J Rse 2021 112414

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Remote Sensing of Environment 259 (2021) 112414

Contents lists available at ScienceDirect

Remote Sensing of Environment


journal homepage: www.elsevier.com/locate/rse

Remote detection of marine debris using satellite observations in the visible


and near infrared spectral range: Challenges and potentials
Chuanmin Hu *
University of South Florida, 140 Seventh Avenue, South, St. Petersburg, FL 33701, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Despite the importance of remote detection of marine debris, nearly all published studies are focused on either
Remote sensing controlled experiments, or Sentinel-2 data with mixed band resolutions that are subject to large uncertainties. To
Marine debris date, key questions such as the following have not been addressed adequately: To what extent can the various
Microplastics
forms of marine debris be remotely detected and differentiated through satellite observations in the visible and
Resolution
Sensitivity
near infrared (NIR) spectral range, and how? Here, using published reflectance spectra of various types of
Spectral bands floating matters, I address these questions through sensitivity analyses, simulations, and spectral analyses of
Sentinel-2 satellite images. While the study is by no means comprehensive, several observations can still be made. First, it
PlanetScope/DOVE appears impossible to remotely detect marine microplastics from all existing and planned optical sensors. This is
simply because the contribution of these particles to the sensor signal, even when they are aggregated on the
water surface at the reported maximum particle density, is at least 60 times lower than the required signal
(~0.2% subpixel coverage) and 20 times lower than the sensor noise for a sensor with a signal-to-noise ratio
(SNR) of 200. In contrast, detecting macroplastics and other debris is possible when they form large patches
along ocean fronts or windrows. Second, assuming a SNR of 200, discriminating large patches of marine debris
from floating algae is only possible with a subpixel coverage of >0.3%. These threshold values are based on the
sensor SNRs only, and they represent the lower bounds of detection and discrimination, respectively. The real
threshold values above which a detection or discrimination is possible also depend on the observing conditions,
and therefore higher. Third, currently, Sentinel-2 MSI (Multi Spectral Instrument) sensors provide an optimal
trade between resolution and coverage, yet MSI sensors have SNRs <200, and interpretation of the MSI spectra
requires extra caution due to variable spatial resolutions in different bands, among other factors. From the
perspective of pure spectroscopy, it is possible to discriminate floating algae from non-algae floating matters but
difficult to differentiate the type of the latter (either plastic or non-plastic debris, foam, etc) because different
non-algae floating matters all show relatively flat reflectance spectral shapes in the vis-NIR spectral range.
Finally, based on these results, recommendations are made on algorithm designs and sensor designs, for example
spectral analysis should be performed over the difference spectra to minimize the impact of variable subpixel
coverage, and certain spectral bands are more important than others for the remote detection of marine debris.

1. Background on remote sensing of marine debris recognized in recent years, together with reports of increased land-based
sources of marine debris (Law et al., 2010). In addressing the pressing
By definition, marine debris refers to any persistent solid material needs of understanding the distributions and fates of various forms of
that is disposed (or abandoned) in the marine environment by natural marine debris, inter-governmental agencies, environmental groups, and
processes (including natural disasters such as Tsunami) and human ac­ the research community have started to work together toward an inte­
tivities, for example microplastic particles, plastic bags or bottles, grated marine debris observing network (Maximenko et al., 2019), of
cigarette butts, foam take-out containers, balloons, fishing gear, tree which remote sensing has been emphasized for its synoptic and frequent
branches/leaves, wood, among others. The impacts of marine debris on coverage (see also Maximenko et al., 2017). The European Space Agency
marine animals and marine ecosystems have been increasingly is particularly keen on funding several pioneering efforts (Topouzelis

* Corresponding author.
E-mail address: huc@usf.edu.

https://doi.org/10.1016/j.rse.2021.112414
Received 3 November 2020; Received in revised form 20 February 2021; Accepted 22 March 2021
Available online 5 April 2021
0034-4257/© 2021 Elsevier Inc. All rights reserved.
C. Hu Remote Sensing of Environment 259 (2021) 112414

et al., 2019 & 2020), with more projects to start in 2020 and new pro­ interpretation to separate debris from Sargassum. The NIR reflec­
jects funded recently by the U.S. National Aeronautics and Space tance of the debris endmember is about ~0.04, suggesting mixed
Administration (NASA). pixels.
Among the key questions are: can the various forms of marine debris • Biermann et al. (2020) also used Sentinel-2 MSI data and examined
be remotely sensed and quantified? If so, how? What are the optimal spectral shapes of various endmembers to classify pixels containing
remote sensing approaches to address the various user needs? marine debris. Similar to those in Kikaki et al. (2020), the NIR
Several pioneering studies attempted to address these questions from reflectance of the debris endmember is about ~0.04, suggesting
different aspects, with reviews provided by Maximenko et al. (2017 & mixed pixels.
2019) and by Martinez-Vicente et al. (2019). Among the various remote • Matthews et al. (2017) showed images from high-resolution com­
sensing techniques such as synthetic aperture radar (Arii et al., 2014) mercial satellites and airborne digital photos where marine debris
and thermal infrared sensors (Garaba et al., 2020; Goddijn-Murphy and after a Tsunami near Japan could be detected as bright pixels in the
Williamson, 2019), passive optical remote sensing in the visible, near- images/photos.
infrared (NIR), and shortwave-infrared (SWIR) spectral range is
perhaps the most used because most medium- (a few hundred meters) or Field surveys:
high-resolution (meters to tens of meters) optical sensors are equipped
with spectral bands in this range. While it is beyond the scope of this • Law et al. (2010) showed distributions and areal density of marine
paper to present a comprehensive review on the vis-NIR-SWIR optical microplastics particles in the North Atlantic, based on >6600 surface
remote sensing on this subject, the following list presents key examples net tows from 1986 to 2010. Most of the data showed <200,000
of these studies: plastic pieces km− 2. There was only one extreme data point with 26
Spectral reflectance measurements in the lab: million pieces km− 2 (Fig. 5 of Maximenko et al., 2017).
• Cózar et al. (2014) analyzed plastic debris data collected from 1127
• Hu et al. (2015) showed the first measurements of spectral reflec­ surface net tows in 442 sites, and found that most plastic pieces have
tance of various large artificial (man-made) marine debris patches their size below 5 mm (i.e., microplastics) with a log-normal distri­
(Fig. 1a). Reflectance is rather “flat” (i.e., without narrow-band bution and a mode of ~2 mm.
features) for up to 900 nm (Fig. 1b). • Eriksen et al. (2014) compiled a dataset of marine plastic debris
• Garaba and Dierssen (2018, 2020) showed the first laboratory measured at 1571 stations from 680 net tows and 891 visual survey
measurements of marine microplastics (Fig. 1b) and also laboratory transects. They showed that the dominant majority of all compiled
measurements of various macroplastics in the vis-NIR-SWIR wave­ and modeled particle density is <1 million pieces km− 2 for particles
lengths. Reflectance is also “flat” for up to 900 nm, with some C–H <4.75 mm in size.
absorption features in the SWIR wavelengths. • van Sebille et al. (2015) compiled a first global inventory of small
floating plastic debris based on field measurements by many groups.
Airborne or satellite measurements of man-made targets under controlled Results indicate that for the vast majority of data, particle density is
environments: <1 million pieces km− 2.

• Garaba et al. (2018) applied an airborne imager in the SWIR wave­ Remote sensing perspectives:
lengths to measure man-made plastic patches.
• Topouzelis et al. (2019 & 2020) used Sentinel-2 MSI images to detect • Goddijn-Murphy et al. (2018) showed a theoretical concept on
man-made plastic targets under a controlled experimental setting. hyperspectral remote sensing of marine macroplastics based on
Spectral shapes of MSI pixels over these targets show a spike around radiative transfer theory. The concept was evaluated later using a
833 or 741 nm, and the NIR reflectance magnitude of ~0.04 in­ carefully designed experiment (Goddijn-Murphy and Dufaur, 2018),
dicates mixed pixels. yet no satellite-based application has been demonstrated.
• Martinez-Vicente et al. (2019) provided a perspective on sensor re­
Applications to satellite measurements over oceanic environments: quirements for remote sensing of marine plastic debris, with a focus
on the NIR and SWIR wavelengths only.
• Kikaki et al. (2020) used Sentinel-2 MSI and Landsat-8 OLI to search
for plastic debris in the Caribbean Sea, based on spectral

Fig. 1. (a) Artificial (man-made) garbage patches in a preliminary experiment. (b) Their corresponding reflectance measured in Tampa Bay using a hand-held
spectrometer (Hu et al., 2015). In (b), reflectance of wet microplastics (red curve, right y-axis) was obtained from Garaba and Dierssen (2020), who provided a
hyperspectral dataset (vis-NIR-SWIR) for different types of plastics (microplastics, macroplastics, plastics with different colors). (For interpretation of the references
to colour in this figure legend, the reader is referred to the web version of this article.)

