the diffuse attenuation coefficient for irradiance reaches an asymptotic value K, where μ¯=aK, and a is the absorption coefficient. This is the asymptotic representation of Gershun’s equation, obtained by integrating the radiative transfer equation in all directions. Attenuation coefficients for all aspects of the light field, i.e., for all radiances and therefore all irradiances, are equivalent in the asymptotic regime and are also IOPs [7,8].

Absorption and scattering control μ¯ in a complex manner. Absorption elongates the radiance distribution with depth about the vertical direction because absorption increases with longer pathlength; the vertical direction is the shortest path and thus experiences the least effect from absorption. Scattering conversely disperses light in all directions, broadening the radiance distribution and decreasingμ¯. If scattering were isotropic, i.e., equal in all directions, μ¯would approach 0 with no absorption. However, the angular distribution of scattering in natural waters is strongly forward-peaked, which reduces such broadening.

Light fields in the asymptotic regime have received considerable attention, particularly in ocean optics literature from 1970’s to 1990’s (see reviews by [6] and [9]), because of the direct link μ¯ and K provide between AOPs and IOPs. In a series of papers, Zaneveld [2,3,10], developed radiative transfer approximations (RTAs) and an inversion to derive IOPs from radiometric measurements in the asymptotic regime that was adapted in [2] to an RTA for remote sensing reflectance RRS ( = Lu/Ed) in surface waters in terms of the IOPs. In this latter case, surface water IOPs define a conceptual μ¯ extrapolated to a hypothetically infinite, homogeneous water column. Even though the light field in surface waters is far from asymptotic, a critical parameter for remote sensing, the attenuation coefficient for upwelling nadir radiance KLu in surface waters, was hypothesized by Zaneveld [2] to be approximately equivalent to K because of the decoupling of upwelling radiances to downwelling radiance distributions. Note this would not apply in actual light fields in deep waters where influence from Raman inelastic scattering becomes apparent in the green and red [11].

Previous attempts to represent μ¯ in terms of IOPs have focused on the single scattering albedo ω ( = b/c), where b is total scattering and c is attenuation [2,3,5,6,12–14]. While convenient in that the ω range is bounded between 0 and 1, a drawback of this approach in the context of RTAs for ocean color remote sensing applications is b and c are not parameters that are readily derived from RRS (e.g., [15]). For algorithm testing and validation, these parameters are also difficult to directly measure without large errors because typical collimated collectors for c measurements have significant acceptance of scattered light at near-forward angles [16,17]. Total scattering b is normally derived from c – a, with the current convention of measuring in situ a and c with WET Labs ac devices [18]. The c measurement error is typically >25% for a device with ~1° acceptance [16,17] and there is no easily applied correction without high quality angular scattering measurements in the near-forward direction.

Gordon et al. [19] have shown, however, that RRS can be closely related to bb / (a + bb), where bb is the backscattering coefficient representing integrated scattering in the backward hemisphere. Furthermore, in cases where a is much greater than bb, the relationship can be simplified to bb/a. It can be easily shown that b/c = bb / (aB + bb) where B is the backscattering ratio, or bb/b, describing the general shape of the volume scattering function (VSF) [20,21]. Therefore bb / (a + bb) and bb/a may have good potential to also describeμ¯, with the advantage that these relationships 1) are directly applicable to remote sensing RTAs and 2) may be directly measured with accuracies within a few percent with commercially in situ available instrumentation [22,23]. Such relationships may have potential for developing novel ocean color inversion algorithms for next generation imagers, such as the Plankton, Aerosol, Cloud, and ocean Ecosystem (PACE) instrument with scheduled launch in 2022, which may also include a polarimeter with angular scattering information (http://pace.gsfc.nasa.gov/). With this application in mind, the relationship between μ¯ and bb/a is investigated across a representative range of VSF shapes.

2. RT simulations

Hydrolight (HL; www.sequoiascientific.com; see [1]) version 5.2.2 was used to solve the radiative transfer equation. HL obtains μ¯ and K by solving the following [12,24]:

(cKcosθ)L(θ)=02π02πL(θ')β(θ',φ'θ,φ)sinθ'dθ'dφ'
where β is the volume scattering function. This is a simplification of radiative transfer theory assuming asymptotic radiance fields are symmetric (only a function of θ) and decay exponentially [7,8,12]. HL solves Eq. (2) using an eigenmatrix equation for eigenfunctions (which give L) and eigenvectors (which give K) (see [1], section 9.6). The equation can also be solved numerically [12,14]. A directly computed solution from IOPs is more accurate than full radiative transfer calculations of light fields to some deep depth, as errors propagate and radiance levels often become too low for HL to manage.

