Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (122)

Search Parameters:
Keywords = bio-optical products

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 9198 KiB  
Article
Estimating Vertical Distribution of Total Suspended Matter in Coastal Waters Using Remote-Sensing Approaches
by Hailong Zhang, Xin Ren, Shengqiang Wang, Xiaofan Li, Deyong Sun and Lulu Wang
Remote Sens. 2024, 16(19), 3736; https://doi.org/10.3390/rs16193736 - 8 Oct 2024
Viewed by 766
Abstract
The vertical distribution of the marine total suspended matter (TSM) concentration significantly influences marine material transport, sedimentation processes, and biogeochemical cycles. Traditional field observations are constrained by limited spatial and temporal coverage, necessitating the use of remote-sensing technology to comprehensively understand TSM variations [...] Read more.
The vertical distribution of the marine total suspended matter (TSM) concentration significantly influences marine material transport, sedimentation processes, and biogeochemical cycles. Traditional field observations are constrained by limited spatial and temporal coverage, necessitating the use of remote-sensing technology to comprehensively understand TSM variations over extensive areas and periods. This study proposes a remote-sensing approach to estimate the vertical distribution of TSM concentrations using MODIS satellite data, with the Bohai Sea and Yellow Sea (BSYS) as a case study. Extensive field measurements across various hydrological conditions and seasons enabled accurate reconstruction of in situ TSM vertical distributions from bio-optical parameters, including the attenuation coefficient, particle backscattering coefficient, particle size, and number concentration, achieving a determination coefficient of 0.90 and a mean absolute percentage error of 26.5%. In situ measurements revealed two distinct TSM vertical profile types (vertically uniform and increasing) and significant variation in TSM profiles in the BSYS. Using surface TSM concentrations, wind speed, and water depth, we developed and validated a remote-sensing approach to classify TSM vertical profile types, achieving an accuracy of 84.3%. Combining this classification with a layer-to-layer regression model, we successfully estimated TSM vertical profiles from MODIS observation. Long-term MODIS product analysis revealed significant spatiotemporal variations in TSM vertical distributions and column-integrated TSM concentrations, particularly in nearshore regions. These findings provide valuable insights for studying marine sedimentation and biological processes and offer a reference for the remote-sensing estimation of the TSM vertical distribution in other marine regions. Full article
Show Figures

Figure 1

Figure 1
<p>Study area and sampling stations in the Bohai Sea and the Yellow Sea during December 2016, April 2018, and July 2018. Six stations (S1, … S6; marked by the red circles) were selected to compare model-derived and in situ TSM vertical profiles shown in Figure 9 below. Transect 1 illustrates the selected transects used to showcase the spatiotemporal variations in satellite-derived TSM vertical profile depicted in Figure 11 below.</p>
Full article ">Figure 2
<p>Workflow for estimating TSM vertical profile data from satellite observations.</p>
Full article ">Figure 3
<p>The TSM vertical distribution in the BSYS: samples of the uniform type (<b>a</b>); samples of the increasing type (<b>b</b>). Schematic diagram of TMS vertical distribution types (<b>c</b>).</p>
Full article ">Figure 4
<p>Correlation between the TSM concentrations and different bio-optical paraments.</p>
Full article ">Figure 5
<p>Comparisons between the measured and estimated TSM concentrations in training (<b>a</b>) and testing dataset (<b>b</b>). Solid lines represent the 1:1 line.</p>
Full article ">Figure 6
<p>Statistics of the TSM vertical type in different seasons across various regions: (<b>a</b>) the whole BSYS; (<b>b</b>) the BS; (<b>c</b>) the NYS; (<b>d</b>) the SYS.</p>
Full article ">Figure 7
<p>Classification criteria of the TVTC method in the BSYS.</p>
Full article ">Figure 8
<p>Correlation between TSM concentrations at adjacent relative water layers.</p>
Full article ">Figure 9
<p>Comparison between the model-derived TSM vertical profile and in situ reconstructed TSM vertical profile for the example stations, with their locations shown in <a href="#remotesensing-16-03736-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 10
<p>Proportional distributions of the TSM vertical types in the BSYS from January to December: (<b>a</b>) the uniform type; (<b>b</b>) the increasing type.</p>
Full article ">Figure 11
<p>Monthly vertical distribution of TSM concentrations at the representative transect (32°N, 122°–126°E), with its location indicated in Transect 1 in <a href="#remotesensing-16-03736-f001" class="html-fig">Figure 1</a>, covering the period from January to December. The numbers on the color bar represent TSM concentration in logarithmic coordinates.</p>
Full article ">Figure 12
<p>Monthly distributions of the column-integrated TSM in the BSYS during the period from 2003 to 2021.</p>
Full article ">
21 pages, 16002 KiB  
Article
Comparative Studies on Nanocellulose as a Bio-Based Consolidating Agent for Ancient Wood
by Anastasia Fornari, Daniele Rocco, Leonardo Mattiello, Martina Bortolami, Marco Rossi, Laura Bergamonti, Claudia Graiff, Stefania Bani, Fabio Morresi and Fabiana Pandolfi
Appl. Sci. 2024, 14(17), 7964; https://doi.org/10.3390/app14177964 - 6 Sep 2024
Viewed by 549
Abstract
In this work, nanocellulose aqueous dispersions were studied as a bio-inspired consolidating agent for the recovery and conservation of ancient wood and compared with two of the most used traditional consolidants: the synthetic resins Paraloid B-72 and Regalrez 1126. The morphology of crystalline [...] Read more.
In this work, nanocellulose aqueous dispersions were studied as a bio-inspired consolidating agent for the recovery and conservation of ancient wood and compared with two of the most used traditional consolidants: the synthetic resins Paraloid B-72 and Regalrez 1126. The morphology of crystalline nanocellulose (CNC), determined using Scanning Electron Microscopy (SEM), presents with a rod-like shape, with a size ranging between 15 and 30 nm in width. Chemical characterization performed using the Fourier-Transform Infrared Spectroscopy (FT-IR) technique provides information on surface modifications, in this case, demonstrating the presence of only the characteristic peaks of nanocellulose. Moreover, conductometric, pH, and dry matter measurements were carried out, showing also in this case values perfectly conforming to what is found in the literature. The treated wood samples were observed under an optical microscope in reflected light and under a scanning electron microscope to determine, respectively, the damage caused by xylophages and the morphology of the treated surfaces. The images acquired show the greater similarity of the surfaces treated with nanocellulose to untreated wood, compared with other consolidating agents. Finally, a colorimetric analysis of these samples was also carried out before and after a first consolidation treatment, and after a second treatment carried out on the same samples three years later. The samples treated with CNC appeared very homogeneous and uniform, without alterations in their final color appearance, compared to other traditional synthetic products. Full article
(This article belongs to the Special Issue Advanced Technologies in Cultural Heritage)
Show Figures