2
C. Hu Remote Sensing of Environment 259 (2021) 112414

• Maximenko et al. (2017) provided a review on remote sensing of pigment absorption. Indeed, once normalized to 25% reflectance in the
marine debris from a NASA workshop, where the strengths and NIR (e.g., around 860 nm), all these spectra appear rather similar
limitations of different technologies are discussed. (Fig. 1b). Therefore, in this study, collectively they may represent one
• Maximenko et al. (2019) provided a comprehensive review of all marine debris endmember.
possible techniques in observing marine debris, including in situ In the ocean, other forms of floating matters also exist. These include
sensing, remote sensing, and modeling, with future approaches Sargassum fluitans and Sargassum natans (Wang et al., 2018), Sargassum
outlined toward addressing the identified challenges (for example, horneri (Qi et al., 2017), Ulva (Hu et al., 2017), cyanobacteria Tricho­
sensor requirement). desmium (Hu et al., 2010; McKinna, 2010), emulsified oil (Lu et al.,
2020), green Noctiluca, red Noctiluca, pumice rafts (Qi et al., 2020), and
Although field surveys of marine debris started in the 1970s, remote foams (whitecaps) (Dierssen, 2019). Some of these have been measured
sensing studies did not start until much more recently. From these pio­ in the field, with in situ hyperspectral reflectance being available (e.g.,
neering studies above, it appears that the feasibility of using routine Fig. 2 of Hu et al., 2015) but others have only been assessed using multi-
satellite platforms (e.g., Sentinel-2, Landsat-8) to detect large patches of band satellite data (e.g., pumice rafts). Some of these spectra are
marine debris has been proven (Biermann et al., 2020; Kikaki et al., compiled in Figs. 2a & 2b, which show typical spectral shapes of non-
2020), and the rest is simply to implement a monitoring system to debris floating matters. Note that except for the spectra in Fig. 2a
combine the published algorithms and satellite data for global moni­ whose magnitudes in the NIR are >0.2 and therefore may be used as
toring. However, this does not appear to be the case, as shown below. In endmember spectra (i.e., 100% coverage within a pixel) for the spectral
particular, it appears impossible to remotely sense microplastics mixing experiments below, the NIR reflectance in Fig. 2b are all <0.1,
regardless of the optical sensors or resolutions. Then, to what extent can suggesting that these spectra derived from medium-resolution (300-m)
we sense macroplastics and other forms of marine debris using satellite image pixels are mixed between floating matters and water (about 16% -
remote sensing in the vis-NIR-SWIR spectral range, and how? 40% subpixel floating matter coverage), therefore cannot be used as
With a focus on the vis-NIR wavelengths, the objective of this paper endmembers. Also note that except for red Noctiluca and pumice rafts, all
is to provide a perspective on whether and under what conditions spectra in Figs. 2a & 2b have the typical reflectance trough around 670
various forms of marine debris can be detected and discriminated nm due to chlorophyll a absorption (black arrows), and they also show
against other floating matters. The paper is based on published end­ the typical vegetation red-edge reflectance above 700 nm. These char­
member spectra, sensor sensitivity, simulation experiments, and spectral acteristics are different from the marine debris endmember spectra, and
analyses of Sentinel-2 data for demonstration purpose, from which such characteristics may form the basis for spectral discrimination be­
recommendations are made on sensor design as well as on algorithms tween floating algae (pigment rich) and floating debris (no pigment,
and approaches toward vis-NIR remote sensing of marine debris. The use Fig. 1) using the vis-NIR bands. Such a concept is demonstrated in the
of SWIR wavelengths is also discussed. Furthermore, because marine simulations below using “plastic bags” (Fig. 1b) and Sargassum (Fig. 2a)
debris is an emerging topic in the scientific literature across many dis­ to represent marine debris and floating algae, respectively. In such
ciplines, satellite remote sensing is only one of the many tools to help simulations, for water endmember spectra, two scenarios were selected
answer multi-disciplinary questions (Maximenko et al., 2019); inter­ to represent clear water and turbid water, respectively (Fig. 2c). These
ested readers are therefore referred to the published literature on other spectra were collected in the Florida Keys (May 12, 2013, 24.5333oN,
aspects of marine debris monitoring. 81.4016oW, chlorophyll concentration ~ 0.1 mg m− 3) and near Tampa
Bay (August 24, 2012, 27.9313oN, 82.6512oW. chlorophyll concentra­
2. Field and satellite data tion ~ 12.7 mg m− 3), respectively, using a hand-held spectrometer
following NASA Ocean Optics protocols.
2.1. Endmember spectra

To date, laboratory or in situ measurements of optical properties of 2.2. Satellite data


marine debris are scarce, with the exception of some artificial (man-made)
garbage patches and field-collected micro plastic particles (Fig. 1). One While there are currently many satellite sensors in orbit, Sentinel-2
common characteristic from these various forms of marine debris is that MSI was selected to represent high-resolution optical sensors because
they are all spectrally flat without narrow-band features in the visible – it provides a trade between spatial resolution (10–20 m) and revisit
NIR wavelengths (400–900 nm). Although only one example of micro­ frequency (2–3 days) for most coastal waters. MSI covers wavelengths of
plastics spectra from Garaba and Dierssen (2020) is shown in Fig. 1b to vis-NIR-SWIR, suitable for detecting and differentiating small floating
illustrate such spectral characteristic, all their miscroplastics spectra show matters. The use of MSI is also to follow a number of recent studies by
the same characteristic. This is because of lack of narrow-band vegetative Kikaki et al. (2020), Biermann et al. (2020), and Topouzelis et al. (2019
& 2020) for both real applications and controlled experiments. MSI has

Fig. 2. Hyperspectral and multi-spectral reflectance of various floating matters from (a) in situ measurements (Hu et al., 2015 for Sargassum; Hu et al., 2017 for Ulva;
McKinna, 2010 for Trichodesmium) and (b) OLCI measurements (Qi et al., 2020), respectively. Other than red Noctiluca and pumice rafts, all floating matters show the
pigment absorption around 670 nm (black arrows). (c) Clear-water and turbid-water endmember spectra used in the simulated mixing experiments. (For inter­
pretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

3
C. Hu Remote Sensing of Environment 259 (2021) 112414

the following spectral bands: 443 (60), 492 (10), 560 (10), 665 (10), 704 estimated using a single wavelength in the NIR (Eq. 3), or a combination
(20), 741 (20), 783 (20), 841 (10), 865 (20), 1614 (20), and 2202 nm of these wavelengths (e.g., floating algae index, Hu, 2009). Because the
(20), where the numbers in the parentheses represent their ground latter involves more bands and therefore more noise (Qi and Hu, 2021),
resolutions in meters. the lower detection limit is from a single band in the NIR, where the
Level-1 MSI data were downloaded from the U. S. Geological Survey, spatial contrast is the highest between floating matters and water.
and processed using the Acolite software to generate Rayleigh corrected Once a pixel is determined to contain a certain type of floating
reflectance (Rrc(λ), dimensionless), from which Red-Green-Blue and matter, there are several ways to discriminate the type, including a
False-colour RGB composite imagery were generated. In the FRGB im­ similarity index between the pixel’s ΔR and (RFM – RW) where RFM and
agery, a NIR band was used to replace the green band in the RGB im­ RW are from the established spectral library (e.g., Figs. 1 & 2) and water
agery, making it suitable to detect floating matters with enhanced NIR pixels, respectively. Because the spectral shapes of RW change the most
reflectance (Qi et al., 2020). in the blue-green wavelengths in natural waters, the simplest way is to
restrict the similarity analysis to the red-NIR-SWIR wavelengths of RFM.
3. Methods Comparing Figs. 1 and 2a, the most prominent difference between ma­
rine debris and floating algae occurs between 670 nm and ~ 750 nm: the
Regardless of the floating matter type (either marine debris, floating former is spectrally flat while the latter has a sharp increase in the NIR (i.
algae, or other types of floating matters), remote detection always re­ e., the red-edge reflectance). Therefore, the reflectance difference be­
quires two steps. Step 1 is to detect a spatial anomaly, i.e., some pixels tween the NIR and red bands (Richardson and Everitt, 1992) can be used
“stand out” from their nearby background waters. Step 2 is to spectrally to discriminate marine debris from floating algae:
differentiate the pixel type from the spatial anomaly. Step 2 is impos­
NRD = RNIR − Rred
sible without Step 1. In simpler words, the two steps can be shortened as:
1) is there “something”? 2) what is that “something”? If the amount of Here NRD stands for NIR-red difference. For a pixel containing χ
floating matter is to be quantified, then a third step is required to address floating matter and (1-χ) water, there is.
the question of how much is that “something.” Here, because Step 3 is [( ) ( )]
ΔNRDFM = ΔRNIR − ΔRred = χ RFM NIR − RFM red + RW NIR –RW red (4)
relatively easy, the focus is on Steps 1 and 2.
Using image examples and simulations, Hu et al. (2015) showed that Then, to be able to separate marine debris (MD) and floating algae
Step 1 relies only on the sensor’s sensitivity (i.e., signal-to-noise ratio or (FA), their difference in ΔNRD should be significantly higher than noise,
SNR), while Step 2 requires specific spectral bands depending on the i.e.,
targeted floating matter type and on the selected algorithms. Although [( ) ( )]
that study was targeted on Sargassum, the principle is the same for ΔNRDFA − ΔNRDMD = χ RFA NIR –RFA red − RMD NIR –RMD red >> noise
marine debris and all other floating matters, therefore these steps will be = 2×2 σ
used for the sensitivity analysis below. (5)
Here the first 2 in Eq. 5 represents statistical significance, and the
3.1. Sensitivity analysis second 2 in Eq. 5 represents the cumulative noise using the square-root
√̅̅̅
rule (two bands, two types, therefore 4). The discrimination limit is
To be able to “stand out” in an image, a pixel needs to be significantly therefore:
different from the surrounding pixels. Mathematically, this requires (Qi /[( ) ( )]
and Hu, 2021): χdis ≥ 4σ RFA NIR –RFA red − RMD NIR –RMD red (6)
√̅̅̅
ΔR > 2 2σ (1) Comparing with Eq. 3, Eq. 6 is very similar except for the factor of 4
√̅̅̅
√̅̅̅ instead of 2 2 because of more spectral bands involved.
where σ is the sensor noise in a pixel, 2 is to account for noise
propagation in pixel differencing between the target pixel and nearby
reference pixel (in this case, noise is the square root of sum squares from 3.2. Simulation experiment
√̅̅̅
two pixels, therefore the 2 term), 2 is to make the difference statisti­
cally significant (i.e., 2 times noise), and ΔR is the difference between In the experiment, reflectance of a pixel mixed with floating matter
the target pixel and nearby reference pixel (i.e., water pixel): and water (i.e., mixed pixel) was estimated with their endmember
spectra and χ. Then, the pixel’s spectra were compared with the end­
ΔR = RT –RW = [χ RFM + (1 − χ) RW ]–RW = χ(RFM –RW ) (2)
member spectra to determine their spectral similarity.
The similarity between two spectra was estimated using a spectral
where “T” stands for target, “FM” is for floating matter, “W” is for water,
angle measure (SAM, Kruse et al., 1993). The choice of SAM over other
and χ (0.0%–100%) is the subpixel proportion of floating matter. For
similarity measures is because SAM is based on spectral shape only. This
simplicity, the wavelength dependence of R is omitted.
is important because χ for marine debris or other floating matters is
From Eqs. 1 and 2, once σ is known, the subpixel detection limit, χdet,
often very small and also variable, thus the reflectance magnitude of the
can be estimated as:
mixed pixel should be deemphasized.
√̅̅̅ /
χdet ≥ 2 2σ (RFM –RW ) (3) Mathematically, SAM is the angle between two spectral vectors,
defined as (Kruse et al., 1993):
From Eq. 2, assuming the endmember spectra of RFM and RW are [( ∑ )/(√̅̅̅̅̅̅̅̅̅̅̅̅ ∑ )]
∑ √̅̅̅̅̅̅̅̅̅̅̅̅
relatively stable, the spectral shape of ΔR is determined between RFM SAM (degrees) = cos− 1 xi yi x2i y2i (7)
and RW with equal weights, and the shape does not change with χ. In
other words, both RFM and RW contribute to ΔR with the same weights where x and y represent two spectra and the summation is for band
regardless of χ. In contrast, their weights to RT are not equal but number i from 1 to N. SAM = 0o means two parallel spectra in log space
determined by χ and (1-χ), respectively. This makes the spectral shape of (i.e., identical spectral shapes), while SAM = 90o means perpendicular
RT being dominated by RW when χ is very small (e.g., < 5% or < 10%) as spectra (i.e., completely different spectral shapes). SAM < 5o indicates
for the case of marine debris. that the two spectra are very similar (Garaba and Dierssen, 2018).
In practice, because the spectral contrast between floating matters Four (4) endmember spectra were selected in the experiment:
and water is mostly in the red-NIR-SWIR wavelengths, χdet can be Sargassum (Fig. 2a); plastic bags (Fig. 1); clear water; turbid water