All computations here were performed in the mid-visible at 532 nm. The interest here is in application of asymptotic light field theory in RTAs to use in remote sensing RTAs and inversion algorithms in the surface ocean. As such, we do not include Raman inelastic effects in the following. The effects of Raman on mesopelagic light fields have been rigorously addressed in other works [11,25]. Results are applicable to other wavelengths once water contributions are taken into account.

To compile the input matrix for IOPs, bb was first set to 0.001, 0.0013, 0.0020, 0.0040, 0.01, 0.1 m−1. Since backscattering of pure seawater bbw(532) at 20°C and 35 PSU is 0.00098 m−1 [23], these values correspond to ηbb = bbw / bb = 0.98, 0.75, 0.49, 0.25, 0.098, 0.0098, where ηbb represents the fractional seawater content of backscattering (cf. Morel and Gentili [26]). While in oceanic waters the contribution of pure seawater to total scattering is negligible, the seawater contribution is usually significant in the backward direction [27,28]. Values for bb/a were then set from 0.001 to 1 in twenty, logarithmically equal increments. Total absorption was calculated from bb/(bb/a) for each permutation. Pure water absorption was set to 0.0441 m−1 at 532 nm after [29].

Fournier-Forand (FF) analytical phase functions for the particulate fraction (βp(θ)/bp [30,31], latest version in [32]) were computed from bulk refractive index np, ranging from 1.04 to 1.24 in 0.02 steps, and particle size distribution (PSD) slope γ, ranging from 3.1 to 4 in 0.1 steps. A subset of aggregate phase functions (particles and seawater) used in HL are plotted in Fig. 1. See [33] for description of discretization of the phase functions and input in HL.

 figure: Fig. 1

Fig. 1 Selected total phase functions used in radiative transfer computations, (A) varying fractional water content for backscattering ηbb, keeping np and γ fixed at 1.1 and 3.4, respectively, for the FF phase function representing the particulate fraction, and (B) varying FF phase function input parameters (np, γ in legend) while keeping ηbb fixed at 0.0098 (i.e., particle dominated).

Download Full Size | PDF

For computations of vertical profiles of K, simple boundary conditions were chosen in the model runs, i.e. 1 m/s wind speed, a semi-empirical sky model based on RADTRAN-X, with 0% cloud coverage. The ocean was assumed to be infinitely deep and optically homogeneous. No other boundary conditions were considered because the asymptotic light field depends only on the IOPs.

3. Results

Figure 2 shows μ¯ as a function of inputs (np and γ) for the FF phase function and bb/a for the extreme values of ηbb considered, 0.98 and 0.0098. The largest bb/a assessed for ηbb = 0.98 was 0.0227, as bb was fixed at 0.001 m−1 and a could not be less than aw = 0.0441 m−1. At low bb/a values, the impact of absorption was dominant for all cases and μ¯ tended toward 1, with negligible effect from phase function shape. As bb/a increased, μ¯decreased as the effects of scattering became more significant. This effect became more significant for phase functions with lower np and γ, i.e., for “softer” particle fields with higher relative fractions of large particles. However, in general, phase function shape had minimal impact onμ¯. This has also been noted by Gordon et al. [6] who used Henyey-Greenstein phase function shapes.

 figure: Fig. 2

Fig. 2 Asymptotic average cosine μ¯ as a function of refractive index np and PSD slope γ inputs to Fournier-Forand phase function model. Horizontal grayscale surfaces represent fixed bb/a values (smallest and largest labeled with arrows at right side of each graph). Note 0.0227 was the largest bb/a possible at ηbb = 0.98 for the input IOP matrix (see text). Subset of results are shown for fractional seawater backscattering ηbb = 0.98 (left panels; bb fixed at 0.001 m−1) and 0.0098 (right panels; bb fixed at 0.1 m−1). Colored vertical surfaces in upper panels correspond to fixed np panels below.

Download Full Size | PDF

Results from all IOP matrix inputs are plotted in Fig. 3(a), with phase function shape variability manifested as the spread of points at any specific bb/a for a specific ηbb (see Fig. 3(a) legend). A 2nd order polynomial provided a satisfactory fit [μ¯=n=02pn(Logbba)n] to individual ηbb relationships (Table 1). With decreasing ηbb, absolute error of the fit increased slightly due to the increasing impact of phase function variability.

 figure: Fig. 3

Fig. 3 (a) All results for0μ¯ plotted with respect to bb/a, including full range of phase function shapes. A curve derived from the relationship obtained by Berwald et al. [5] is also shown (see text). (b) 4th order polynomial fits to all data and for ηbb ≤ 0.49. Vertical lines show maximum error for each bb/a for each fit.