Figure 1

Figure 1
<p>SEM micrograph of CNC.</p>
Full article ">Figure 2
<p>FTIR spectrum of CNC.</p>
Full article ">Figure 3
<p>Wood samples treated with CNC (A), Paraloid B-72 (B), and Regalrez 1126 (C), and samples untreated (NT) during the processes of impregnation: (<b>1</b>) before treatment; (<b>2</b>) immediately after; (<b>3</b>) after 19 h; and (<b>4</b>) after 50 days.</p>
Full article ">Figure 4
<p>Detailed images of samples treated with (<b>a</b>) CNC, (<b>b</b>) Paraloid B-72, and (<b>c</b>) Regalrez 1126, immediately after application.</p>
Full article ">Figure 5
<p>Reflected light microscope images of samples that were not-treated (NT) and those treated with CNC (A), Paraloid B-72 (B), and Regalrez 1126 (C). Images with UV light and 5× magnification (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>); images with visible light and 2.5× magnification (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>). White arrows indicate in all samples the signs of degradation due to the action of xylophagous insects.</p>
Full article ">Figure 6
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode 24 h after the consolidation treatment.</p>
Full article ">Figure 7
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode, one month after the consolidation treatment.</p>
Full article ">Figure 8
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode, three years after the first consolidating treatment.</p>
Full article ">Figure 9
<p>Reflectance spectra of untreated (NT) and treated samples (CNC, Paraloid B-72, Regalrez 1126), acquired in SCI and SCE mode, one week after the second consolidating treatment, carried out three years after the first treatment.</p>
Full article ">Figure 10
<p>SEM images of untreated wood sample; cross section in correspondence of a woodworm hole (<b>a</b>); longitudinal sections (<b>b</b>,<b>c</b>); magnification of a fiber channel (<b>c</b>).</p>
Full article ">Figure 11
<p>SEM images of untreated sample (<b>1</b>) and of the consolidant coating films of CNC (<b>2</b>), Paraloid B-72 (<b>3</b>), and Regalrez 1126 (<b>4</b>).</p>
Full article ">Figure 12
<p>SEM images of longitudinal section of wood samples, where it is possible to see the fibers channels: untreated sample (<b>1</b>); sample treated with CNC (<b>2</b>), Paraloid B-72 (<b>3</b>), Regalrez 1126 (<b>4</b>).</p>
Full article ">
25 pages, 8689 KiB  
Article
Assessment of Atmospheric Correction Algorithms for Sentinel-3 OLCI in the Amazon River Continuum
by Aline M. Valerio, Milton Kampel, Vincent Vantrepotte, Victoria Ballester and Jeffrey Richey
Remote Sens. 2024, 16(14), 2663; https://doi.org/10.3390/rs16142663 - 20 Jul 2024
Viewed by 1121
Abstract
Water colour remote sensing is a valuable tool for assessing bio-optical and biogeochemical parameters across the vast extent of the Amazon River Continuum (ARC). However, accurate retrieval depends on selecting the best atmospheric correction (AC). Four AC processors (Acolite, Polymer, C2RCC, OC-SMART) were [...] Read more.
Water colour remote sensing is a valuable tool for assessing bio-optical and biogeochemical parameters across the vast extent of the Amazon River Continuum (ARC). However, accurate retrieval depends on selecting the best atmospheric correction (AC). Four AC processors (Acolite, Polymer, C2RCC, OC-SMART) were evaluated against in situ remote sensing reflectance (Rrs) measurements. K-means classification identified four optical water types (OWTs) that are affected by the ARC. Two OWTs showed seasonal differences in the Lower Amazon River, influenced by the increase in suspended sediment concentration with river discharge. The other OWTs in the Amazon River Plume are dominated by phytoplankton or by a mixture of optically significant constituents. The Quality Water Index Polynomial method used to assess the quality of in situ and orbital Rrs had a high failure rate when the Apparent Visible Wavelength was >580 nm for in situ Rrs. OC-SMART Rrs products showed better spectral quality compared to Rrs derived from other AC processors evaluated in this study. These results improve our understanding of remotely sensing very turbid waters, such as those in the Amazon River Continuum. Full article
(This article belongs to the Special Issue Remote Sensing for the Study of the Changes in Wetlands)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sampling stations along the Amazon River Continuum used in this study. Each campaign has a different colour (see <a href="#remotesensing-16-02663-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 2
<p>Flowchart of the present study illustrating the overall methodology. The orange, blue and green lines represent the statistical comparisons made between the in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> and the satellite-derived <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> values after atmospheric correction.</p>
Full article ">Figure 3
<p>Match-ups of simulated in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> for S3-OLCI bands and <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> derived from different atmospheric correction processors for the same day: (<b>A</b>) Acolite; (<b>B</b>) Polymer; (<b>C</b>) C2RCC; (<b>D</b>) OC-SMART. The data were measured at the Amazon River Continuum, and the in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> were derived following the methodology proposed by [<a href="#B37-remotesensing-16-02663" class="html-bibr">37</a>] with the elimination of sun and sky glint as recommended by [<a href="#B38-remotesensing-16-02663" class="html-bibr">38</a>]. Scatter plots are presented in a log–log scale. Circles represent match-ups within a 3 h satellite pass window, while triangles represent match-ups outside the 3 h satellite pass window.</p>
Full article ">Figure 4
<p>Spectral variation in the statistical parameters between 400 and 779 nm: (<b>A</b>) Coefficient of determination (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">R</mi> <mn>2</mn> </msup> </mrow> </semantics></math>), (<b>B</b>) Root Mean Square Deviation (RMSD), (<b>C</b>) slope, (<b>D</b>) Mean Relative Absolute Difference (MRAD), (<b>E</b>) Mean Bias (MB) and (<b>F</b>) Valid Pixel (VP).</p>
Full article ">Figure 5
<p>Radar plot illustrating the statistical metrics used to evaluate the accuracy of the remote sensing reflectance for each atmospheric correction processor, using different approaches to estimate in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>. The green line represents the in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> processed according to [<a href="#B37-remotesensing-16-02663" class="html-bibr">37</a>] with sun and sky glint corrected according to the method proposed in [<a href="#B38-remotesensing-16-02663" class="html-bibr">38</a>] (M99 + R06). The blue line shows the same approach, not only considering the day of the in situ measurement but also one day before or after to increase the number of match-ups. The red line corresponds to the in situ<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> processed with the 3C model.</p>
Full article ">Figure 6
<p>(<b>A</b>) Optical Water Types (OWT) identified in the Amazon River Continuum and (<b>B</b>) their location for the different campaigns carried out (see <a href="#remotesensing-16-02663-f001" class="html-fig">Figure 1</a> and <a href="#remotesensing-16-02663-t001" class="html-table">Table 1</a> for more information on the field campaigns).</p>
Full article ">Figure 7
<p>(<b>A</b>) The QWIP relationship between Apparent Visible Wavelength (AVW) and the Normalised Difference Index (NDI) at blue-green and red bands, as described in [<a href="#B31-remotesensing-16-02663" class="html-bibr">31</a>], with the Amazon River Continuum (ARC) in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> dataset showing the different levels of QWIP values (±0.2 dashed grey line and ±0.3 dash-dotted grey line). Each optical water type found at the ARC is represented by a different colour. (<b>B</b>) Histogram of the AVW for our in situ ARC dataset.</p>
Full article ">Figure 8
<p>Mapped S3-OLCI image as an example (8 November 2019), where Apparent Visible Wavelength has been applied after using different atmospheric corrections: (<b>A</b>) Acolite; (<b>B</b>) Polymer; (<b>C</b>) C2RCC and (<b>D</b>) OC-SMART. The white line in the C2RCC image represents the transect used to extract pixels for evaluation. The same transect was applied to all four images processed with different atmospheric corrections.</p>
Full article ">Figure 9
<p>Apparent Visible Wavelength (AVW) values for the same pixel according to different atmospheric correction approaches, with longitudinal variability.</p>
Full article ">Figure 10
<p>Mapped S3-OLCI image as an example (8 November 2019), where Quality Water Index Polynomial score was calculated after applying different atmospheric corrections: (<b>A</b>) Acolite; (<b>B</b>) Polymer; (<b>C</b>) C2RCC and (<b>D</b>) OC-SMART. Black pixels are those outside the range of −0.2 to 0.2, as recommended by [<a href="#B31-remotesensing-16-02663" class="html-bibr">31</a>]. Pixels outside this range are considered as not passing the spectral quality.</p>
Full article ">Figure 11
<p>Match-ups of simulated in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> for S3-OLCI bands and <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> derived from different atmospheric correction processors for the same day: (<b>A</b>) Acolite; (<b>B</b>) Polymer; (<b>C</b>) C2RCC; (<b>D</b>) OC-SMART. The in situ and satellite data used for the match-up passed the QWIP score with an interval of ±0.3. The data were measured at the Amazon River Continuum, and the in situ <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> were derived following the methodology proposed by [<a href="#B37-remotesensing-16-02663" class="html-bibr">37</a>] and the elimination of sun and sky glint as recommended by [<a href="#B38-remotesensing-16-02663" class="html-bibr">38</a>]. Scatter plots are presented on a log–log scale. Circles represent match-ups within a 3 h satellite pass window, while triangles represent match-ups outside the 3 h satellite pass window.</p>
Full article ">Figure 12
<p>Spectral variation in the statistical parameters between 400 and 779 nm for the in situ and satellite data that passed the QWIP score with an interval of ±0.3: (<b>A</b>) Coefficient of determination (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">R</mi> <mn>2</mn> </msup> </mrow> </semantics></math>), (<b>B</b>) Root Mean Square Deviation (RMSD), (<b>C</b>) Slope, (<b>D</b>) Mean Relative Absolute Difference (MRAD), (<b>E</b>) Mean Bias (MB) and (<b>F</b>) Valid Pixel (VP).</p>
Full article ">Figure 13
<p>Performance evaluation according to the statistical metrics (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">R</mi> <mn>2</mn> </msup> </mrow> </semantics></math>, RMSD, MRAD, MB, VP). Light colours (white or yellow, closer to 0) are likely to have accurate <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> for a given optical water type (OWT: K1, K2, K3 and K4) and S3-OLCI band.</p>
Full article ">
26 pages, 4692 KiB  
Article
Development of a Greenhouse Wastewater Stream Utilization System for On-Site Microalgae-Based Biostimulant Production
by Sofia Faliagka, Georgios Kountrias, Eleni Dimitriou, Maria Álvarez-Gil, Mario Blanco-Vieites, Fabio Magrassi, Marta Notari, Eleftheria Maria Pechlivani and Nikolaos Katsoulas
AgriEngineering 2024, 6(3), 1898-1923; https://doi.org/10.3390/agriengineering6030111 - 21 Jun 2024
Viewed by 1213
Abstract
The challenges to feed the world in 2050 are becoming more and more apparent. This calls for producing more with fewer inputs (most of them under scarcity), higher resource efficiency, minimum or zero effect on the environment, and higher sustainability. Therefore, increasing the [...] Read more.
The challenges to feed the world in 2050 are becoming more and more apparent. This calls for producing more with fewer inputs (most of them under scarcity), higher resource efficiency, minimum or zero effect on the environment, and higher sustainability. Therefore, increasing the circularity of production systems is highly significant for their sustainability. This study investigates the utilization of waste streams from greenhouse hydroponic drainage nutrient solutions for the cultivation of the microalgae Desmodesmus sp. The cultivation was done in an automatically controlled container-scale closed tubular Photo Bio-Reactor (PBR). The study included lab-scale open-pond system experiments and in situ container-scale experiments in the greenhouse wastewater system to assess biomass growth, optical density, nitrogen consumption, and the influence of enzymatic complexes on microalgae cell breakdown. A batch-harvesting process was followed, and the harvested microalgae biomass was pre-concentrated using FeCl3 as a flocculant that has demonstrated efficient sedimentation and biomass recovery. Following microalgae sedimentation, the produced biomass was used for biostimulant production by means of a biocatalysis process. The enzymatic complexes, “EnzProt”, “EnzCell”, and “EnzMix” were tested for cell breakdown, with “EnzMix” at a dosage of 10% showing the most promising results. The results demonstrate successful biomass production and nitrogen uptake in the lab-scale open-pond system, with promising upscaling results within container-scale cultivation. The findings contribute to a better assessment of the needs of Desmodesmus sp. culture and highlight the importance in optimizing culture conditions and enzymatic processes for the production of biostimulants. Full article
Show Figures