4
C. Hu Remote Sensing of Environment 259 (2021) 112414

(Fig. 2c). While in nature other types of floating matters also exist and /
water reflectance can also change in both shape and magnitude, for χdis H ≥ 4σH 0.25 ≈ 0.3%
demonstration purpose, the first two are used to represent floating /
vegetation and floating debris, respectively, and the last two are used to χdis MSI ≥ 4σMSI 0.25 ≈ 1.0% (9)
represent typical clear open-ocean waters and turbid coastal waters, These estimates are based on the assumption that 1) for both floating
respectively. plastics (or other debris) and floating algae, their NIR reflectance is
The hyperspectral data of the 4 endmembers were first resampled to ~0.25 for χ = 100% (Fig. 1b & Fig. 2a), and 2) floating plastics (or other
MSI wavelengths using their relative spectral response (RSR) functions, debris) are spectrally flat in the red and NIR wavelengths (Fig. 1b).
and then mixed using different subpixel coverage (χ from 1% to 20%). Indeed, Garaba and Dierssen (2018 & 2020) showed that wet micro­
Then, both the mixed spectra, Rχ, and their contrasts from water, ΔRχ, plastics pieces have reflectance ~0.25 in the NIR and < 0.1 in the SWIR
were compared with the 2 floating matter endmembers to determine wavelengths for χ = 100%, and all spectra are rather flat between red
their similarities using Eq. 7. and NIR wavelengths. Therefore, such assumptions appear realistic.
In the above simulation experiment, because MSI bands have Then, from Eqs. 8 & 9, the detection and discrimination of microplastics
different spatial resolutions, the same experiment was conducted twice. and other marine debris are discussed separately below.
The first used imaginary MSI bands where their resolutions were all set
to 10 m. The second used realistic resolutions for individual bands 4.1.1. Microplastics
(either 60, 10, or 20 m). In the latter case, if a 10-m band had χ = 20%, From 11,854 surface trawls between 1971 and 2013, van Sebille
the 20-m band had χ = 20% / 4 = 5%. Therefore, χ varied between bands et al. (2015) compiled and analyzed microplastics distributions in global
in the same mixed-pixel spectra, causing distorted spectral shapes (see oceans (microplastics defined by particle size <5 mm, see also Garaba
below). and Dierssen, 2018). The dominant majority showed surface density of
<1 M pieces km− 2, and nearly the entire data archive showed <10 M
4. Results pieces km− 2. So far, the highest reported density is 26 M pieces km− 2
(Fig. 5 of Matthews et al., 2017). In an independent study, Eriksen et al.
4.1. Sensitivity (2014) compiled a dataset of marine plastic debris measured at 1571
stations from 680 net tows and 891 visual survey transects. They also
Table 1 shows σ from the proposed NASA mission (HyspIRI, showed that the dominant majority of all compiled particle density and
currently Surface Biology and Geology or SBG) assuming an SNR of 200, modeled particle density is <1 M pieces km− 2 for particles <4.75 mm in
and σ estimated from MSI measurements over clear-water scenes (Qi and size. Therefore, it is reasonable to assume that the maximum density of
Hu, 2021) following the approach of Hu et al. (2012). Here, σ represents microplastics in natural waters is ~10 M pieces km− 2. In another study
noise estimated from R instead of total at-sensor radiance. Only several by Cózar et al. (2014), the size of plastics from a compiled dataset
MSI bands in the green, red, and NIR wavelengths are listed because showed a log-normal distribution with most particles of <5 mm and the
these are the most relevant bands to detect floating matters. Rt,typical is histogram mode of ~2 mm.
the typical total reflectance over oceans under cloud-free and glint-free Assuming a mean size of 2.5 mm per piece, the maximum density of
conditions (Hu et al., 2012). MSI SNRs are lower than the proposed 10 M pieces km− 2 is equivalent to about 50 m2 microplastics km− 2 (i.e.,
HyspIRI SNRs, and the corresponding σ is 2–4 times higher than the χ = 0.005% of a pixel) if all pieces are laid on the very surface without
proposed HyspIRI σ. For simplicity, σMSI in the NIR is assumed to be blocking each other. Clearly, χ = 0.005% is << χH det (0.2%) and also <<
mean + 2 standard deviations (6 × 10− 4); Likewise, σH in the NIR is χMSI − 5
det (0.8%, Eq. 8). For χ = 0.005%, ΔR = 0.25 × χ ≈ 1 × 10 . Such a
assumed to be 2 × 10− 4. signal, corresponding to the maximum microplastics density reported in
Then, for a HyspIRI-like sensor, Eq. 1 suggests that in order for a pixel the literature, is 60 times lower than the required 6 × 10− 4 and 20 times
to “stand out” from the nearby background water pixels, ΔR needs to be lower than sensor noise for a sensor with SNRs of 200. In turn, in order
> ~ 6 × 10− 4. For MSI-like sensors with σ ~ 6 × 10− 4, ΔR needs to be > for microplastics particles to be detected, their density needs to be at
~ 2 × 10− 3. Assuming RNIR NIR
FM ≈ 0.25 (Figs. 1b & 2a) and RW ≈ 0 (Fig. 2c), least >600 M pieces km− 2 or 600 pieces m− 2. Even though, all these
from Eq. 3, the following lower detection limits can be derived for a particles need to be aggregated on the very surface without blocking
HyspIRI-like sensor and for MSI, respectively: each other in order to achieve a maximum reflectance signal.
√̅̅̅ / In the marine environment, because microplastics particles may
χdet H ≥ 2 2σH (RFM –RW ) ≈ 0.2%
actually be below water surface due to mixing or other processes
√̅̅̅ / (Kukulka et al., 2012; Kooi et al., 2016; Martinez-Vicente et al., 2019),
χdet MSI ≥ 2 2σMSI (RFM –RW ) ≈ 0.8% (8) their NIR reflectance can be much lower than when they are all aggre­
gated at the ocean surface. Therefore, the requirement is even higher
Similarly, assuming RNIR red NIR red
FA – RFA ≈ 0.25 (Fig. 2a) and RMD – RMD ≈ 0.0 than shown above, reinforcing the argument that remote detection of
(Fig. 1b), from Eq. 6, the following lower discrimination limits can be microplactics is nearly impossible.
derived for a HyspIRI-like sensor and for MSI, respectively: The requirement of >600 microplastics pieces m− 2 or χ > 0.2% for a
sensor with SNRs of 200 is way beyond the reported maximum density of
10 pieces m− 2. For the same SNRs, in order to detect a single micro­
Table 1 plastics piece with an average size of 2.5 mm, the pixel size must be 25
Reflectance noise (σ) used in the sensitivity analysis. Rt,typical is typical top-of- cm2 (i.e., 5-cm resolution) or smaller. Even if such a resolution is
atmosphere reflectance over the ocean (Hu et al., 2012). “H” is for HyspIRI possible, its SNRs are usually much lower than 200. Therefore, for all
specification. σMSI is from Qi and Hu (2021) estimated from clear-water scenes. existing and future non-military satellite sensors, such a requirement is
For simplicity, σMSI in the NIR is assumed to be mean + 2 standard deviations,
simply impossible to meet. Such a conclusion would not be altered even
about 6 × 10− 4. Likewise, σH in the NIR is assumed to be 2 × 10− 4.
if the microplastic particle size were assumed to be the maximum of its
Band (nm) Rt,typical SNRH σH σMSI category (5 mm) instead of the assumed average size of 2.5 mm. In this
(×10− 2) (×10− 4) (×10− 4)
case, the enhanced reflectance due to microplastics is 4 times higher, but
560 6.90 200 3.45 7.14 ± 2.70 still much lower than sensor noise.
665 3.71 200 1.86 6.58 ± 3.92 The take home message from the sensitivity analysis is simple: the
741 2.73 200 1.37 4.15 ± 1.00
865 1.85 200 0.93 3.70 ± 0.85
enhanced signal due to microplastics at the reported maximum density