Download Full Size | PDF

Tables Icon

Table 1. Fitting statistics for the average cosine for each ηbb as a function of bb/a after 2nd order polynomial fits, including full range of phase function shapes.

Good agreement was observed for results with ηbb ≤ 0.49, representative of coastal waters and the majority of surface waters in the open ocean [27,28]. Combining all results and results for ηbb ≤ 0.49, 4th order polynomial fits [μ¯=n=04pn(Logbba)n] were satisfactory (Fig. 3(b)), with fitting parameters provided in Table 2. Percent absolute error %δ for the fit to all data was a reasonable 3.4%. Residuals of μ¯ to fitted values were approximately normally distributed with standard deviation of 0.04 (not shown), indicating insignificant contribution of outliers and a highly representative relationship. A previous relationship between μ¯ and ω found by Berwald et al. [5] could also be overlaid in Fig. 3 by knowing B = 0.0183 for the Petzold phase function used in their HL simulations, where bb/a = 0.0183 / (ω−1 – 1).

Tables Icon

Table 2. Fitting statistics for the average cosine for combined ηbb relationships as a function of bb/a after 4th order polynomial fits, including full range of phase function shapes.

A further relationship was also developed that explicitly included the ηbb parameter for all data, which could be useful if ηbb could be known or estimated a priori. For example, estimation of ηbb from chlorophyll concentration derived from ocean color RRS would be comparable to the current convention for choosing a bidirectional reflectance distribution function (BRDF) correction to retrieve normalized water-leaving radiance from satellite imagery, following [34]. The resulting fit for μ¯ for all results was as follows:

μ¯(bba,ηbb)=n=02pn(ηbb)(Logbba)n=n=02m=02pn,mηbbm(Logbba)n=0.06790.020Logbba0.0808(Logbba)2+0.3859ηbb0.0837ηbbLogbba0.490ηbb(Logbba)2+0.573ηbb20.0743ηbb2Logbba+0.233ηbb2(Logbba)2.
Absolute errors %δ for this fit were 0.31 to 3.7 for ηbb ranging from 0.98 to 0.0098, respectively.

4. Discussion

A key motivation of this work was to assess the dependence of μ¯ on bb/a for use in RTAs for ocean color RRS algorithms such as those proposed by Zaneveld [2]. Results in Table 2 show μ¯ is a strong function of bb/a with minimal dependence on variability in the shape of phase functions for representative particle fields. Varying contribution of molecular scattering from pure seawater for different turbidity levels complicates the relationship, but %δ was still only 3.4% when considering the full range of turbidities, i.e., the full range of ηbb. If ηbb can be estimated a priori [34], then errors may be further reduced, particularly for very clear waters (Table 1). If remotely sensed multi-angle polarimetry becomes available, as is proposed for the future PACE mission, then it may be possible to glean enough phase function shape information with that sensor to directly estimate ηbb, as the angular shapes of the water and particle contributions in the backward trend in opposite directions (e.g [35].). For the majority of waters, however, the strong agreement in results for ηbb ≤ 0.49 will be representative.

HL computation results here also allow assessment of a central assumption in the Zaneveld approximations, namely that KLu near the surface is approximately equivalent toK. Zaneveld [2] argued these parameters should be in close agreement based on measurements made in the 1960’s and 1970’s of upward radiance fields, showing a near constant shape and attenuation rate with depth [9,36–38]. Figure 4(a) shows profiles of the % difference between KLu just below the surface (0-) and K for a relatively turbid case (ηbb = 0.098), for a range of bb/a and particle phase function shapes. For these phase functions, B varied between 0.5% and 2%. Over the full range of variables tested, maximal percent difference was within 3% for the majority of conditions, with the sun at zenith. Another recent study by [11], also employing HL computations, found similar results in the asymptotic regime for water columns with vertical structure in IOPs, but only in the blue spectral region, as their work included the effects of inelastic Raman scattering. Figure 4(b) shows the effects of varying solar zenith angle θz for a range of ηbb, bb/a and particle phase function shapes. As θz increased, % differences also increased, but in a predictable way for all cases (4th order polynomial fit provided in Fig. 4(b)). Good predictability implies KLu can be directly tied to K and thus to bb and a with low uncertainties.

 figure: Fig. 4

Fig. 4 Percent difference between the attenuation of upwelling nadir radiance KLu and asymptotic attenuation K (a) for ηbb = 0.098 and γ = 3.4, with a range of np and bb/a given in legend; (b) as a function of θz for a range in ηbb, just below the air-sea interface (0-), with symbol colors matching legend in (a). Fit provided for ηbb = 0.098 (solid black curve). The physical approximation from Eq. (4) is also plotted (dashed red curve).