Figure 1

Figure 1
<p>Agro-wastewater treatment plant based on a circular economy approach. Hydroponic crops generate a drainage solution that undergoes pre-treatment before entering the microalgae PBR. The microalgae harvest system (discharge/sedimentation tank) collects algae biomass, which can then be processed via enzymatic hydrolysis or biocatalysis to create a plant enabler to be applied on the crop.</p>
Full article ">Figure 2
<p>The main structure housing all components of the system, including (<b>a</b>) a unit control panel to monitor and control various systems within the container, (<b>b</b>) two air conditioners to provide temperature regulation, (<b>c</b>) a heat exchanger system to manage the temperature of the photobioreactors, (<b>d</b>) sensors to monitor environmental parameters, (<b>e</b>) LED panels to provide artificial illumination to support microalgae growth, (<b>f</b>) photobioreactor Loop 1 and Loop 2, (<b>g</b>) a storage wastewater tank, (<b>h</b>) a medium algae tank to store the culture medium, (<b>i</b>) a discharge/sedimentation tank for harvesting and biomass separation, (<b>j</b>) a cleaning water system to treat and clean water within the container, (<b>k</b>) a biofertilizer reactor to produce biofertilizers/biostimulants, and (<b>l</b>) an external unit for container acclimatization to control the internal climate conditions.</p>
Full article ">Figure 3
<p>Illustration of the biostimulant production process. The algae cultures cultivated in the PBR are transferred to the sedimentation tank, where they settle and concentrate, effectively separating the algae biomass from the water. The concentrated algal culture is then pumped into the fertilizer tank, which processes it into biofertilizer. This tank is equipped with a stirrer, a heating system, and a temperature controller (max. 60 °C) to ensure optimal processing conditions.</p>
Full article ">Figure 4
<p><span class="html-italic">Desmodesmus</span> sp. grown at (<b>a</b>) container scale for on-site biostimulant production while using (<b>b</b>) white and (<b>c</b>) RGB lights.</p>
Full article ">Figure 5
<p>Glass graduated cylinders at the end of the flocculation-coagulation experiments in <span class="html-italic">Desmodesmus</span> sp. cultures.</p>
Full article ">Figure 6
<p>Chlorophyll concentration expressed in absorbance at 680 nm of the “EnzMix”, “EnzCell”, “EnzProt”, and control treatment.</p>
Full article ">Figure 7
<p>Variation of the chlorophyll content (OD680) according to the percentage of “EnzMix” in relation to the biomass present in the sample.</p>
Full article ">Figure 8
<p>(<b>a</b>) Optical density at 750 nm (OD750, black dots) and at 680 nm (OD680, white dots), (<b>b</b>) dry biomass productivity (g L<sup>−1</sup>), and (<b>c</b>) nitrogen content (NO<sub>3</sub><sup>−</sup>-N, mg L<sup>−1</sup>) in <span class="html-italic">Desmodesmus</span> sp. culture grown in open-pond bioreactor for 50 days.</p>
Full article ">Figure 9
<p>(<b>a</b>) Optical density at 750 nm (OD750, black dots) and at 680 nm (OD680, white dots), and (<b>b</b>) dry biomass productivity (g L<sup>−1</sup>) in <span class="html-italic">Desmodesmus</span> sp. culture grown in open-pond bioreactor for 15 days.</p>
Full article ">Figure 10
<p>(<b>a</b>) Optical density at 750 nm (OD750, black dots) and at 680 nm (OD680, white dots) and (<b>b</b>) dry biomass productivity (g L<sup>−1</sup>) in the <span class="html-italic">Desmodesmus</span> sp. culture grown in the closed PBR for 33 days.</p>
Full article ">Figure 11
<p>Optical density at 750 nm (OD750, black dots) and at 680 nm (OD680, white dots) in the <span class="html-italic">Desmodesmus</span> sp. culture grown in the open-pond bioreactor for 40 days using the biomass initially grown in the container.</p>
Full article ">Figure 12
<p>(<b>a</b>) Optical density at 750 nm (OD750, black dots) and at 680 nm (OD680, white dots) and (<b>b</b>) dry weight (g L<sup>−</sup><sup>1</sup>) of the <span class="html-italic">Desmodesmus</span> sp. culture grown for 20 days in the open-pond bioreactor.</p>
Full article ">Figure 13
<p>(<b>a</b>) Optical density at 750 nm (OD750, black dots) and at 680 nm (OD680, white dots) and (<b>b</b>) dry weight (g L<sup>−</sup><sup>1</sup>) of the culture in the PBR system with CO<sub>2</sub> infusion.</p>
Full article ">Figure 14
<p>Black dots indicate the pH values in the PBR during the whole culture after CO<sub>2</sub> introduction to the system.</p>
Full article ">Figure 15
<p>Black dots indicate the (<b>a</b>) Calcium, (<b>b</b>) potassium, and (<b>c</b>) nitrogen content (mmol L<sup>−</sup><sup>1</sup>) in the <span class="html-italic">Desmodesmus</span> sp. culture in the PBR.</p>
Full article ">
14 pages, 2799 KiB  
Article
Evaluation of Ocean Color Algorithms to Retrieve Chlorophyll-a Concentration in the Mexican Pacific Ocean off the Baja California Peninsula, Mexico
by Patricia Alvarado-Graef, Beatriz Martín-Atienza, Ramón Sosa-Ávalos, Reginaldo Durazo and Rafael Hernández-Walls
Remote Sens. 2024, 16(10), 1748; https://doi.org/10.3390/rs16101748 - 15 May 2024
Viewed by 916
Abstract
Mathematical algorithms relate satellite data of ocean color with the surface Chlorophyll-a concentration (Chl-a), a proxy of phytoplankton biomass. These mathematical tools work best when they are adapted to the unique bio-optical properties of a particular oceanic province. Ocean color [...] Read more.
Mathematical algorithms relate satellite data of ocean color with the surface Chlorophyll-a concentration (Chl-a), a proxy of phytoplankton biomass. These mathematical tools work best when they are adapted to the unique bio-optical properties of a particular oceanic province. Ocean color algorithms should also consider that there are significant differences between datasets derived from different sensors. Common solutions are to provide different parameters for each sensor or use merged satellite data. In this paper, we use satellite data from the Copernicus merged product suite and in situ data from the southernmost part of the California Current System to test two widely used global algorithms, OCx and CI, and a regional algorithm, CalCOFI2. The OCx algorithm yielded the most favorable results. Consequently, we regionalized it and conducted further testing, leading to significant improvements, especially in eutrophic and oligotrophic waters. The database was then separated according to (a) dynamic boundaries in the area, (b) bio-optical properties, and (c) climatic conditions (El Niño/La Niña). Regional algorithms were obtained and tested for each partition. The Chl-a retrievals for each model were tested and compared. The best fit for the data was for the regional algorithms that considered the climatic conditions (El Niño/La Niña). These results will allow for the construction of consistent regionally adapted time series and, therefore, will demonstrate the importance of El Niño/La Niña events on the bio-optical properties of the area. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>General grid of IMECOCAL stations where in situ samples were collected. The line number increases from 100 toward the south on line 137. Numbers above transects indicate station numbers, which increase with distance from the coast. The blue arrows indicate the main directions of the California Current System in the region.</p>
Full article ">Figure 2
<p>Results from applying OCx (blue squares), CI (yellow circles), and CalCOFI2 (pink triangles) algorithms to remote sensing reflectance data (<math display="inline"><semantics> <mrow> <mi mathvariant="normal">F</mi> <mo>=</mo> <msub> <mrow> <mi mathvariant="normal">l</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <mfenced open="[" close="]" separators="|"> <mrow> <mi mathvariant="fraktur">R</mi> </mrow> </mfenced> </mrow> </semantics></math>, Equation (2)) off the coast of Baja California compared to the log10 of in situ Chl-<span class="html-italic">a</span> data (gray dots).</p>
Full article ">Figure 3
<p>Regional (red circles, Equation (7)) and OCx (blue squares) algorithm retrievals applied to remotely sensed reflectance data (<math display="inline"><semantics> <mrow> <mi mathvariant="normal">F</mi> <mo>=</mo> <msub> <mrow> <mi mathvariant="normal">l</mi> <mi mathvariant="normal">o</mi> <mi mathvariant="normal">g</mi> </mrow> <mrow> <mn>10</mn> </mrow> </msub> <mfenced open="[" close="]" separators="|"> <mrow> <mi mathvariant="fraktur">R</mi> </mrow> </mfenced> </mrow> </semantics></math>, Equation (2)) off the coast of Baja California compared to the log10 of in situ Chl-<span class="html-italic">a</span> data (gray dots).</p>
Full article ">Figure 4
<p>(<b>a</b>) Variation of green/blue reflectance ratio (<math display="inline"><semantics> <mrow> <mi mathvariant="fraktur">R</mi> </mrow> </semantics></math><sup>−1</sup>) with the station number (<a href="#remotesensing-16-01748-f001" class="html-fig">Figure 1</a>) that increases with the distance from the coast. The stations were classified as coastal waters with a green/blue ratio of reflectance of 1 or higher (red squares, 105 stations), transitional waters with a green/blue ratio between 0.5 and 1 (green crosses, 653 stations), and oceanic waters with a green/blue ratio of 0.5 and under (blue circles, 2607 stations). (<b>b</b>) Station locations.</p>
Full article ">Figure 5
<p>Retrievals of log<sub>10</sub> of mChl-<span class="html-italic">a</span>, using Equation (8), for coastal waters (dotted line), Equation (9) for transitional waters (solid line), and Equation (10) for oceanic waters (dashed line), compared to the log<sub>10</sub> of in situ Chl-<span class="html-italic">a</span> data (red squares for coastal waters, green crosses for transitional waters, and blue circles for oceanic waters).</p>
Full article ">Figure 6
<p>Algorithm retrievals for data classified as (<b>a</b>) La Niña, (<b>b</b>) El Niño, (<b>c</b>) or normal compared to the log<sub>10</sub> of in situ Chl-<span class="html-italic">a</span> data.</p>
Full article ">
21 pages, 4333 KiB  
Article
Tracking Water Quality and Macrophyte Changes in Lake Trasimeno (Italy) from Spaceborne Hyperspectral Imagery
by Alice Fabbretto, Mariano Bresciani, Andrea Pellegrino, Krista Alikas, Monica Pinardi, Salvatore Mangano, Rosalba Padula and Claudia Giardino
Remote Sens. 2024, 16(10), 1704; https://doi.org/10.3390/rs16101704 - 11 May 2024
Viewed by 1940
Abstract
This work aims to show the potential of imaging spectroscopy in assessing water quality and aquatic vegetation in Lake Trasimeno, Italy. Hyperspectral reflectance data from the PRISMA, DESIS and EnMAP missions (2019–2022, summer periods) were compared with in situ measurements from WISPStation and [...] Read more.
This work aims to show the potential of imaging spectroscopy in assessing water quality and aquatic vegetation in Lake Trasimeno, Italy. Hyperspectral reflectance data from the PRISMA, DESIS and EnMAP missions (2019–2022, summer periods) were compared with in situ measurements from WISPStation and used as inputs for water quality product generation algorithms. The bio-optical model BOMBER was run to simultaneously retrieve water quality parameters (Chlorophyll-a (Chl-a) and Total Suspended Matter, (TSM)) and the coverage of submerged and emergent macrophytes (SM, EM); value-added products, such as Phycocyanin concentration maps, were generated through a machine learning approach. The results showed radiometric agreement between satellite and in situ data, with R2 > 0.9, a Spectral Angle < 10° and water quality mapping errors < 30%. Both SM and EM coverage varied significantly from 2019 (135 ha, 0 ha, respectively) to 2022 (2672 ha, 343 ha), likely influenced by changes in rainfall and lake levels. The areas of greatest variability in Chl-a and TSM were identified in the littoral zones in the western side of the lake, while the highest variation in the fractional cover of SM and density of EM were observed in the south-eastern region; this information could support the water authorities’ monitoring activities. To this end, further developments to improve the reference field data for the validation of water quality products are recommended. Full article
Show Figures