5
C. Hu Remote Sensing of Environment 259 (2021) 112414

even with all particles on the very surface without blocking each other 4.2. Simulation experiment: Spectral shape variations
(in reality, this is unlikely to happen due to mixing or other processes in
the ocean) is 60 times lower than the required signal and 20 times lower While Section 4.1 is focused on the sensor’s sensitivity or SNRs to
than sensor noise (assuming SNRs = 200). To make them detectable, the detect and discriminate marine debris, this section illustrates how
particle density needs to be at least 60 times higher than the reported spectral shape changes with spectral endmember, water type (clear or
maximum, or the pixel size must be 25 cm2 (5-cm resolution) or smaller. turbid), and χ. Fig. 3 illustrates how R and ΔR from a mixed Sargassum-
In the end, it would be extremely difficult, if not completely impossible, water pixel change with χ for both clear and turbid waters. The same for
to remotely sense microplastics from space using vis-NIR wavelengths. the “plastic bags” endmember is presented in Fig. 4. In these simulations,
Note that the above sensitivity analysis is only on whether a spatial Sargassum endmember spectrum is from Fig. 2a, plastic endmember
anomaly can be detected (i.e., is there “something”?) because if the spectrum is from Fig. 1b (“plastic bags”), and clear/turbid water end­
presence/absence detection is impossible, there is no need to discuss members spectra are from Fig. 2c.
discrimination. In both figures, the left column represents the results with imaginary
These findings are certainly disappointing. However, a critical MSI bands all having 10-m resolution, while the right column is for the
assumption in the above calculations is that the microplastics pieces are results with realistic MSI bands where the 704, 741, 783, and 865-nm
uniformly distributed on the surface. When the assumption is violated, bands have 20-m resolution but other bands (except 443 nm) have 10-
for example when the particles are heavily concentrated along narrow m resolution. In the latter case, the subpixel coverage of floating mat­
ocean fronts, windrows, or in small-scale eddy convergence zones so ters in the 20-m bands is only ¼ of the 10-m bands, creating spectral
that particle density is >0.2% (i.e., > 600 m− 2), it might be possible to distortion, for example an artificial local maximum at 841 nm and an
detect the reflectance anomalies and therefore presence/absence of artificial local minimum at 704 nm. Worse than this, the degree of the
microplastics particles. spectral distortion varies with water type and floating matter type as
The estimates above are based on the assumed SNR of ~200, as well as χ. Thus, it would be difficult to use spectra of mixed pixels to
proposed for the HyspIRI mission (currently SBG). High spatial- represent spectral endmembers, especially when χ is small (e.g., ~20%
resolution sensors typically have SNRs much lower than 200, resulting in Topouzelis et al. (2019 & 2020), Kikaki et al. (2020), and Biermann
in much higher sensor noise (Table 1). In such realistic cases, the et al. (2020)). It is believed that in nearly all published papers on the use
detection limit is also higher, for example χdet > 0.8% (or particle of MSI to detect marine debris, the resolution mismatch among different
density > 2400 m− 2) for MSI. bands was not considered enough, and the spectral endmembers in the
Finally, the above arguments are purely from the perspective of in­ published papers appear to be distorted in a similar fashion as shown
strument sensitivity. In some special cases, microplastics may aggregate above, thus may result in large uncertainties in their interpretations.
among other larger floating matters, for example Sargassum. Because we The second observation from Figs. 3 & 4 is that, consistent with the
can currently estimate Sargassum density and distributions using both argument in Section 3.1, the use of ΔR is superior to R when evaluating
coarse- and medium-resolution satellite sensors (Wang and Hu, 2016, & spectral shapes of mixed pixels. Indeed, in the log-scale plots, the
2020; Wang et al., 2018), if a relationship between microplastics and spectral shape in ΔR does not change with χ regardless of whether
Sargassum density can be established from field surveys, the relationship imaginary or realistic MSI bands are used as long as ΔR is positive.
may be applied to the synoptic observations of Sargassum in the Atlantic The third observation is that, when the blue bands of 443-nm and
to estimate the total amount of microplastics around these large mac­ 492-nm are excluded, the spectral shape in ΔR of mixed pixels is also
roalgae mats. stable between clear and turbid waters (i.e., the empty and solid symbols
almost overlap with each other). This has significant implications to
4.1.2. Macroplastics and other debris applications of satellite imagery in different water environments, as
Although made of different materials, both macroplastics (> 5 mm) water type (i.e., clear or turbid) may be excluded from consideration
and other non-plastic debris have broad-band spectral response (e.g., when performing spectral analysis.
Fig. 1b) similar to microplastics, and thus may be difficult to differen­
tiate from each other spectrally. For this reason, they are treated as the 4.3. Simulation experiment: spectral similarity
same type, termed as macro debris.
Similar to the detection of microplastics, Step 1 in macro debris While Figs. 3 & 4 show the spectral shapes after spectral mixing,
detection is also to detect a spatial anomaly, where the requirement on Figs. 5 & 6 show the spectral similarity measures between the mixed
subpixel coverage is the same: χdet > 0.2% for a SNR of 200, assuming pixel and floating matter endmember, expressed in SAM (Eq. 7). Lower
macro debris is on the very surface as opposed to be submersed in water. SAM values indicate higher similarity. To understand whether SAM is
For a 10-m pixel, this means that the macro debris patch within the pixel different between the two endmembers (Sargassum and plastic), each
must be at least >0.2 m2. For MSI, the detection limit is χdet > 0.8% or mixed pixel is compared to both endmembers.
0.8 m2. Once a spatial anomaly is detected, spectral analysis is still From these results, the following can be summarized.
required in Step 2 to tell whether the anomaly is due to macro debris or One, consistent with the findings from Figs. 3 & 4, ΔR (solid bars) is
other floating matters (i.e., floating algae). From Eq. 9, χdis needs to be way more efficient than R (empty bars) in differentiating floating mat­
>0.3% for a HyspIRI-like sensor, and > 1% for MSI. For a 10-m pixel, ters, either Sargassum or plastic (i.e., compare the empty and solid bars
this means 0.3 m2 and 1 m2, respectively. in the left column of both Figs. 5 & 6).
Although still difficult, these detection and discrimination limits can Two, such an ability is compromised for MSI spectra with mixed
certainly be met under certain circumstances, for example around river band resolutions because their spectral shapes are distorted to the var­
mouths, in frontal convergence zones, or along windrows of the ocean, iable χ in different bands (right column in both Figs. 5 & 6). Indeed, with
as shown in various digital photos of Biermann et al. (2020) and Kikaki mixed band resolutions, it is difficult to determine whether the
et al. (2020). However, in practice, as shown below, while detection and Sargassum-containing pixel is spectrally more similar to Sargassum
discrimination of floating algae and non-algae features are possible, endmember or to plastic endmember (Fig. 5 right column), or whether
discrimination of macro debris type is actually more demanding than the plastic-containing pixel is spectrally more similar to plastic end­
shown above due to spectral similarity among different types of macro member or to Sargassum endmember (Fig. 6 right column).
debris. Three, in contrast, if all MSI bands are forced to have the same 10-m
resolution, Sargassum-containing pixels and plastic-containing pixels
can be easily separated through comparing their SAM values to both
endmembers, and such a separation is possible down to at least 1%

6
C. Hu Remote Sensing of Environment 259 (2021) 112414

Fig. 3. Simulated mixing experiment for Sargassum


using Sentinel-2 MSI band settings. The blue and
green dashed lines represent clear water and turbid
water, respectively. The legend shows χ from 1% to
20% for the 10-m bands. Left column represents re­
sults with imaginary MSI bands where all bands have
the same 10-m resolution. Right column represents
results with realistic MSI bands where the following
bands have 20-m resolution: 704, 741, 783, and 865
nm, and the 443-nm band has 60-m resolution; (a) R
for mixed pixels; (b) ΔR between mixed pixels and
water endmembers; (c) same as in (b) but ΔR is
plotted in log scale. Note the artificial peak at 841 nm
in the right column due to spectral distortion caused
by mixed band resolutions. (For interpretation of the
references to colour in this figure legend, the reader is
referred to the web version of this article.)

subpixel coverage. This is shown by the solid bars between the two facilitate image visualization, but the detection limit may be compro­
colors in the left column of both Figs. 5 and 6. mised to a higher value, for example to ~1% for floating algae with a
Although the simulation experiment used only two endmembers of SNR of 200 (Hu et al., 2015). Likewise, in practice, a single pixel above
Sargassum and plastic, because their spectral shapes can represent the detection limit is difficult to interpret (e.g., Gower et al., 2014, also
floating algae and macro debris, respectively, the findings above can see Fig. 7). Rather, a spatially coherent feature (e.g., an image slick)
help guide spectral analysis and algorithm development when applying from at least several pixels is required, thus requiring higher χ than the
MSI imagery to detect marine debris and other floating matters, as sensitivity-based estimates of 0.2% and 0.3% (for SNRs of 200) and 0.8%
shown below. Indeed, most floating algae have similar red-edge reflec­ and 1% (for MSI). For example, χdet and χdis for MSI may be 2–3 times
tance and similar 670-nm reflectance trough as in the Sargassum end­ higher than the sensitivity-based estimates, being 2% and 3%,
member (Fig. 2), and nearly all macro plastics (except those monotonic respectively.
green or blue plastic bottles, Garaba and Dierssen, 2020) have similar Furthermore, different sensors have different artifacts, which should
flat spectral shapes as those shown in Fig. 1b. Considering that hyper­ be considered carefully when analyzing the spatial and spectral anom­
spectral and high-resolution sensors are currently unavailable, the most alies. One example is the hardware parallax effect in push-broom sensors
achievable appears to be separating floating algae from non-algae while such as MSI (ESA, 2015) and Landsat-8 OLI, where different bands are
further differentiating the non-algae type (e.g., plastics or non-plastic not co-registered in time for a given pixel. Such an effect can create
debris or other objects) is rather challenging, as shown below. colorful pepper noise in RGB composite imagery over moving targets (e.
g., waves, Fig. 7) when sun glint or sky glint is apparent (Liu et al.,
4.4. Practical considerations 2020). Other disturbing image features include pepper noise due to
cosmic rays (Gower et al., 2014) and waves that can also confuse al­
The above sensitivity analysis (Eqs. 8 and 9) is based on the ideal gorithms to detect small features (Wang and Hu, 2020). Under these
situations where image noise is assumed to come from the sensor noise circumstances, detecting marine debris becomes more difficult because
only, and spectral shapes in either the marine debris endmembers or the although a noise removal algorithm may be used to remove these arti­
floating algae endmembers are stable. Therefore, both χdet and χdis facts (Wang and Hu, 2020), the same algorithm may also remove the
represent the lower bounds, i.e., below which detection and discrimi­ small features caused by marine debris as they are expected to be at
nation are impossible, but above which whether or not they are possible subpixel scales.
still depend on other factors. Similar to the sensitivity analysis, the simulation results in Sections
For example, the detection limit is estimated from a single band in 4.2 and 4.3 are also simplified to demonstrate the concept of 1) why ΔR
the NIR because this is where the maximum contrast occurs between is preferred over R to differentiate floating matter type and 2) for MSI,
floating matters and water, but single-band images are difficult to why the use of single pixels is not a practical way for spectral discrim­
interpret as reflectance of the water background may change substan­ ination even if ΔR is used. In real applications, other types of floating
tially across the image (Wang and Hu, 2016). The use of floating algae matters as well as spectral modulations by image noise also need to be
index (FAI, Hu, 2009) or other indexes (e.g., FRGB, Qi et al., 2020) may considered. However, even such conceptual demonstrations may