Download Full Size | PDF

The general relationship found in Fig. 4(b) has direct physical meaning. We may assume Kd and KLu may be similar, as nadir radiance is derived from backscattered downward irradiance. For a flat water surface, we may assume Kdw) ≈Kd(0°) / cos(θw), where θw is the solar zenith angle in the water, so that KLuw) ≈KLu(0°) / cos(θw). Setting θw = 0° (as in Fig. 4(a)), results in KLu(0°) ≈K, so that:

[KLu,0(θw)K]K[1cos(θw)1]
This approximation is plotted in Fig. 4(b) after refracting θz to θw through the air-water interface, showing reasonable agreement with HL results for θz < ~65°.

Results thus provide justification for the targeted application of K and μ¯ in ocean color RTAs for the surface ocean. Assessing the performance of ocean color RTAs using the above representations of μ¯ in terms of bb/a with associated uncertainties is the subject of other ongoing work.

Funding

NASA Ocean Biology and Biogeochemistry, PACE Science Team (NNX15AN17G); Harbor Branch Oceanographic Institute Foundation.

Acknowledgments

The authors thank Ron Zaneveld and Curt Mobley for helpful discussions. Three anonymous reviewers provided valuable comments on the manuscript; in particular, we greatly appreciate the input of one reviewer who suggested the physical relationship provided in Eq. (4).

References and links

1. C. D. Mobley, Light and Water: Radiative Transfer in Natural Waters (Academic, 1994), p. 592.

2. J. R. Zaneveld, “A theoretical derivation of the dependence of the remotely sensed reflectance of the ocean on the inherent optical properties,” J. Geophysical Res. 100, 13135 (1995).

3. J. R. Zaneveld, “An asymptotic closure theory for irradiance in the sea and its inversion to obtain the inherent optical properties,” Limnol. Oceanogr. 34(8), 1442–1452 (1989). [CrossRef]  

4. R. W. Preisendorferr, “Theoretical proof of the existence of characteristic diffuse light in natural waters,” J. Mar. Res. 18, 1–9 (1959).

5. J. Berwald, D. Stramski, C. D. Mobley, and D. A. Kiefer, “Influences of absorption and scattering on vertical changes in the average cosine of the underwater light field,” Limnol. Oceanogr. 40(8), 1347–1357 (1995). [CrossRef]  

6. H. R. Gordon, K. Ding, and W. Gong, “Radiative transfer in the ocean: computations relating to the asymptotic and near-asymptotic daylight field,” Appl. Opt. 32(9), 1606–1619 (1993). [CrossRef]   [PubMed]  

7. J. R. V. Zaneveld and H. Pak, “Some aspects of the axially symmetric submarine daylight field,” J. Geophys. Res. 77(15), 2677–2680 (1972). [CrossRef]  

8. N. Højerslev and J. R. V. Zaneveld, “A theoretical proof of the existence of the submarine asymtotic daylight field,” (Københavns Univ. Inst. Fysisk Oceanog., Copenhagen, 1977).

9. N. G. Jerlov, Marine Optics (Elsevier, 1976).

10. J. Ronald and V. Zaneveld, “Remotely sensed reflectance and its dependence on vertical structure: a theoretical derivation,” Appl. Opt. 21(22), 4146–4150 (1982). [CrossRef]   [PubMed]  

11. L. Li, D. Stramski, and R. A. Reynolds, “Characterization of the solar light field within the ocean mesopelagic zone based on radiative transfer simulations,” Deep-Sea Res. 87, 53–69 (2014). [CrossRef]  

12. L. Prieur and A. Y. Morel, “Etude theorique du regime asymptotique: Relations entre characteristiques optiques et coefficient d’extinction relatif a la penetration de la lumiere du jour,” Cah. Oceanogr. 23, 35–48 (1971).

13. V. A. Timofeeva, “Optical characteristics of turbid media of the sea-water type,” Izv., Atmos. Ocean. Phys. 7, 863–865 (1971).