Figure 1

Figure 1
<p>On the left, a true color composition image as acquired by DESIS on 4 August 2019 showing the position of the WISPStation (43.122, 12.134—red box); on the right, a picture of the platform with the WISPStation.</p>
Full article ">Figure 2
<p>Time series of the environmental data: air temperature (°C) shown in orange, precipitation (mm) shown in light blue and lake level (m) shown in green. The four summer periods in which satellite images were acquired are highlighted in grey with dashed lines.</p>
Full article ">Figure 3
<p>Flowchart of the methodology applied in the study. The oval with a black outline represents the input products where the study started from. Grey boxes indicate the methodology applied. Green diamond shapes stand for decision-making steps in the process. Blue parallelograms represent products generated, and the violet oval indicates the end point of the process.</p>
Full article ">Figure 4
<p>Comparison of DESIS and in situ Rrs data, before (“DESIS”, dark blue) and after (“DESIS deglint”, light blue) sun glint removal. In situ data are displayed in orange. Statistical results are displayed in the boxes.</p>
Full article ">Figure 5
<p>Comparisons of the average Rrs values gathered from the spaceborne data and corresponding in situ Rrs data. The variability in the mean spectra of PRISMA (6 images) and DESIS (6 images) is displayed as blue curves, with the shaded blue area representing the standard deviation. The mean and standard deviation of the in situ data are equivalently shown in orange. In the case of the EnMAP data, the comparison is limited to a single image, and it is shown with the same color configuration. In this case, the standard deviation refers to the variability present in the ROI and in the set of three measurements of the in situ data. The statistical results are displayed in the boxes.</p>
Full article ">Figure 6
<p>Water quality maps and bottom characterization for the 13 images of the available dataset. From <b>left</b> to <b>right</b>: Chl-a, TSM and PC maps; bottom characterization products, in terms of emergent macrophytes and the three cover classes: b0 (semi-emergent macrophytes), b1 (permanently submerged macrophytes), b2 (sand).</p>
Full article ">Figure 6 Cont.
<p>Water quality maps and bottom characterization for the 13 images of the available dataset. From <b>left</b> to <b>right</b>: Chl-a, TSM and PC maps; bottom characterization products, in terms of emergent macrophytes and the three cover classes: b0 (semi-emergent macrophytes), b1 (permanently submerged macrophytes), b2 (sand).</p>
Full article ">Figure 7
<p>Distributions of sand and submerged macrophyte cover classes (sparse, moderate and dense) in the four-year study period. Un-colonized (sand) pixel percentage is represented in yellow; sparse, moderate and dense submerged macrophytes are shown with a gradient of green from lightest to darkest.</p>
Full article ">Figure 8
<p>From <b>left</b> to <b>right</b>, standard deviation maps of Chl-a, PC, TSM, submerged macrophytes’ fractional cover and emergent macrophytes’ density (WAVI).</p>
Full article ">
12 pages, 3245 KiB  
Article
Transparent and Efficient Wood-Based Triboelectric Nanogenerators for Energy Harvesting and Self-Powered Sensing
by Ting Cheng, Kunli Cao, Yidan Jing, Hongyan Wang and Yan Wu
Polymers 2024, 16(9), 1208; https://doi.org/10.3390/polym16091208 - 26 Apr 2024
Cited by 1 | Viewed by 1095
Abstract
Wood possesses several advantageous qualities including innocuity, low cost, aesthetic appeal, and excellent biocompatibility, and its naturally abundant functional groups and diverse structural forms facilitate functionalization modification. As the most sustainable bio-based material, the combination of wood with triboelectric nanogenerators (TENGs) stands poised [...] Read more.
Wood possesses several advantageous qualities including innocuity, low cost, aesthetic appeal, and excellent biocompatibility, and its naturally abundant functional groups and diverse structural forms facilitate functionalization modification. As the most sustainable bio-based material, the combination of wood with triboelectric nanogenerators (TENGs) stands poised to significantly advance the cause of green sustainable production while mitigating the escalating challenges of energy consumption. However, the inherent weak polarizability of natural wood limits its development for TENGs. Herein, we present the pioneering development of a flexible transparent wood-based triboelectric nanogenerator (TW-TENG) combining excellent triboelectrical properties, optical properties, and wood aesthetics through sodium chlorite delignification and epoxy resin impregnation. Thanks to the strong electron-donating groups in the epoxy resin, the TW-TENG obtained an open-circuit voltage of up to ~127 V, marking a remarkable 530% enhancement compared to the original wood. Furthermore, durability and stability were substantiated through 10,000 working cycles. In addition, the introduction of epoxy resin and lignin removal endowed the TW-TENG with excellent optical characteristics, with optical transmittance of up to 88.8%, while preserving the unique texture and aesthetics of the wood completely. Finally, we show the application prospects of TW-TENGs in the fields of self-power supply, motion sensing, and smart home through the demonstration of a TW-TENG in the charging and discharging of capacitors and the output of electrical signals in different scenarios. Full article
(This article belongs to the Special Issue High Proformance Wood Coating)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Fabrication process of the TW-TENG; (<b>b</b>) <b>i</b>:Optical photographs of TW with different delignification time, <b>ii</b>: Flexibility of DW vs TW.</p>
Full article ">Figure 2
<p>Morphological and chemical compositional characterization of NW before and after treatment. (<b>a</b>,<b>b</b>) Cross-sectional SEM images of NW; (<b>c</b>,<b>d</b>) cross-sectional SEM images of DW120; (<b>e</b>,<b>f</b>) cross-sectional SEM images of TW120. (<b>g</b>) FTIR infrared spectra of NW, DW120, and TW120 and magnified images within the dashed line. (<b>h</b>) Relative contents of cellulose, hemicellulose, and lignin in DW with different degrees of delignification treated using the sodium chlorite method.</p>
Full article ">Figure 3
<p>Characterization of the optical and mechanical properties of TW with different degrees of delignification. (<b>a</b>) Optical transmittance of TW with different degrees of delignification. (<b>b</b>) Haze of TW with different degrees of delignification. (<b>c</b>) Tensile strength and elongation at break of TW with different degrees of delignification. (<b>d</b>) Stress–strain curves of TW with different degrees of delignification.</p>
Full article ">Figure 4
<p>Electrical output performance tests for the TW-TENGs. (<b>a</b>) Triboelectrical performance Voc versus delignification treatment length. (<b>b</b>–<b>d</b>) Open-circuit voltage (Voc), short-circuit current (Isc), and transfer charge (Qsc) of TW120-TENG at different working frequencies. (<b>e</b>–<b>g</b>) Open-circuit voltage (Voc), short-circuit current (Isc), and transfer charge (Qsc) of TW120-TENGs at different working displacements. (<b>h</b>) The 10,000 contact-separation cycles of TW120-TENG.</p>
Full article ">Figure 5
<p>Demonstration of the practical application of the TW-TENG. (<b>a</b>) The TW-TENG lights up 19 small LED bulbs. (<b>b</b>) Open-circuit voltage (Voc) is generated by triggering the TW-TENG in different motion modes. (<b>c</b>) The TW-TENG transmits the message of “NJFU” in Morse code. (<b>d</b>) The TW-TENG charges a capacitor and lights up an LED screen with the capacitor as the power source. (<b>e</b>) The TW-TENG serves as a motion sensor to monitor human movement.</p>
Full article ">
20 pages, 5657 KiB  
Article
Revalorization of Yerba Mate Residues: Biopolymers-Based Films of Dual Wettability as Potential Mulching Materials
by Laura M. Sanchez, Jorge de Haro, Eva Domínguez, Alejandro Rodríguez, Antonio Heredia and José J. Benítez
Polymers 2024, 16(6), 815; https://doi.org/10.3390/polym16060815 - 14 Mar 2024
Viewed by 1445
Abstract
Biodegradable mulching films are a very attractive solution to agronomical practices intended to achieve more successful crop results. And, in this context, the employment of agricultural and industrial food residues as starting material for their production is an alternative with economic and environmental [...] Read more.
Biodegradable mulching films are a very attractive solution to agronomical practices intended to achieve more successful crop results. And, in this context, the employment of agricultural and industrial food residues as starting material for their production is an alternative with economic and environmental advantages. This work reports the preparation of bilayer films having two different wettability characteristics from three bio-derived biopolymers: TEMPO-oxidized cellulose nanofibers isolated from infused Yerba Mate residues, Chitosan and Polylactic acid. The infused Yerba Mate residues, the isolated and oxidized cellulose nanofibers, and the films were characterized. Nanofibrillation yield, optical transmittance, cationic demand, carboxyl content, intrinsic viscosity, degree of polymerization, specific surface area and length were studied for the (ligno)cellulose nanofibers. Textural and chemical analysis, thermal and mechanical properties studies, as well as water and light interactions were included in the characterization of the films. The bilayer films are promising materials to be used as mulching films. Full article
(This article belongs to the Section Biobased and Biodegradable Polymers)
Show Figures