7
C. Hu Remote Sensing of Environment 259 (2021) 112414

Fig. 4. Same as Fig. 3, but the endmember is “plastic bags” (Fig. 1b) instead of Sargassum. Note the artificial peak at 841 nm in the right column due to spectral
distortion caused by the mixed band resolutions. Such a peak does not exist if all bands have the same resolution (left column).

provide some guidance on how to perform the spectral analysis. mixture of marine debris and foam, or marine debris alone, especially if
For example, to avoid MSI spectral distortion in single pixels, mean the debris is partially submersed in water to cause a reflectance decrease
spectra from 5 × 5 pixels instead of a single pixel may be used to derive in the NIR wavelengths.
the spectral shape in a more reliable way, as shown in Fig. 8. Although The results in Fig. 8 are for demonstration purpose only, as in reality
the magnitude of the mean spectra is lower than from some of the 5 × 5 the plastic endmember may have slightly different spectra than used
pixels, the spectral shape of the floating matter is largely retained, thus here. If the plastic endmember spectra were replaced by the microplastic
facilitating analysis for spectral similarity. The two MSI images in Fig. 8 spectra collected by Garaba and Dierssen (2020) (red curve in Fig. 1b),
were collected from the southern Caribbean Sea where marine debris the SAM to plastic would be 9.8o (Fig. 8a) and 5.3o (Fig. 8c), respec­
has been reported (Kikaki et al., 2020). However, from the spectral tively. This new result actually makes it easier to discriminate the two
shapes alone, it is difficult to conclude whether the pixels contain ma­ image slicks between Sargassum-like and plastic-like, but it does not alter
rine debris, especially for the first case in Fig. 8a. the principles presented above.
In Fig. 8a, the image slick appears greenish, indicating high NIR Note that the reflectance magnitude in both cases is very small.
reflectance because in the FRGB image, a NIR band is used to represent Assuming a NIR reflectance of 0.25 in both endmembers for χ = 100%,
the green channel. The mean ΔRrc spectrum in Fig. 8b from 5 × 5 pixels Fig. 8b suggests mean χ ~ 6% while Fig. 8d suggests mean χ ~ 2%. Even
(centered at the red “x” location) indeed shows red edge reflectance in though, the slick features can be detected and discriminated.
the NIR, strongly suggesting floating vegetation and very likely From the analysis demonstrated here, once a spatially coherent
Sargassum. Its SAM to the known Sargassum endmember is 3.8o, but 7.5o feature is detected in the MSI imagery, two techniques are recom­
to the plastic endmember. In contrast to the mean ΔRrc spectrum, ΔRrc mended to avoid the spectral “distortion” in spectral analysis or algo­
from 3 randomly selected pixels at the same location (Fig. 8b) show high rithm development. The first is to use ΔR instead of RT in Eq. 2. This
spectral noise, making them difficult to interpret. way, both floating matter and water contribute equal weights to ΔR
Similar observations are obtained in Fig. 8c where the image slick regardless of χ (Figs. 3c & 4c). The second is to average several pixels to
appears whitish, indicating lack of red-edge reflectance. The 5 × 5 mean minimize the impact of resolution mismatch. Without averaging, the
ΔRrc spectrum from the feature indicates non-algae materials (Fig. 8d). individual pixels in Figs. 8b & 8d show spectral shapes that are highly
Its SAM to plastic is half of that to Sargassum (8.2o versus 18.5o), but its noisy, making them impossible to compare with known floating matter
NIR reflectance is lower than the green-red reflectance. This charac­ endmember spectra.
teristic actually does not appear to resemble plastic, but more similar to
foam (whitecap) spectra (Dierssen, 2019), especially when considering
that the local minima at 741 nm are near the reported 756-nm minimum 4.5. Case study over the West Florida shelf
in whitecap reflectance. However, it is also possible that the feature is a
With the findings above, a case study using MSI data is presented

8
C. Hu Remote Sensing of Environment 259 (2021) 112414

Fig. 5. Spectral similarity between a mixed Sargassum pixel and Sargassum endmember (orange bars), and between a mixed Sargassum pixel and plastics endmember
(blue bars) using the following MSI bands: 560, 665, 704, 741, 841, 865 nm. (a) and (b): all MSI bands are 10 m; (c) and (d): the following bands are 20 m: 705, 741,
and 865. The solid orange bars (annotated with arrows) in (a) and (b) are close to 0 degrees, indicating high similarity. These results correspond to the spectra shown
in Fig. 3. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 6. Same as Fig. 5, but the mixed pixel is between water and plastic. The solid blue bars in (a) and (b) (annotated with arrows) are close to 0 degrees, indicating
high similarity. These results correspond to the spectra shown in Fig. 4. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)

9
C. Hu Remote Sensing of Environment 259 (2021) 112414

as indicated by the 665-nm absorption feature (arrows in Figs. 9j & 9k).


Digital photos taken on the same day in Area 4 confirms that these slicks
are floating Trichodesmium mats (Fig. 9b). Indeed, even without such a
field validation, from knowledge of regional oceanography (Hu et al.,
2010) and from the spectral shapes in Figs. 9j & 9k, one can still
conclude that these slicks are very likely due to surface aggregation of
Trichodesmium.
In contrast, although it is clear that the slicks in Areas 1 & 2 are not
floating algae, it is very difficult to discriminate their type based on the
spectral shapes. The 443-band may be excluded because the spectral
distortion by this 60-m band cannot be removed even after 5 × 5 pixel
averaging. Then, except for the residual band-resolution effect in several
NIR bands (704-nm, 741-nm, 783-nm, 865-nm, all having 20-m reso­
lution), all spectra are featureless (Figs. 9h & 9i). The question becomes
Fig. 7. MSI FRGB image near Japan (centered around 31.7593oN 142.2510◦ E) what could cause these image slicks, marine debris or other floating
showing colorful pixels due to the sensor parallax effect. The spectral shapes of matters. Unfortunately, there is no easy answer.
these pixels in the inset figure appear abnormal as they do not resemble any First, foams or white caps may be ruled out because these slicks show
known spectral shapes, including those from marine debris. red-rich spectra (in contrast, whitecaps are blue-rich, Dierssen, 2019)
and because the ocean was very calm (wind is mostly <3 m s− 1 for two
here to demonstrate how to detect image features (spatial anomaly) and
consecutive days before the imaging time, Fig. 9c). Weather data
how to discriminate the feature type.
showed almost no rainfall prior to the imaging time, and there is no
For simplicity, the first step in remote detection of marine debris, i.e.,
major river nearby, suggesting these materials are unlikely to be riverine
detecting a spatial anomaly, is through visual inspection while sophis­
origin. Furthermore, there was no news report of large debris patches in
ticated image segmentation may be implemented in the future (e.g.,
this region, and routine cruise surveys to this region in the past (almost
Wang and Hu, 2020). Fig. 9a shows an MSI RGB sub-image west and
monthly, as part of the red tide monitoring effort) never encountered
northwest of Tampa Bay (Florida, USA), where the full-resolution image
large debris patches in this region. Therefore, although from the
in several areas (denoted as “1” – “4”) reveals image slicks (Figs. 9d – 9
perspective of image spectroscopy these slicks appear like marine debris,
g). These slicks indicate the presence of floating matters.
it is difficult to conclude without a direct field validation, let alone
Spectral analysis of representative pixels from the slicks, through the
further differentiating the debris type (e.g., plastic or non-plastic). At the
use of ΔRrc as in Fig. 8, indicate different spectral shapes from these
present, whether the slicks in Areas 1 and 2 are due to marine debris or
slicks (Figs. 9h – 9 k). While slicks in Areas 1 & 2 show flat spectral
other type of floating matters remains to be a puzzle. On the other hand,
shapes, those in Areas 3 and 4 show typical shapes from floating algae,
detecting and discriminating floating matters (algae versus non-algae)

Fig. 8. (a) MSI FRGB image on 29 November 2015 in the SW Caribbean Sea. The red “x” (16.0649oN, 86.4025oW) and black “x” symbols mark the locations where
spectra were extracted for analysis in (b). The Rrc spectra of 5 × 5 pixels from the two locations are shown to the right y-axis, while their ΔRrc spectra are shown to the
left y-axis. In addition to the 5 × 5 mean spectrum (solid green circle), spectra from 3 individual pixels are also shown. (c) MSI FRGB image on 30 August 2018 in the
SW Caribbean Sea. The red “x” (15.7479oN, 88.1656oW) and black “x” symbols mark the locations where spectra were extracted for analysis in (d), where the spectral
legend is the same as in (b). Note that the spectral shapes from individual pixels are all distorted (compared to the mean spectral shapes) due to mixed band res­
olutions. Annotated on the images are the SAM values to Sargassum and plastic endmembers, respectively, and mean χ of the marked pixels. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