14. G. W. Kattawar and G. N. Plass, “Asymptotic radiance and polarization in optically thick media: ocean and clouds,” Appl. Opt. 15(12), 3166–3178 (1976). [CrossRef]   [PubMed]  

15. H. R. Gordon, O. B. Brown, and M. M. Jacobs, “Computed relationships between the inherent and apparent optical properties of a flat homogeneous ocean,” Appl. Opt. 14(2), 417–427 (1975). [CrossRef]   [PubMed]  

16. E. Boss, W. H. Slade, M. Behrenfeld, and G. Dall’Olmo, “Acceptance angle effects on the beam attenuation in the ocean,” Opt. Express 17(3), 1535–1550 (2009). [CrossRef]   [PubMed]  

17. K. J. Voss and R. W. Austin, “Beam-attenuation measurements error due to small-angle scattering acceptance,” J. Atmos. Ocean. Technol. 10(1), 113 (1993). [CrossRef]  

18. J. R. Zaneveld, J. C. Kitchen, A. Bricaud, and C. M. Moore, “Anlysis of in situ spectral absorption meter data,” in Ocean Optics XI, (SPIE, 1992).

19. H. R. Gordon, O. B. Brown, R. H. Evans, J. W. Brown, R. C. Smith, K. S. Baker, and D. K. Clark, “A semi-analytic radiance model of ocean color,” J. Geophys. Res. 93(D9), 10909–10924 (1988). [CrossRef]  

20. A. Morel, “Diffusion de la lumière par les eaux de mer; résultats expérimentaux et approche théorique,” AGARD Lect. Ser., 3.1.1.-3.1.76 (1973).

21. M. S. Twardowski, E. Boss, J. B. Macdonald, S. Pegau, A. H. Barnard, and J. R. V. Zaneveld, “A model for estimating bulk refractive index from the optical backscattering ratio and the implications for understanding particle composition in case I and case II waters,” J. Geophys. Res. 106(C7), 14129–14142 (2001). [CrossRef]  

22. N. Stockley, R. Röttgers, D. McKee, I. Lefering, J. Sullivan, and M. Twardowski, “Improving in situ absorption measurement uncertainties for ocean color remote sensing validation,” Opt. Express. submitted.

23. J. M. Sullivan, M. S. Twardowski, J. R. V. Zaneveld, and C. M. Moore, Measuring Optical Backscattering in Water, Light Scattering Reviews 7 (Springer Praxis Books, 2013).

24. C. D. Mobley, “Radiative Transfer Theory - The Asymptotic Radiance Distribution” (2010), retrieved http://www.oceanopticsbook.info/view/radiative_transfer_theory/the_asymptotic_radiance_distribution.

25. J. Berwald, D. Stramski, C. D. Mobley, and D. A. Kiefer, “Effect of Raman scattering on the average cosine and diffuse attenuation coefficient of irradiance in the ocean,” Limnol. Oceanogr. 43(4), 564–576 (1998). [CrossRef]  

26. A. Morel and B. Gentili, “Diffuse reflectance of oceanic waters: its dependence on Sun angle as influenced by the molecular scattering contribution,” Appl. Opt. 30(30), 4427–4438 (1991). [CrossRef]   [PubMed]  

27. M. Twardowski, H. Claustre, S. Freeman, D. Stramski, and Y. Huot, “Optical backscattering properties of the “clearest” natural waters,” Biogeosciences 4(6), 1041–1058 (2007). [CrossRef]  

28. P. J. Werdell, B. A. Franz, J. T. Lefler, W. D. Robinson, and E. Boss, “Retrieving marine inherent optical properties from satellites using temperature and salinity-dependent backscattering by seawater,” Opt. Express 21(26), 32611–32622 (2013). [CrossRef]   [PubMed]  

29. R. M. Pope and E. S. Fry, “Absorption spectrum (380-700 nm) of pure water. II. Integrating cavity measurements,” Appl. Opt. 36(33), 8710–8723 (1997). [CrossRef]   [PubMed]  

30. G. Fournier and J. L. Forand, “Analytic phase function for ocean water,” in Ocean Optics XII, (Proc. SPIE, 1994), pp. 194–201.

31. G. Fournier and M. Jonasz, “Computer-based underwater imaging analysis,” in Airborne and In-water Underwater Imaging, (Proc. SPIE, 1999), pp. 62–77.

32. M. Jonasz and G. Fournier, Light Scattering by Particles in Water: Theoretical and Experimental Foundations, 1st ed. (Elsevier, 2007).