Figure 1

Figure 1
<p>Thickness values of all prepared films.</p>
Full article ">Figure 2
<p>SEM micrographs of the films’ hydrophilic surface: bilayer film ((<b>a</b>) ×1000 and (<b>c</b>) ×5000) and monolayer hydrophilic film ((<b>b</b>) ×1000 and (<b>d</b>) ×5000). Scale: 10 µm.</p>
Full article ">Figure 3
<p>SEM micrographs of the films’ hydrophobic surface: bilayer film ((<b>a</b>) ×1000 and (<b>c</b>) ×5000) and monolayer hydrophobic film ((<b>b</b>) ×1000 and (<b>d</b>) ×5000). Scale: 10 µm.</p>
Full article ">Figure 4
<p>SEM micrographs of the films’ cross-section: (<b>a</b>) bilayer film, (<b>b</b>) monolayer hydrophilic film and (<b>c</b>) monolayer hydrophobic film. Scale: 10 µm.</p>
Full article ">Figure 5
<p>FTIR spectra of films prepared from just the starting materials (TO-YCNF and CH), and of the prepared mono- and bilayer films. The bilayer film was irradiated by the hydrophobic and hydrophilic sides, respectively.</p>
Full article ">Figure 6
<p>DSC thermograms of the mono- and bilayer films.</p>
Full article ">Figure 7
<p>(<b>a</b>) TGA thermograms of the mono- and bilayer films and (<b>b</b>) their corresponding first derivatives (DTGA).</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) TGA thermograms of the mono- and bilayer films and (<b>b</b>) their corresponding first derivatives (DTGA).</p>
Full article ">Figure 8
<p>(<b>a</b>) Tensile tests of the mono- and bilayer films; (<b>b</b>) Tensile tests of both the original and the hot-pressed bilayer material.</p>
Full article ">Figure 9
<p>WCA of the external and internal surfaces of the bilayer material.</p>
Full article ">Figure 10
<p>Water vapor absorption tests of the prepared mono- and bilayer films.</p>
Full article ">Figure 11
<p>UV–Visible transmittance of the prepared mono- and bilayer films. The bilayer film was irradiated by the hydrophobic and hydrophilic sides, respectively. UV-C, UV-B, UV-A, and Visible regions of the electromagnetic spectrum are indicated by using shades of blue and white colors in the figure.</p>
Full article ">Scheme 1
<p>Workflow for the preparation and study of the bilayer films.</p>
Full article ">
19 pages, 2875 KiB  
Article
Estimating the Colored Dissolved Organic Matter in the Negro River, Amazon Basin, with In Situ Remote Sensing Data
by Rogério Ribeiro Marinho, Jean-Michel Martinez, Tereza Cristina Souza de Oliveira, Wagner Picanço Moreira, Lino A. Sander de Carvalho, Patricia Moreira-Turcq and Tristan Harmel
Remote Sens. 2024, 16(4), 613; https://doi.org/10.3390/rs16040613 - 6 Feb 2024
Cited by 3 | Viewed by 1517
Abstract
Dissolved organic matter (DOM) is a crucial component of continental aquatic ecosystems. It plays a vital role in the carbon cycle by serving as a significant source and reservoir of carbon in water. DOM provides energy and nutrients to organisms, affecting primary productivity, [...] Read more.
Dissolved organic matter (DOM) is a crucial component of continental aquatic ecosystems. It plays a vital role in the carbon cycle by serving as a significant source and reservoir of carbon in water. DOM provides energy and nutrients to organisms, affecting primary productivity, organic composition, and the food chain. This study presents empirical bio-optical models for estimating the absorption of colored dissolved organic matter (aCDOM) in the Negro River using in situ remote sensing reflectance (Rrs) data. Physical–chemical data (TSS, DOC, and POC) and optical data (aCDOM and Rrs) were collected from the Negro River, its tributaries, and lakes and empirical relationships between aCDOM at 440 nm, single band, and the ratio bands of Rrs were assessed. The analysis of spectral slope shows no statistically significant correlations with DOC concentration or aCDOM absorption coefficient. However, strong relationships were observed between DOC and aCDOM (R2 = 0.72), aCDOM and Rrs at 650 nm (R2 > 0.80 and RMSE < 1.75 m−1), as well as aCDOM and the green/red band ratio (R2 > 0.80 and RMSE < 2.30 m−1). aCDOM displayed large spatial and temporal variations, varying from 1.9 up to 20.1 m−1, with higher values in rivers of the upper course of the Negro basin and lower values in rivers with total solids suspended > 10 mg·L−1. Environmental factors that influence the production of dissolved organic matter include soil type, dense forest cover, high precipitation, and low erosion rates. This study demonstrated that aCDOM can serve as an indicator of DOC, and Rrs can serve as an indicator of aCDOM in the Negro basin. Our findings offer a starting point for future research on the optical properties of Amazonian black-water rivers. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study area and sampling sections.</p>
Full article ">Figure 2
<p>Examples of Rrs collected in different sections of the Negro River (black-water), its tributaries (black- and clear-water), and the Amazon River (white-water). (<b>a</b>) High-water period; (<b>b</b>) low-water period.</p>
Full article ">Figure 3
<p>aCDOM of the Negro River (<b>a</b>) and tributaries (<b>b</b>). Note that the <span class="html-italic">Y</span>-axis scale in (<b>a</b>) is different from the <span class="html-italic">Y</span>-axis scale in <a href="#remotesensing-16-00613-f003" class="html-fig">Figure 3</a>b. Refer to <a href="#remotesensing-16-00613-t001" class="html-table">Table 1</a> for the names of the sections in (<b>b</b>).</p>
Full article ">Figure 4
<p>DOC versus aCDOM in the Negro River Basin. The dashed black line indicates a 1:1 relationship. Please refer to <a href="#remotesensing-16-00613-t001" class="html-table">Table 1</a> for the names of the sections.</p>
Full article ">Figure 5
<p>Relationship between aCDOM at 440 nm and the ratio of Rrs for MSI bands simulated at 560 nm and 665 nm in the high-water period (<b>a</b>) and low-water period (<b>b</b>). The blue lines are 95% confidence intervals.</p>
Full article ">Figure 6
<p>The relationship between aCDOM at 440 nm and the Rrs for MSI band simulated at 665 nm for the high-water period (<b>a</b>) and low-water period (<b>b</b>). The blue lines represent the 95% confidence intervals.</p>
Full article ">Figure 7
<p>Spatial distribution of aCDOM at 440 nm in the Negro basin, Amazon River, and the coast of French Guiana. Source: This study and GLORIA [<a href="#B73-remotesensing-16-00613" class="html-bibr">73</a>].</p>
Full article ">
22 pages, 615 KiB  
Article
The Nutritional Value of Plant Drink against Bovine Milk—Analysis of the Total Concentrations and the Bio-Accessible Fraction of Elements in Cow Milk and Plant-Based Beverages
by Maja Welna, Anna Szymczycha-Madeja, Anna Lesniewicz and Pawel Pohl
Processes 2024, 12(1), 231; https://doi.org/10.3390/pr12010231 - 21 Jan 2024
Cited by 2 | Viewed by 1390
Abstract
Four types of non-dairy (plant) drinks—almond, oat, rice, and soy—as well as cow milk with varying fat contents (1.5%, 2.0%, and 3.2%), were examined and compared in terms of the total concentrations of Al, As, B, Ba, Ca, Cd, Cr, Cu, Fe, K, [...] Read more.
Four types of non-dairy (plant) drinks—almond, oat, rice, and soy—as well as cow milk with varying fat contents (1.5%, 2.0%, and 3.2%), were examined and compared in terms of the total concentrations of Al, As, B, Ba, Ca, Cd, Cr, Cu, Fe, K, Mg, Na, Mn, Ni, P, Pb, Sb, Se, Sr, and Zn using inductively coupled optical emission spectrometry (ICP OES). Additionally, in vitro gastrointestinal digestion was used to determine the bio-accessible fraction of selected elements, evaluating the nutritional value and risk assessment involved with the consumption of these beverages. A significant difference in the mineral profile was observed depending on the type of plant drink, with the highest content of elements noted in the soy drink and the lowest in the rice drink. Except for Ca and P, the soy drink appears to be a much better source of essential nutrients, including Cu, Fe, and Mn, than cow’s milk. A similar Ca content in plant beverages can be obtained only by adding calcium salt at the stage of its production. Interestingly, by using the multivariate data analysis, the average content of the selected elements (Cu, K, Na, P, and Zn) can be used both to differentiate dairy and non-dairy milk samples according to their type and to distinguish plant drinks from milk of animal origin. The bio-accessibility of essential elements (Ca, Cu, Fe, Mg, Mn, P, Zn) in cow milk was within 8.37–98.2% and increased with an increase in its fat content. Accordingly, by drinking 1 L of this milk daily, it is possible to contribute to the recommended dietary intakes of Ca, P, Cu, Mg, and Zn between 5.6–68%. Although the bio-accessibility of elements in the rice drink was the highest (9.0–90.8%), the soy drink seems to be the best source of nutrients in bioavailable forms; its consumption (1 L/day) covers the requirements of Cu, Mn, Mg, Ca, P, and Zn in 7.0–67%. Unfortunately, both groups of beverages are not important sources of Fe (plant drink) and Mn or Fe (cow milk) in the human diet. On the other hand, potentially toxic elements (Al, B, Ba) were found in them in a relatively inert form. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Biplots for the two first principal components (PC1 and PC2) showing beverage samples grouped into cow milk (CM) and almond (PDA), oat (PDO), rice (PDR), and soy (PDS) plant-type drinks. A different number of variables was used to establish the PC1 and the PC2: (<b>a</b>) 17 (mean concentrations of Al, As, B, Ba, Ca, Cr, Cu, Fe, K, Mg, Mn, Na, Ni, P, Se, Sr, and Zn), (<b>b</b>) 14 (mean concentrations of Al, B, Ba, Ca, Cr, Cu, Fe, K, Mg, Mn, Na, P, Sr, and Zn), (<b>c</b>) 7 (mean concentrations of B, Cu, K, Mg, Na, P, and Zn), and (<b>d</b>) 3 (mean concentrations of B, Mg, and P).</p>