10
C. Hu Remote Sensing of Environment 259 (2021) 112414

Fig. 9. An example showing the possibility of detecting and discriminating floating algae and non-algae matters as well as the challenge in identifying the type of
non-algae floating matters using MSI data. (a) MSI RGB image on 10 February 2021 over part of the west Florida shelf (NW off Tampa Bay); (b) Digital photo taken on
the same day in area “4” of the image showing surface Trichodesmium mats (photo credit: Charles Tilney and Matt Garrett of Florida Fish and Wildlife Conservation
Commission; photo location annotated as “X” in (g)); (c) Windspeed measured from a nearby marine buoy, where the marked dates represent the beginning of the day
(0:00 GMT); (d) – (g) Enlarged sub-images corresponding to areas 1–4 in (a), where surface slicks are visible; (h) – (k) ΔRrc spectra (5 × 5 pixels, vertical bars indicate
standard deviations) extracted from MSI pixels of the image slicks in (d) – (g), respectively, where the pixel locations are marked in red. The pigment absorption
features in (j) and (k) are marked with arrows. In (h) – (k), mean χ of each 5 × 5 pixels is annotated. (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

are shown to be possible for χ ≥ 3% from a group of pixels, a result Therefore, band-combination indexes, such as FAI (Hu, 2009), may be
consistent with the sensitivity analysis and practical considerations used to mitigate such effects at the price of reduced sensitivity to detect
above. spatial anomalies. This is exactly the reason why the same 200 SNR led
to the ~1% detectability estimate in Hu et al. (2015) but ~0.2% in this
5. Discussions: Looking forward study. The FAI design can be changed to use different band combinations
depending on band availability and application needs, for example
5.1. Which wavelengths (bands) to use through the alternative FAI (AFAI, Wang and Hu, 2016; Qi et al., 2017)
or the floating debris index (FDI, Biermann et al., 2020). In the end, a
While some current and future satellite sensors are equipped with combination of a red band (665 nm), a NIR band (754 nm), and another
spectral bands in the SWIR wavelengths and SWIR bands have proven NIR band (865 nm) or a SWIR band (1.2 μm or 1.6 μm) should be suf­
useful in detecting and quantifying emulsified oil (Clark et al., 2010; Lu ficient in detecting the presence of floating matters.
et al., 2020; Hu et al., 2021), this study is focused on the vis-NIR bands for Then, the inclusion of a green band (560 nm) with the 665-nm band
several reasons. First, the spectral contrast between floating matters and and NIR bands makes it easy to calculate SAM in order to discriminate
background waters is mostly in the NIR, regardless of the type of floating between floating algae and non-algae floating matters (Figs. 5 & 6). The
matters (either Sargassum, Ulva, Trichodesmium, or plastics, see references chlorophyll absorption feature around 665 nm (arrows in Figs. 2a & 2b,
in Figs. 1 & 2). Reflectance of floating matters in the SWIR bands is several Figs. 9j & 9k) represents a key difference between most floating algae
times lower than in the NIR bands, making it more difficult to detect if and non-algae floating matters, therefore can be used in either the SAM
SWIR bands are to be used. This is different from emulsified oil that may or other indexes to differentiate the two types. Once an image feature is
have higher SWIR reflectance than in the NIR. Second, due to mixing and being identified as floating algae, spectral shapes in the visible wave­
other processes, floating matters may be slightly below surface, making the lengths can be used to further differentiate whether the floating algae is
SWIR reflectance quickly disappear due to strong water absorption. For Sargassum, Ulva, Trichodesmium, or others (Qi et al., 2020) as long as χ is
example, water absorption coefficient is ~130 m− 1 at 1.2 μm and ~ 700 above the discrimination threshold (Qi et al., 2019).
m− 1 at 1.6 μm. If the floating matter is 1 cm below surface, its SWIR
reflectance would be reduced by 14 times at 1.2 μm and 1 million times at
1.6 μm. In contrast, the 765-nm reflectance would be reduced by only 6%. 5.2. Challenges in discriminating the type of non-algae floating matters
Furthermore, when implementing a detection scheme, although the
use of a single NIR may maximum the pixel-to-pixel contrast at local Once non-algae floating matters are identified using the above SAM-
scale, interpretating single-band images is usually difficult because of based or other similar approaches, further discriminating the type of
relatively large gradience (compared to noise) across the image. non-algae floating matters represents a technical challenge as most non-
algae debris appear to be similar in spectral shapes (Fig. 1b, Figs. 9h &

11
C. Hu Remote Sensing of Environment 259 (2021) 112414

9i). This is especially true considering the usually small χ values . violate atmospheric correction assumptions on negligible or predictable
Indeed, the pixels from the same image slick in Fig. 9h are likely to be of (predicted from the red band) NIR-SWIR surface reflectance (Stumpf
the same type, yet their pixel-to-pixel variations in their spectral shapes et al., 2003; Wang and Shi, 2007). Therefore, a nearest-neighbor at­
appear to encompass those from different plastic and non-plastic ma­ mospheric correction (Hu et al., 2000) or a dark-target based image-wise
terials (Fig. 1b, also see more examples in Garaba and Dierssen, 2020). atmospheric correction approach, such as that implemented in the
Such pixel-to-pixel variations also appear to overwhelm the spectral Acolite software (Vanhellemont and Ruddick, 2018), may be used. On
variations in the visible due to changes in the background waters. the other hand, because ΔR rather than R is used in Steps 2 and 3, both
Therefore, at the present, it appears that discrimination of non-algae Rrc and Rt can be used. This is because atmospheric properties over
floating matters from floating algae is doable but further pinpointing adjacent pixels are assumed to be the same, leading to ΔRrc = ΔRt = ΔR.
the non-algae floating matter type based on spectroscopy alone is not. Automation also requires cloud masking and other steps to mask
Adding more spectral bands, for example from hyperspectral measure­ pixels that are impossible to determine whether they contain floating
ments, might help, but this is a subject of future research. matters. These pixels are treated as no-observation pixels to avoid
One exception might be the separation of plastic versus non-plastic biasing statistics. This is perhaps more difficult than implementing the
marine debris, as the former shows specific, narrow, hydrocarbon ab­ above steps, as there is no universal way to mask these pixels. For
sorption features in the SWIR wavelengths (Garaba and Dierssen, 2020). example, the standard cloudmasking algorithm in SeaDAS uses a
Once hyperspectral data at those wavelengths are available, it might be threshold of Rrc(865 nm) > 0.027 to mask clouds, and the modified al­
possible to fingerprint floating plastics. This is similar to the use of these gorithm uses a threshold of Rrc(2130 nm) > 0.0175 to mask clouds
features in detecting and quantifying emulsified oil on the ocean surface (Wang and Shi, 2006). Because of the enhanced NIR and SWIR reflec­
(Clark et al., 2010; Lu et al., 2020; Hu et al., 2021). However, one tance due to floating matters, these cloud masking schemes may falsely
drawback is the small magnitude of these features from marine debris, mask some floating matter pixels as clouds. To overcome this difficulty,
which may be extremely difficult to detect for the reasons outlined customized cloudmasking algorithms have been used for different sen­
above. sors (Wang and Hu, 2016 & 2020), yet their global applicability needs to
Despite these difficulties, discriminating marine debris from other be evaluated. Likewise, cloud shadows in high-resolution imagery,
non-algae floating matters may still be possible through non- among other “artifacts”, also need to be identified and masked. These
spectroscopy methods. For example, most of the non-algae floating detailed image processing requirements are beyond the scope of this
matters are rare in the ocean, with often known locations, thus can be paper, and interested readers may refer to the relevant literature for
easily ruled out with some a priori knowledge. Inspection of wind data more information. In the end, a regional near real-time system may be
can also help rule out the possibility of whitecaps. developed to monitor both floating algae and non-algae floating matters,
On the other hand, although discrimination between floating algae similar to the Sargassum Watch System (SaWS, Hu et al., 2016) estab­
and floating debris (and further discrimination of the type of floating lished for the Atlantic Ocean and Gulf of Mexico.
debris) is desirable, it is not always necessary in an ecological perspec­
tive. This is because biofouling of marine debris is common, which has 5.4. “Optimal” optical sensors
implications on both the ecosystem and the fate of marine debris
(Maximenko et al., 2019). Nevertheless, the feasibility of discriminating In a recent paper, Martinez-Vicente et al. (2019) provided an initial
the type of non-algae floating matters needs to be explored more in a review of the current satellite sensor capability, including SAR and LIDAR,
practical sense with the existing satellite data. with the focus on sensors equipped with NIR and SWIR bands to emphasize
the contrast between marine debris and water in these bands. For passive
5.3. Automation considerations optical sensors, none of the existing satellite sensors was designed to
monitor floating matters, especially for the case of marine debris. While
Based on these observations, it appears possible to implement a step- designing a “perfect” optical sensor (including satellite orbital character­
wise approach to automate the detection and quantification of both istics) is beyond the scope of this paper, the observations above may pro­
floating algae and non-algae floating matters (including macro debris), vide some general guide if an “optimal” sensor/mission is to be designed.
for example: The characteristics of passive optical remote sensing can be gener­
Step 1 – detecting spatial anomaly and delineating image features. alized in the following 4 resolutions: spatial, spectral, radiometric, and
This can be based on a single band or a combination of bands using temporal resolutions, with the last two defined by SNRs and site revisit
image segmentation techniques (e.g., Wang and Hu, 2020); frequency. The importance of the first three resolutions has been dis­
Step 2 – discriminating between floating algae and non-algae floating cussed in this paper, and the last resolution depends on both sensor and
matters. This is based on their difference around 665 nm, which can be satellite orbital designs. Because there is always a trade-off between all
quantified through the use of SAM or other indexes. In this step, the four resolutions, for a coastal region, an optimal sensor/mission should
spectral shape around 665 nm is derived from reflectance difference be able to detect and discriminate non-algae floating matter patches of
(ΔRrc) in order to maximize the spectral contrast, where the background several m2 in size every few days. Such a capacity may enable the sensor
water pixels for the individual slick pixels can be found using a nearest- to search for missing fishing gears or large solid objects in the ocean due
neighbor approach (Hu et al., 2000). to Tsunami or other disasters. Thus, the following may serve as an
Step 3 – quantifying χ in each mixed pixel. This is through the use of optimal trade: 3–4 m spatial resolution, 4–6 spectral bands (443 nm,
locally tuned lower-bound threshold to represent χ = 0% and pre- 560 nm, 620 nm, 665 nm, 754 nm, 865 nm, with 665-nm band having
defined upper-bound threshold to represent χ = 100% (e.g., upper 10–20 nm bandwidth), SNRs of 50–100 at typical ocean radiance inputs,
bound for a single NIR band may be 0.25, but for AFAI may be 0.1 ac­ and revisit frequency of 3–4 days. The 865-nm band is used to form a
cording to the endmember spectra presented in Figs. 1b & 2a). baseline with the 670-nm band to calculate AFAI. If only 5 bands are
In all steps above, the fundamental property is the spectral reflec­ allowed, the 865-nm band may be removed because the 754-nm and
tance derived from satellite measurements. Ideally, atmospherically 670-nm bands can be used to calculate the normalized difference
corrected surface reflectance should be used to remove the variable ef­ vegetation index (NDVI). The 620-nm band is to differentiate floating
fects due to Rayleigh scattering, gaseous absorption, aerosol scattering, algae colors (green-rich or orange-rich) but can be sacrificed if only 4
sun glint, and solar/viewing geometry. However, this is not always bands are allowed. The other 4 bands (443, 560, 670, and 754 nm)
possible from a pixel-wise atmospheric correction approach (e.g., represent the core requirements for effective detection and discrimina­
currently being used in the NASA processing software SeaDAS and ESA tion of floating algae and non-algae floating matters.
processing software SNAP) because the presence of floating matter will Currently, the Sentinel-2 MSI sensors almost meet these