33. A. Tonizzo, M. Twardowski, S. McLean, K. Voss, M. Lewis, and C. Trees, “Closure and uncertainty assessment for ocean color reflectance using measured volume scattering functions and reflective tube absorption coefficients with novel correction for scattering,” Appl. Opt. 56(1), 130–146 (2017). [CrossRef]  

34. A. Morel, D. Antoine, and B. Gentili, “Bidirectional reflectance of oceanic waters: accounting for Raman emission and varying particle scattering phase function,” Appl. Opt. 41(30), 6289–6306 (2002). [CrossRef]   [PubMed]  

35. E. Boss and W. S. Pegau, “Relationship of light scattering at an angle in the backward direction to the backscattering coefficient,” Appl. Opt. 40(30), 5503–5507 (2001). [CrossRef]   [PubMed]  

36. J. E. Tyler, “Observed and computed path radiance in the underwater light field,” J. Mar. Res. 18, 157–167 (1960).

37. N. G. Jerlov and M. Fukuda, “Radiance distribution in the upper layers of the ocean,” Tellus 12(3), 348–355 (1960). [CrossRef]  

38. V. A. Timofeeva, Optics of Turbid Waters (Results of Laboratory Studies), Optical Aspects of Oceanography (Academic, 1974).

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 (4)

Fig. 1
Fig. 1 Selected total phase functions used in radiative transfer computations, (A) varying fractional water content for backscattering ηbb, keeping np and γ fixed at 1.1 and 3.4, respectively, for the FF phase function representing the particulate fraction, and (B) varying FF phase function input parameters (np, γ in legend) while keeping ηbb fixed at 0.0098 (i.e., particle dominated).
Fig. 2
Fig. 2 Asymptotic average cosine μ ¯ as a function of refractive index np and PSD slope γ inputs to Fournier-Forand phase function model. Horizontal grayscale surfaces represent fixed bb/a values (smallest and largest labeled with arrows at right side of each graph). Note 0.0227 was the largest bb/a possible at ηbb = 0.98 for the input IOP matrix (see text). Subset of results are shown for fractional seawater backscattering ηbb = 0.98 (left panels; bb fixed at 0.001 m−1) and 0.0098 (right panels; bb fixed at 0.1 m−1). Colored vertical surfaces in upper panels correspond to fixed np panels below.
Fig. 3
Fig. 3 (a) All results for0 μ ¯ plotted with respect to bb/a, including full range of phase function shapes. A curve derived from the relationship obtained by Berwald et al. [5] is also shown (see text). (b) 4th order polynomial fits to all data and for ηbb ≤ 0.49. Vertical lines show maximum error for each bb/a for each fit.
Fig. 4
Fig. 4 Percent difference between the attenuation of upwelling nadir radiance KLu and asymptotic attenuation K (a) for ηbb = 0.098 and γ = 3.4, with a range of np and bb/a given in legend; (b) as a function of θz for a range in ηbb, just below the air-sea interface (0-), with symbol colors matching legend in (a). Fit provided for ηbb = 0.098 (solid black curve). The physical approximation from Eq. (4) is also plotted (dashed red curve).

Tables (2)

Tables Icon

Table 1 Fitting statistics for the average cosine for each ηbb as a function of bb/a after 2nd order polynomial fits, including full range of phase function shapes.

Tables Icon

Table 2 Fitting statistics for the average cosine for combined ηbb relationships as a function of bb/a after 4th order polynomial fits, including full range of phase function shapes.

Equations (4)

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

μ ¯ ( z )= 0 2π 0 π L( θ,φ,z )cosθsinθdθdφ 0 2π 0 π L( θ,φ,z )sinθdθdφ = E d E u E od + E ou   ,
( c K cosθ ) L ( θ )= 0 2π 0 2π L ( θ' )β( θ ' , φ ' θ,φ )sinθ'dθ'dφ'
μ ¯ ( b b a , η bb )= n=0 2 p n ( η bb ) ( Log b b a ) n = n=0 2 m=0 2 p n,m η bb m ( Log b b a ) n =0.0679 0.020Log b b a 0.0808 ( Log b b a ) 2 +0.3859 η bb 0.0837 η bb Log b b a 0.490 η bb ( Log b b a ) 2 + 0.573 η bb 2 0.0743 η bb 2 Log b b a +0.233 η bb 2 ( Log b b a ) 2 .
[ K Lu,0 ( θ w ) K ] K [ 1 cos( θ w ) 1 ]
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.