Full article ">
20 pages, 2441 KiB  
Article
Soft Epidermal Paperfluidics for Sweat Analysis by Ratiometric Raman Spectroscopy
by Ata Golparvar, Lucie Thenot, Assim Boukhayma and Sandro Carrara
Biosensors 2024, 14(1), 12; https://doi.org/10.3390/bios14010012 - 25 Dec 2023
Cited by 3 | Viewed by 3882
Abstract
The expanding interest in digital biomarker analysis focused on non-invasive human bodily fluids, such as sweat, highlights the pressing need for easily manufactured and highly efficient soft lab-on-skin solutions. Here, we report, for the first time, the integration of microfluidic paper-based devices (μPAD) [...] Read more.
The expanding interest in digital biomarker analysis focused on non-invasive human bodily fluids, such as sweat, highlights the pressing need for easily manufactured and highly efficient soft lab-on-skin solutions. Here, we report, for the first time, the integration of microfluidic paper-based devices (μPAD) and non-enhanced Raman-scattering-enabled optical biochemical sensing (Raman biosensing). Their integration merges the enormous benefits of μPAD, with high potential for commercialization and use in resource-limited settings, with biorecognition-element-free (but highly selective) optical Raman biosensing. The introduced thin (0.36 mm), ultra-lightweight (0.19 g), and compact footprint (3 cm2) opto-paperfluidic sweat patch is flexible, stretchable, and conforms, irritation-free, to hairless or minimally haired body regions to enable swift sweat collection. As a great advantage, this new bio-chemical sensory system excels through its absence of onboard biorecognition elements (bioreceptor-free) and omission of plasmonic nanomaterials. The proposed easy fabrication process is adaptable to mass production by following a fully sustainable and cost-effective process utilizing only basic tools by avoiding typically employed printing or laser patterning. Furthermore, efficient collection and transportation of precise sweat volumes, driven exclusively by the wicking properties of porous materials, shows high efficiency in liquid transportation and reduces biosensing latency by a factor of 5 compared to state-of-the-art epidermal microfluidics. The proposed unit enables electronic chip-free and imaging-less visual sweat loss quantification as well as optical biochemical analysis when coupled with Raman spectroscopy. We investigated the multimodal quantification of sweat urea and lactate levels ex vivo (with syntactic sweat including +30 sweat analytes on porcine skin) and achieved a linear dynamic range from 0 to 100 mmol/L during fully dynamic continuous flow characterization. Full article
(This article belongs to the Special Issue SERS-Based Biosensors: Design and Biomedical Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The introduced paper-based epidermal soft optofluidic device (referred to as “epi-opto-paper-fluidics”) positioned on the epidermis to facilitate non-invasive and label-free biochemical analysis of sweat components (when coupled with a Raman spectroscopy) and simultaneous visual quantification of sweat loss volume. (<b>b</b>) It comprises sequential layers: a PDMS encapsulator layer atop, a precisely patterned cellulose paper-based channel layer at the core, an integrated, flexible laser blocker tape, and an underlying adhesive layer. (<b>c</b>) The resulting congested Raman scattering spectra are dominated by cellulose and PDMS-linked Raman shifts. The influence of sweat samples manifested with selective Raman shifts correlated linearly in intensity with concentrations of sweat biomarkers such as lactate and urea. To improve the visualization of the unit, a diluted Rhodamine 6G solution is introduced in the inlet, which is wicked entirely by the paperfluidic changing its color to pink. In reality, the white channel color transitions to grayish upon sweat absorption, like typical wetted paper.</p>
Full article ">Figure 2
<p>The three phases of developing the introduced soft microfluidic device: design and optimization through mechanical simulations, printing-free facile fabrication using elementary tools, and in situ applications. (<b>a</b>) The initial design of the channel layer, (<b>b</b>) optimized design directed to enhance mechanical robustness. The evaluation of mechanical deformation is demonstrated through in vitro experiments following the removal of protective packaging: (<b>c</b>) involves axial stretching, (<b>d</b>) in-plan bending, and (<b>e</b>) twisting tests. Subsequently, the mechanical resilience is further investigated post-attachment to a female epidermis over varying temporal frames: (<b>f</b>–<b>i</b>) visually elucidates the sustained resilience of the device under applied stress. The channel incorporates discrete black markings within its PDMS layer, which serve as markers enabling real-time visualization of fluid front progression. This visualization is harnessed for accurately determining sweat volume intake (sweat loss quantification) to estimate sweat rate.</p>
Full article ">Figure 3
<p>(<b>a</b>) Unprocessed Raman spectra (without baseline correction and smoothing) as acquired from human sweat induced by varying exercise intensities, manifesting different sweat rates. Highlighted within these spectra are the shifts corresponding to the characteristic Raman shifts of lactate and urea, positioned at around 855 and 1005 cm<sup>−1</sup>, respectively, and a broad water-related band centered around 1650 cm<sup>−1</sup>. Offsets were added to spectra for visualization. (<b>b</b>) The intensity changes in lactate and urea Raman bands, which correspond to the concentration change in sweat collected with diverse sweat rates, reveal the discernible impact of the sweat rate on the analyte concentration.</p>
Full article ">Figure 4
<p>Flow kinetics analysis in the developed paper-based optofluidics. (<b>a</b>) Fluid front position over time for 20 μL sweat droplet, which required ~319 s to traverse a distance of ~70 mm. (<b>b</b>) The relationship between travel distance and time is examined for a 20 μL droplet within the microfluidic and its theoretical estimation based on Lucas–Washburn equation. (<b>c</b>) Travel distance over time for droplets of varying volumes. The investigation spans a range of droplet sizes, effectively highlighting the diverse rates at which these droplets navigate the paperfluidic landscape. (<b>d</b>) Highly linear calibration curve for volume uptake using the position of the terminal point of each droplet (fluid front) as shown in (<b>c</b>).</p>
Full article ">Figure 5
<p>The bioreceptor- and label-free optical Raman scattering biochemical sensing (Raman biosensing) application in sweat analysis utilizing proposed paper-based soft optofluidics for swift sweat collection and transportation. (<b>a</b>) Raman scattering spectra from each component of the soft microfluidic platform separately. (<b>b</b>) The acquired ex vivo Raman spectra (unprocessed), obtained under dynamic conditions, emerge as a composite of PDMS, sweat, and cellulose contributions. Distinct small-scale Raman bands associated with sweat are discernible. Offsets were added for clarity. (<b>c</b>) The calibration curve was derived from the analysis of sweat lactate content. (<b>d</b>) Urea biosensing during dynamic in-flow measurements.</p>
Full article ">
16 pages, 6066 KiB  
Article
Development and Application of an Automated Raman Sensor for Bioprocess Monitoring: From the Laboratory to an Algae Production Platform
by Wiviane Wieser, Antony Ali Assaf, Benjamin Le Gouic, Emmanuel Dechandol, Laura Herve, Thomas Louineau, Omar Hussein Dib, Olivier Gonçalves, Mariana Titica, Aurélie Couzinet-Mossion, Gaetane Wielgosz-Collin, Marine Bittel and Gerald Thouand
Sensors 2023, 23(24), 9746; https://doi.org/10.3390/s23249746 - 11 Dec 2023
Cited by 3 | Viewed by 1622
Abstract
Microalgae provide valuable bio-components with economic and environmental benefits. The monitoring of microalgal production is mostly performed using different sensors and analytical methods that, although very powerful, are limited to qualified users. This study proposes an automated Raman spectroscopy-based sensor for the online [...] Read more.
Microalgae provide valuable bio-components with economic and environmental benefits. The monitoring of microalgal production is mostly performed using different sensors and analytical methods that, although very powerful, are limited to qualified users. This study proposes an automated Raman spectroscopy-based sensor for the online monitoring of microalgal production. For this purpose, an in situ system with a sampling station was made of a light-tight optical chamber connected to a Raman probe. Microalgal cultures were routed to this chamber by pipes connected to pumps and valves controlled and programmed by a computer. The developed approach was evaluated on Parachlorella kessleri under different culture conditions at a laboratory and an industrial algal platform. As a result, more than 4000 Raman spectra were generated and analysed by statistical methods. These spectra reflected the physiological state of the cells and demonstrate the ability of the developed sensor to monitor the physiology of microalgal cells and their intracellular molecules of interest in a complex production environment. Full article
Show Figures