12
C. Hu Remote Sensing of Environment 259 (2021) 112414

requirements, yet the mixed band resolutions degrade their capacity in waves, and sensor artifacts (e.g., parallax effect). This is true even for
discriminating floating matter types. On the other hand, the Planet the same type of floating matter, not to mention the possibility that
Scope/DOVE constellation of hundreds of miniature satellite sensors more than one type of floating matter may exist in the same region.
(CubeSats) provides 3–4 m resolution data in 3–4 spectral bands with For the reasons mentioned above, in the case of lack of pure spectral
2–3 day revisit frequency in many coastal areas (http://www.planet. endmember (i.e., χ = 100%), it is better to use ΔR rather than R to
com), thus almost meeting the “optimal” requirements above. Unfortu­ represent the endmember.
nately, DOVE spectral bands are too wide (60–90 nm) to discriminate 6) Because the spectral similarity between macro plastics and non-
floating algae against non-algae floating matters. Nevertheless, these plastics debris (i.e., relatively flat, wide-band reflectance in the vis-
existing satellite sensors, combined with other high-resolution sensors NIR wavelengths), it appears difficult to separate them spectrally.
such as those discussed in Martinez-Vicente et al. (2019), may provide a This is also partially due to the lack of a relatively complete spectral
practical solution to address the challenges in sensor requirements, library of various marine debris. More measurements are therefore
while the challenges in algorithm development, as demonstrated above, required to complement those reported in previous studies focused
still need to be addressed. on plastics. On the other hand, because marine debris may be a
mixture of different types of materials, discriminating a specific type
6. Summary may be unnecessary.
7) The spectral analysis here using Sargassum and plastic is for
By no means is this paper meant to provide a comprehensive guide on demonstration purpose only. In the marine environment, other
how to remotely detect marine debris from space especially when known (e.g., Fig. 2) and unknown floating matters may exist, and
considering its focus on passive optical remote sensing. Instead, the paper some of them may have similar spectral shapes as marine debris (e.g.,
addresses two challenges identified by Martinez-Vicente et al. (2019) on Figs. 9d & 9e; the thought-to-be sea jelly schools in Qi et al., 2020). In
SNR requirements and methodology requirements, respectively, through the end, it may be impossible to discriminate marine debris from
answering the following questions posed earlier: to what extent can marine other unknown floating matters from a perspective of pure spec­
debris be remotely sensed, and how? troscopy. In such cases, other ancillary information (e.g., knowledge
From the above sensitivity analysis, simulation experiments, and of local oceanography, news reports of pollution events, etc.) may be
case studies using MSI imagery, the following observations and sug­ used to help interpret the detected image features.
gestions can be generalized: 8) Although the paper demonstrates under what conditions remote
detection of marine debris might be possible using the reflected vis-
1) Regardless of the floating matter type, either marine debris, oil NIR sunlight, it also shows conditions that make such a detection
slicks, or floating vegetation, from the perspective of spectroscopy, impossible through the sensitivity analysis. Clearly, there is still a lot
remote detection is always through spatial anomaly and remote to do after the recent pioneering studies on this topic, including
discrimination is always through spectral shape similarity as establishment of a more complete spectral library of various floating
opposed to spectral magnitude. Both depend on sensor’s SNRs and matters and development of more robust algorithms to detect debris-
band settings. like features in noisy remote sensing images.
2) From a theoretical basis, with only 4 bands around 560 nm, 665 nm,
and two NIR wavelengths, both χdet and χdis depend only on sensor Finally, the paper uses MSI data to demonstrate the concept and chal­
sensitivity (SNRs). They are 0.2% and 0.3% for a sensor with SNR of lenges in remote sensing of marine debris in the marine environment. The
200, and 0.8% and 1.0% for MSI. Below these limits, floating matters arguments may change substantially for remote sensing of marine debris
are simply not detectable. Considering other practical requirements when they may be heavily concentrated on beaches. In both environments,
(e.g., a spatially coherent image feature instead of a single pixel sensors with finer spatial resolution such as WorldView (2 m) or Planet­
needs to “stand out” from the background, these limits may be Scope/Dove (3–4 m) offer a better capacity in detecting smaller debris
increased by 2–3 times in order to detect and discriminate floating patches, although detecting such features in the marine environment re­
matters. For example, χdet and χdis for MSI may actually be 2% and quires more effort because these sensors were not designed for marine
3%, respectively (Figs. 8 & 9). applications. In any case, given the increased reports of marine debris and
3) The detection limits suggest that detecting microplastics is likely their increasingly recognized importance in the marine environment, more
impossible even for the reported maximum density and even when effort is required to develop reliable algorithms and feasible approaches to
all particles are at the very surface. Therefore, future efforts may be monitor and track marine debris, thus contributing to a final roadmap in
dedicated to frontal zones or windrows only, or to other surrogates remote sensing of marine debris (Maximenko et al., 2019).
(e.g., Sargassum) that may aggregate microplastics. Whether these
approaches are feasible, however, still requires more research. Credit author statement
4) While the detection of presence/absence of floating matters can be
through single-band images or band-combination indexes with each Chuanmin Hu: Conceptualization; Methodology; Data curation;
having its own strengths and weaknesses, spectral similarity (or Writing- Original draft preparation, revision, editing; Funding
anomaly) needs to be analyzed through the difference spectra (ΔR in acquisition.
Eq. 2) because this is the only way to retain the spectral shape of the
floating matter endmember when χ is typically small (e.g., <10%). Declaration of Competing Interest
Such a practice actually started from Gower et al. (2006) when
differentiating Sargassum from Trichodesmium in the Gulf of Mexico. The authors declare that they have no known competing financial
For the same reason, using spectral shapes derived from mixed pixels interests or personal relationships that could have appeared to influence
as endmembers are subject to large uncertainties because these the work reported in this paper.
shapes depend not only on the floating matter endmember, but also The authors declare the following financial interests/personal re­
on the unknown χ as well as on changes in the water endmember in lationships which may be considered as potential competing interests:
the real environment.
5) While detecting macro debris appears possible from MSI, more work Acknowledgements
is required to make the previously reported cases more convincing,
as spectral shapes of a mixed pixel can be modulated by the variable This work was supported by the U.S. NASA through its Ocean Biology
χ, variable water reflectance, mismatch in band spatial resolution, and Biogeochemistry program (NNX16AR74G, 80NSSC20M0264) and