Figure 1

Figure 1
<p>Development of a Raman measurement approach: from laboratory to pilot-scale applications.</p>
Full article ">Figure 2
<p>Monitoring of <span class="html-italic">Parachlorella kessleri</span> in Bold Basal Medium in a 1-L airlift photobioreactor. (<b>A</b>) Overview of 1793 spectra obtained over 36 days. (<b>B</b>) Median of 50 spectra recorded on day 0, day 4, day 8 and day 32. (<b>C</b>) Concentrations of chlorophyll <span class="html-italic">a</span>, chlorophyll <span class="html-italic">b</span> and carotenoids increase during cell growth until nitrogen starvation. Pigment concentration decreases when NaNO<sub>3</sub> = 0 mg/L.</p>
Full article ">Figure 3
<p>(<b>A</b>) 2D correlation map of all spectra with a correlation table covering the 36 days of culture in a 1-L airlift photobioreactor. The colour of each map point represents the level of correlation between two spectra, from red (highest correlation) to blue (lowest correlation). (<b>B</b>) Repeatability of spectra measured by the autocorrelation level between 50 spectra recorded in the same time window. (<b>C</b>) Similarity of the spectra to those of the first day assessed by calculating the correlation between them.</p>
Full article ">Figure 4
<p>Monitoring of <span class="html-italic">Parachlorella kessleri</span> in Bold Basal Medium in a 100-L tubular airlift photobioreactor from day 0 to day 13. (<b>A</b>) Overview of 2720 spectra obtained at 13 days. (<b>B</b>) Median of 50 spectra recorded on day 0, day 4, day 8 and day 13 (<b>C</b>) Concentrations of chlorophyll <span class="html-italic">a</span>, chlorophyll <span class="html-italic">b</span> and carotenoids increase during cell growth until nitrogen limitation. After significant nitrogen limitation, the concentration of pigments decreases. (<b>D</b>) Lipid analyses of samples collected between day 7 and day 13. Abbreviations indicate: NL, neutral lipids; GL, glycolipids; PL, phospholipids.</p>
Full article ">Figure 5
<p>Raman band intensity over 14 days at pilot scale in a 100-L photobioreactor: v(C=C)<sub>1660cm<sup>−1</sup></sub>/δ(CH<sub>2</sub>)<sub>1444cm<sup>−1</sup></sub>, v<sub>as</sub>(C−H<sub>2</sub>)<sub>2940cm<sup>−1</sup></sub>, v<sub>as</sub> (=C−H)<sub>3008cm<sup>−1</sup></sub>, v(C=O)<sub>1750cm<sup>−1</sup></sub>, v(C=C)<sub>1524cm<sup>−1</sup></sub> and v(C−C)<sub>1157cm<sup>−1</sup></sub>/v(C=C) <sub>1524cm<sup>−1</sup></sub>.</p>
Full article ">Figure 6
<p>(<b>A</b>) Three-dimensional representation of the principal component analysis (PCA) of spectra from a 14-day culture in a 100-L tubular airlift bioreactor (PC1 26.3%; PC2 11.4%; PC4 7.4%) and their three respective loadings. (<b>B</b>) Kruskal–Wallis one-way ANOVA test, based on PCA loading 1, representing the variance of the spectra over the 13 days.</p>
Full article ">
28 pages, 8900 KiB  
Review
Optical Properties of Laccases and Their Use for Phenolic Compound Detection and Quantification: A Brief Review
by Pauline Conigliaro, Marianna Portaccio, Maria Lepore and Ines Delfino
Appl. Sci. 2023, 13(23), 12929; https://doi.org/10.3390/app132312929 - 3 Dec 2023
Cited by 1 | Viewed by 1413
Abstract
Phenolic compounds (PheCs) are particularly relevant in many different frameworks due to their pro-oxidant and antioxidant activities. In fact, on the one hand, they are considered very dangerous pro-oxidant agents that can be present in the environment as pollutants in wastewater and soil [...] Read more.
Phenolic compounds (PheCs) are particularly relevant in many different frameworks due to their pro-oxidant and antioxidant activities. In fact, on the one hand, they are considered very dangerous pro-oxidant agents that can be present in the environment as pollutants in wastewater and soil from different industrial and agricultural industries. On the other hand, the antioxidant influence of PheCs available in natural products (including foods) is nowadays considered essential for preserving human health. Conventional techniques for detecting PheCs present some disadvantages, such as requiring expensive instrumentation and expert users and not allowing in situ measurements. This is the reason why there is a high interest in the development of simple, sensitive, specific, and accurate sensing methods for PheCs. Enzymes are often used for this purpose, and laccases with unique optical properties are adopted as bio-elements for sensing schemes. The present paper aims to revise the optical properties of laccases and their use for developing PheC detection and quantification methods used in different fields such as environment monitoring, food characterization and medical applications. In particular, the results offered by UV, visible and infrared absorption, fluorescence, Raman, and surface-enhanced Raman spectroscopy (SERS) have been considered. The enzymatic biosensing devices developed using the related optical signals have been reported, and a comparison of their performances has carried out. A brief description of the main characteristics of laccase and phenols is also given. Full article
Show Figures