13
C. Hu Remote Sensing of Environment 259 (2021) 112414

Ecological Forecast program (NNX17AF57G, 80NSSC21K0422). The Kikaki, A., Karantzalos, K., Power, C.A., Raitsos, D.E., 2020. Remotely sensing the source
and transport of marine plastic debris in bay islands of honduras (Caribbean Sea).
author thanks Drs. Eric Hochberg, Lianbo Hu, Lachlan McKinna, and
Remote Sens. 12, 1727. https://doi.org/10.3390/rs12111727.
Shungudzemwoyo Garaba for providing endmember spectra for Sargassum, Kooi, M., Reisser, J., Slat, B., et al., 2016. The effect of particle properties on the depth
Ulva, Trichodesmium, and microplastics, respectively. The author is grateful profile of buoyant plastics in the ocean. Sci. Rep. 6, 33882.
to Dr. Aikaterini Kikaki (National Technical University of Athens, Greece) Kruse, F.A., Lefkoff, A.B., Boardman, J.W., Heidebrecht, K.B., Shapiro, A.T., Barloon, P.
J., Goetz, A.F.H., 1993. The spectral image processing system (SIPS)—interactive
for the discussions on marine debris detection in the Caribbean Sea. Three visualization and analysis of imaging spectrometer data. Remote Sens. Environ. 44,
anonymous reviewers provided insightful comments to help improve the 145–163.
presentation of this work, whose efforts are appreciated. Kukulka, T., Proskurowski, G., Morét-Ferguson, S., Meyer, D.W., Law, K.L., 2012. The
effect of wind mixing on the vertical distribution of buoyant plastic debris. Geophys.
Res. Lett. 39 (L07601).
References Law, K.L., Moret-Ferguson, S., Maximenko, N.A., Proskurowsk, G., Peacock, E.E.,
Hafner, J., Reddy, C.M., 2010. Plastic accumulation in the North Atlantic subtropical
Arii, M., Koiwa, M., Aoki, Y., 2014. Applicability of Sar to marine debris surveillance gyre. Science 329, 1185–1188.
after the great East Japan earthquake. IEEE J. Sel. Top. Appl. Earth ObsRemote Sens. Liu, Y., Xu, B., Zhi, W., Hu, C., Dong, Y., Jin, S., Lu, Y., Chen, T., Xue, W., Liu, Y.,
7, 1729–1744. Zhao, B., Lu, W., 2020. Space eye on flying aircraft: from Sentinel-2 MSI parallax to
Biermann, L., Clewley, D., Martinez-Vicente, V., Topouzelis, K., 2020. Finding plastic hybrid computing. Remote Sens. Environ. 246, 111867. https://doi.org/10.1016/j.
patches in coastal waters using optical satellite data. Sci. Rep. 10, 1–10. rse.2020.111867.
Clark, R.N., Swayze, G.A., Leifer, I., Livo, K.E., Kokaly, R., Hoefen, T., Lundeen, S., Lu, Y., Shi, J., Hu, C., Zhang, M., Sun, S., Liu, Y., 2020. Optical interpretation of oil
Eastwood, M., Green, R.O., Pearson, N., 2010. A method for quantitative mapping of emulsions in the ocean – Part II: applications to multi-band coarse-resolution
thick oil spills using imaging spectroscopy. US Geological Survey Open-File Report imagery. Remote Sens. Environ. 242, 111778. https://doi.org/10.1016/j.
1167, 1–51. rse.2020.111778.
Cózar, Andrés, Fidel, Echevarría, Ignacio González-Gordillo, J., Irigoien, Xabier, Martinez-Vicente, V., Clark, J.R., Corradi, P., et al., 2019. Measuring marine plastic
Úbeda, Bárbara, Hernández-León, Santiago, Álvaro T. Palma, et al., 2014. Plastic debris from space: initial assessment of observation requirements. Remote Sens. 11,
debris in the Open Ocean. Proc. Natl. Acad. Sci. 111 (28), 10239–10244. 2443. https://doi.org/10.3390/rs11202443.
Dierssen, H., 2019. Hyperspectral measurements, parameterizations, and atmospheric Matthews, J.P., Ostrovsky, L., Yoshikawa, Y., Komori, S., Tamura, H., 2017. Dynamics
correction of whitecaps and foam from visible to shortwave infrared for ocean color and early post-tsunami evolution of floating marine debris near Fukushima Daiichi.
remote sensing. Front. Earth Sci. 7, 14. https://doi.org/10.3389/feart.2019.00014. Nat. Geosci. 10, 598–605.
Eriksen, M., Lebreton, L.C.M., Carson, H.S., Thiel, M., Moore, C.J., Borerro, J.C., Maximenko, N., Arvesen, J., Asner, G., Carlton, J., Castrence, M., Centurioni, L.,
Galgani, F., Ryan, P.G., Reisser, J., 2014. Plastic pollution in the world’s oceans: Chao, Y., Chapman, J., Chirayath, V., Corradi, P., et al., 2017. Remote sensing of
more than 5 trillion plastic pieces weighing over 250,000 tons afloat at sea. PLoS marine debris to study dynamics, balances and trends. Available online. https://ec
One 2014 (9), e111913. ocast.arc.nasa.gov/las/Reports%20and%20Papers/Marine-Debris-Workshop-2017.
ESA (2015). Sentinel-2 User Handbook. https://sentinel.esa.int/documents/247904/ pdf.
685211/Sentinel-2_User_Handbook Accessed date: 24 October 2020. Maximenko, N., Corradi, P., Law, K.L., Van Sebille, E., Garaba, S.P., Lampitt, R.S.,
Garaba, S.P., Dierssen, H.M., 2018. An airborne remote sensing case study of synthetic Galgani, F., Martinez-Vicente, V., Goddijn-Murphy, L., Veiga, J.M., et al., 2019.
hydrocarbon detection using short-wave infrared absorption features identified from Toward the integrated marine debris observing system. Front. Mar. Sci. 6, 447.
marine-harvested macro- and microplastics. Remote Sens. Environ. 205, 224–235. https://doi.org/10.3389/fmars.2019.00447.
Garaba, S.P., Dierssen, H.M., 2020. Hyperspectral ultraviolet to shortwave infrared McKinna, L.I.W., 2010. Optical Detection and Quantification of Trichodesmium spp.
characteristics of marine-harvested, washed-ashore and virgin plastics. Earth Syst. within the Great Barrier Reef. PhD dissertation. James Cook University, Australia,
Sci. Data 12 (77–86), 2020. https://doi.org/10.5194/essd-12-77-2020. 312pp.
Garaba, S.P., Aitken, J., Slat, B., Dierssen, H.M., Lebreton, L., Zielinski, O., Reisser, J., Qi, L., Hu, C., 2021. To What Extent Can Ulva and Sargassum Be Detected and Separated
2018. Sensing ocean plastics with an airborne hyperspectral shortwave infrared in Satellite Imagery? Harmful Algae doi:10.1016/j.hal.2021.102001.
imager. Environmental Sci. & Techol. 52, 11699–11707. Qi, L., Hu, C., Wang, M., Shang, S., Wilson, C., 2017. Floating algae blooms in the East
Garaba, S.P., Acuña-Ruz, T., Mattar, C.B., 2020. Hyperspectral longwave infrared China Sea. Geophys. Res. Lett. 44 https://doi.org/10.1002/2017GL075525.
reflectance spectra of naturally dried algae, anthropogenic plastics, sands and shells. Qi, L., Tsai, S.F., Chen, Y., Le, C., Hu, C., 2019. In search of red Noctiluca scintillans
Earth Syst. Sci. Data 12, 2665–2678. blooms in the East China Sea (2019). Geophys. Res. Lett. 46, 5997–6004. https://
Goddijn-Murphy, L., Dufaur, J., 2018. Proof of concept for a model of light reflectance of doi.org/10.1029/2019GL082667.
plastics floating on natural waters. Mar. Pollut. Bull. 135, 1145–1157. https://doi. Qi, L., Hu, C., MIkelsons, K., Wang, M., Lance, V., Sun, S., Barnes, B.B., Zhao, J., der
org/10.1016/j.marpolbul.2018.08.044. Zande, D.V., 2020. In search of floating algae and other organisms in global oceans
Goddijn-Murphy, L., Williamson, B., 2019. On thermal infrared remote sensing of plastic and lakes. Remote Sens. Environ. 239, 111659. https://doi.org/10.1016/j.
pollution in natural waters. Remote Sens. 11, 2159. rse.2020.111659.
Goddijn-Murphy, L., Peters, S., van Sebille, E., James, N.A., Gibb, S., 2018. Concept for a Richardson, A.J., Everitt, J.H., 1992. Using spectral vegetation indices to estimate
hyperspectral remote sensing algorithm for floating marine macroplastics. Mar. rangeland productivity. Geocarto International 7 (1), 63–69.
Pollut. Bull. 126, 255–262. Stumpf, R.P., Arnone, R.A., Gould, R.W., Martinolich, P.M., Ransibrahmanakul, V., 2003.
Gower, J., Hu, C., Borstad, G., King, S., 2006. Ocean color satellites show extensive lines A partially coupled ocean-atmosphere model for retrieval of water-leaving radiance
of floating Sargassum in the Gulf of Mexico. IEEE Trans. Geosci. Remote Sens. 44, from SeaWiFS in coastal waters. In: Hooker, S.B., Firestone, E.R. (Eds.), SeaWiFS Post
3619–3625. Launch Technical Report Series, NASA Tech. Memo. 2003-206892, Vol. 22. NASA
Gower, J., King, S., Young, E., 2014. Global remote sensing of Trichodesmium. Int. J. God- dard Space Flight Center, Greenbelt, Maryland, pp. 51–59.
Remote Sens. 35 (14), 5459–5466. https://doi.org/10.1080/ Topouzelis, K., Papakonstantinou, A., Garaba, S.P., 2019. Detection of floating plastics
01431161.2014.926422. from satellite and unmanned aerial systems (plastic litter project 2018). Int. J. Appl.
Hu, C., 2009. A novel ocean color index to detect floating algae in the global oceans. Earth Obs. Geoinf. 79, 175–183.
Remote Sens. Environ. 113, 2118–2129. Topouzelis, K., Papageorgiou, D., Karagaitanakis, A., Papakonstantinou, A.,
Hu, C., Carder, K.L., Muller-Karger, F.E., 2000. Atmospheric correction of SeaWiFS Ballesteros, M.A., 2020. Remote sensing of sea surface artificial floating plastic
imagery over turbid coastal waters: a practical method. Remote Sens. Environ. 74, targets with Sentinel-2 and unmanned aerial systems (plastic litter project 2019).
195–206. Remote Sens. 12, 2013. https://doi.org/10.3390/rs12122013.
Hu, C., Cannizzaro, J., Carder, K.L., Muller-Karger, F.E., Hardy, R., 2010. Remote van Sebille, E., Wilcox, C., Lebreton, L., Maximenko, N.A., Hardesty, B.D., van
detection of Trichodesmium blooms in optically complex coastal waters: examples Franeker, J.A., Eriksen, M., Siegel, D., Galgani, F., Law, K.L., 2015. A global
with MODIS full-spectral data. Remote Sens. Environ. 114, 2048–2058. inventory of small floating plastic debris. Environ. Res. Lett. 10 (12), 124006.
Hu, C., Feng, L., Lee, Z., Davis, C.O., Mannino, A., McClain, C.R., Franz, B.A., 2012. https://doi.org/10.1088/1748-9326/10/12/124006.
Dynamic range and sensitivity requirements of satellite ocean color sensors: learning Vanhellemont, Q., Ruddick, K., 2018. Atmospheric correction of metre-scale optical
from the past. Appl. Opt. 51, 6045–6062. satellite data for inland and coastal water applications. Remote Sens. Environ. 216,
Hu, C., Feng, L., Hardy, R.F., Hochberg, E.J., 2015. Spectral and spatial requirements of 586–597.
remote measurements of pelagic Sargassum macro algae. Remote Sens. Environ. 167, Wang, M., Hu, C., 2016. Mapping and quantifying Sargassum distribution and coverage
229–246. https://doi.org/10.1016/j.rse.2015.05.022. in the central West Atlantic using MODIS observations. Remote Sens. Environ. 183,
Hu, C., Murch, B., Barnes, B.B., Wang, M., Marechal, J.P., Franks, J., Johnson, D., 356–367. https://doi.org/10.1016/j.rse.2016.04.019.
Lapointe, B., Goodwin, D.S., Schell, J.M., Siuda, A.N.S., 2016. Sargassum watch Wang, M., Hu, C., 2020. Automatic extraction of Sargassum features on Sentinel-2 MSI
warns of incoming seaweed. Eos 97 (22), 10–15. https://doi.org/10.1029/ images. IEEE Trans. Geosci. Remote Sens. https://doi.org/10.1109/
2016EO058355. TGRS.2020.3002929.
Hu, L., Hu, C., He, M.X., 2017. Remote estimation of biomass of Ulva prolifera Wang, M., Hu, C., Cannizzaro, J., English, D., Han, X., Naar, D., et al., 2018. Remote
macroalgae in the Yellow Sea. Remote Sens. Environ. 192, 217–227. https://doi.org/ sensing of Sargassum biomass, nutrients, and pigments. Geophys. Res. Lett. 45
10.1016/j.rse.2017.01.037. https://doi.org/10.1029/2018GL078858.
Hu, C., Lu, Y., Sun, S., Liu, Y., 2021. Optical remote sensing of oil spills in the ocean: Wang, M., Shi, W., 2006. Cloud masking for ocean color data processing in the coastal
What’s really possible? J. Remote Sens. 9141902. https://doi.org/10.34133/2021/ regions. IEEE Trans. Geosci. Remote Sens. 44 (11), 3196–3205.
9141902. Wang, M., Shi, W., 2007. The NIR-SWIR combined atmospheric correction approach for
MODIS Ocean color data processing. Opt. Express 15, 15722–15733.

14

You might also like