Figure 1

Figure 1
<p>The redox cycle for substrate oxidation is catalyzed by laccase.</p>
Full article ">Figure 2
<p>Representation of the laccase structure of <span class="html-italic">Trametes versicolor</span>; copper atoms (T1, T2 and T3) are highlighted in brown (reprinted from Ref. [<a href="#B44-applsci-13-12929" class="html-bibr">44</a>] under open access conditions).</p>
Full article ">Figure 3
<p>Absorption spectrum of <span class="html-italic">Coriolus hirsutus</span> laccase; investigated samples were prepared by dissolving 0.7 mg of the enzyme in 20 mM sodium acetate buffer pH 4.5 (reprinted with permission from Ref. [<a href="#B62-applsci-13-12929" class="html-bibr">62</a>]).</p>
Full article ">Figure 4
<p>Absorption spectra of (a) <span class="html-italic">T. ochracea</span>, (b) <span class="html-italic">T. hirsuta</span>, (c) <span class="html-italic">Coriolopsis fulvocinerea</span> and (d) <span class="html-italic">Cerrena maxima</span> laccases (1 mg/mL in phosphate buffer pH 6.0) (reprinted with permission from Ref. [<a href="#B64-applsci-13-12929" class="html-bibr">64</a>]).</p>
Full article ">Figure 5
<p>Fluorescence spectra of laccase from fungus <span class="html-italic">Sclerotinia sclerotiorum</span>. The emission spectrum shows a visible signal with the maximum located at 440 nm, while the excitation spectrum shows a peak located at 280 nm and a band centered at 330 nm. The laccase solutions were prepared using 2-(N-Morpholino) ethanesulfonic acid 4-Morpholineethanesulfonic acid (MES) buffer (reprinted with permission from Ref. [<a href="#B68-applsci-13-12929" class="html-bibr">68</a>]).</p>
Full article ">Figure 6
<p>Emission spectra (I: with an excitation wavelength of 330 nm) and excitation spectra (II: with an emission wavelength of 420 nm) of laccases from (a) <span class="html-italic">T. ochracea</span>, (b) <span class="html-italic">T. hirsuta</span>, (c) <span class="html-italic">Coriolopsis fulvocinerea</span> and (d) <span class="html-italic">Cerrena maxima</span> (reprinted with permission from Ref. [<a href="#B64-applsci-13-12929" class="html-bibr">64</a>]).</p>
Full article ">Figure 7
<p>FT-IR spectra of yellow laccase (red line), blue laccase (blue line) and white laccase (black line) (reprinted from Ref. [<a href="#B53-applsci-13-12929" class="html-bibr">53</a>] under open access conditions).</p>
Full article ">Figure 8
<p>Resonant Raman spectra of laccase from <span class="html-italic">Rhus vernicifera</span> in H<sub>2</sub>O at room temperature (300 K upper spectrum) and in the frozen state (200 K lower spectrum) (reprinted from Ref. [<a href="#B83-applsci-13-12929" class="html-bibr">83</a>] under PMC open access conditions).</p>
Full article ">Figure 9
<p>Resonant Raman spectrum of <span class="html-italic">Pleurotus ostreatus</span> laccase obtained with an excitation of 633 nm (reprinted with permission from Ref. [<a href="#B65-applsci-13-12929" class="html-bibr">65</a>]).</p>
Full article ">Figure 10
<p>To the left, the absorption spectrum of the oxidized form of syringaldazine with laccase is reported. To the right, the SERS spectrum with laccase adsorbed on Ag (a) and (b) the spectrum after the addition of syringaldazine on Ag/laccase surface are shown (reprinted with permission from Ref. [<a href="#B93-applsci-13-12929" class="html-bibr">93</a>]).</p>
Full article ">Figure 11
<p>Structure of phenol.</p>
Full article ">Figure 12
<p>Structure of some phenolic compounds: bisphenol A (BPA), catechol, resorcinol, and hydroquinone.</p>
Full article ">Figure 13
<p>Calibration curve and linear range (inset) of the optical biosensors proposed by Abdullah et al. for catechol detection (reprinted from Ref. [<a href="#B127-applsci-13-12929" class="html-bibr">127</a>] under an open access condition).</p>
Full article ">Figure 14
<p>In the left of the Figure see panels (<b>a</b>–<b>c</b>). (<b>a</b>) Absorption spectra of hydroquinone and a solution of laccase and hydroquinone; (<b>b</b>) absorption spectra of resorcinol and a solution of laccase and resorcinol; (<b>c</b>) absorption spectra of catechol and a solution of laccase and catechol. The (<b>a</b>,<b>b</b>) panels on the right show the calibration curve of the optical biosensors based on sol–gel immobilized laccase at increasing concentrations of resorcinol and catechol (reprinted from Ref. [<a href="#B131-applsci-13-12929" class="html-bibr">131</a>] under open access conditions).</p>
Full article ">Figure 15
<p>The linear range for an LTCC sensor exploiting the changes in absorbance due to solutions with different concentrations of ABTS (measurement conditions: flow rate = 25 µL/min, LED current: 15 mA) (reprinted with permission from Ref. [<a href="#B133-applsci-13-12929" class="html-bibr">133</a>]).</p>
Full article ">Figure 16
<p>Calibration plot for catechol concentration determination. In the inset, the linear range is shown (reprinted with permission from Ref. [<a href="#B136-applsci-13-12929" class="html-bibr">136</a>] under open access conditions).</p>
Full article ">Figure 17
<p>Schematic drawing of Papkovsky’s enzyme sensor. Modified from Ref. [<a href="#B138-applsci-13-12929" class="html-bibr">138</a>].</p>
Full article ">Figure 18
<p>Schematic of the instrumental system: S1 and S2, driving syringes containing laccase + AuNPs and indocyanine green + polyphenolic compound, respectively; S3, stopping syringe; FD, spectrofluorometer; C, observation cell; T, thermostat; PC, computer (reprinted with permission from Ref. [<a href="#B139-applsci-13-12929" class="html-bibr">139</a>]).</p>
Full article ">Figure 19
<p>Kinetic curves have been acquired for the laccase–indocyanine green system alone (1) and in the presence of AuNPs (2), gallic acid (3) and AuNPs + gallic acid (4). [laccase] = 0.1 U mL<sup>−1</sup>, [indocyanine green] = 5.2 µmol L<sup>−1</sup>, [gallic acid] = 2 µmol L<sup>−1</sup>, temperature = 20 °C, pH 7.5 and [Tris] = 25 mmol L<sup>−1</sup> (reprinted with permission from Ref. [<a href="#B139-applsci-13-12929" class="html-bibr">139</a>]).</p>
Full article ">Figure 20
<p>Sensing scheme for detection of polyphenols based on “turn-off” photoluminescence using enzyme immobilized CdTe QDs (reproduced with permission from Ref. [<a href="#B140-applsci-13-12929" class="html-bibr">140</a>]).</p>
Full article ">Figure 21
<p>Sensing scheme of the oxygen-detection-based optic system (reproduced from Ref. [<a href="#B142-applsci-13-12929" class="html-bibr">142</a>] under open access conditions).</p>
Full article ">Figure 22
<p>Calibration curve for catechol (reproduced from Ref. [<a href="#B142-applsci-13-12929" class="html-bibr">142</a>] under open access conditions).</p>
Full article ">
25 pages, 32152 KiB  
Article
Assessing Planet Nanosatellite Sensors for Ocean Color Usage
by Mark D. Lewis, Brittney Jarreau, Jason Jolliff, Sherwin Ladner, Timothy A. Lawson, Sean McCarthy, Paul Martinolich and Marcos Montes
Remote Sens. 2023, 15(22), 5359; https://doi.org/10.3390/rs15225359 - 14 Nov 2023
Cited by 1 | Viewed by 1414
Abstract
An increasing number of commercial nanosatellite-based Earth-observing sensors are providing high-resolution images for much of the coastal ocean region. Traditionally, to improve the accuracy of normalized water-leaving radiance (nLw) estimates, sensor gains are computed using in-orbit vicarious calibration methods. [...] Read more.
An increasing number of commercial nanosatellite-based Earth-observing sensors are providing high-resolution images for much of the coastal ocean region. Traditionally, to improve the accuracy of normalized water-leaving radiance (nLw) estimates, sensor gains are computed using in-orbit vicarious calibration methods. The initial series of Planet nanosatellite sensors were primarily designed for land applications and are missing a second near-infrared band, which is typically used in selecting aerosol models for atmospheric correction over oceanographic regions. This study focuses on the vicarious calibration of Planet sensors and the duplication of its red band for use in both the aerosol model selection process and as input to bio-optical ocean product algorithms. Error measurements show the calibration performed well at the Marine Optical Buoy location near Lanai, Hawaii. Further validation was performed using in situ data from the Aerosol Robotic Network—Ocean Color platform in the northern Adriatic Sea. Bio-optical ocean color products were generated and compared with products from the Visual Infrared Imaging Radiometric Suite sensor. This approach for sensor gain generation and usage proved effective in increasing the accuracy of nLw measurements for bio-optical ocean product algorithms. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>True-color RGB Dove scene of Venetian Lagoon on 6 December 2017.</p>
Full article ">Figure 2
<p>Dove, VIIRS and AERONET AAOT <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> for (<b>a</b>) 21 April 2017, (<b>b</b>) 27 November 2017, (<b>c</b>) 6 December 2017, (<b>d</b>) 19 December 2017, and (<b>e</b>) 21 December 2017.</p>
Full article ">Figure 3
<p>Ratio graphs between satellite sensors and the in situ sensor for Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at (<b>a</b>) 494 nm, (<b>b</b>) 545 nm, and (<b>c</b>) 635 nm.</p>
Full article ">Figure 4
<p>Chlorophyll graphs of (<b>a</b>) estimated values and (<b>b</b>) ratio of Dove and VIIRS to AERONET AAOT chlorophyll.</p>
Full article ">Figure 5
<p>VIIRS <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 486 nm, unity Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 494 nm, and calibrated Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 494 nm on 6 December 2017.</p>
Full article ">Figure 6
<p>VIIRS <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 486 nm, unity Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 494 nm, and calibrated Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 494 nm on 21 December 2017.</p>
Full article ">Figure 7
<p>VIIRS <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 551 nm, unity Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 545 nm, and calibrated Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 545 nm on 6 December 2017.</p>
Full article ">Figure 8
<p>VIIRS <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 551 nm, unity Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 545 nm, and calibrated Dove <math display="inline"><semantics> <mrow> <mi>n</mi> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>w</mi> </mrow> </msub> </mrow> </semantics></math> at 545 nm on 21 December 2017.</p>
Full article ">Figure 9
<p>VIIRS chlorophyll, unity Dove chlorophyll, and calibrated Dove chlorophyll on 6 December 2017.</p>
Full article ">Figure 10
<p>VIIRS chlorophyll, unity Dove chlorophyll, and calibrated Dove chlorophyll on 21 December 2017.</p>
Full article ">Figure 11
<p>VIIRS diffuse attenuation at 551 nm, unity Dove, and calibrated Dove diffuse attenuation at 545 nm on 6 December 2017.</p>
Full article ">Figure 12
<p>VIIRS diffuse attenuation at 551 nm, unity Dove, and calibrated Dove diffuse attenuation at 545 nm on 21 December 2017.</p>
Full article ">Figure 13
<p>VIIRS absorption at 551 nm, unity Dove, and calibrated Dove absorption at 545 nm on 6 December 2017.</p>
Full article ">Figure 14
<p>VIIRS absorption at 551 nm, unity Dove, and calibrated Dove absorption at 545 nm on 21 December 2017.</p>
Full article ">Figure 15
<p>VIIRS backscatter at 551 nm, unity Dove, and calibrated Dove backscatter at 545 nm on 6 December 2017.</p>
Full article ">Figure 16
<p>VIIRS backscatter at 551 nm, unity Dove, and calibrated Dove backscatter at 545 nm on 21 December 2017.</p>
Full article ">Figure 17
<p>Scatter plots of VIIRS and calibrated Dove on 21 December 2017 for (<b>a</b>) chlorophyll, (<b>b</b>) Kd, (<b>c</b>) absorption, and (<b>d</b>) backscatter.</p>
Full article ">
36 pages, 8597 KiB  
Review
Recent Advances in Dietary Sources, Health Benefits, Emerging Encapsulation Methods, Food Fortification, and New Sensor-Based Monitoring of Vitamin B12: A Critical Review
by Seyed Mohammad Taghi Gharibzahedi, Maryam Moghadam, Jonas Amft, Aysu Tolun, Gauri Hasabnis and Zeynep Altintas
Molecules 2023, 28(22), 7469; https://doi.org/10.3390/molecules28227469 - 7 Nov 2023
Cited by 9 | Viewed by 3996
Abstract
In this overview, the latest achievements in dietary origins, absorption mechanism, bioavailability assay, health advantages, cutting-edge encapsulation techniques, fortification approaches, and innovative highly sensitive sensor-based detection methods of vitamin B12 (VB12) were addressed. The cobalt-centered vitamin B is mainly found [...] Read more.
In this overview, the latest achievements in dietary origins, absorption mechanism, bioavailability assay, health advantages, cutting-edge encapsulation techniques, fortification approaches, and innovative highly sensitive sensor-based detection methods of vitamin B12 (VB12) were addressed. The cobalt-centered vitamin B is mainly found in animal products, posing challenges for strict vegetarians and vegans. Its bioavailability is highly influenced by intrinsic factor, absorption in the ileum, and liver reabsorption. VB12 mainly contributes to blood cell synthesis, cognitive function, and cardiovascular health, and potentially reduces anemia and optic neuropathy. Microencapsulation techniques improve the stability and controlled release of VB12. Co-microencapsulation of VB12 with other vitamins and bioactive compounds enhances bioavailability and controlled release, providing versatile initiatives for improving bio-functionality. Nanotechnology, including nanovesicles, nanoemulsions, and nanoparticles can enhance the delivery, stability, and bioavailability of VB12 in diverse applications, ranging from antimicrobial agents to skincare and oral insulin delivery. Staple food fortification with encapsulated and free VB12 emerges as a prominent strategy to combat deficiency and promote nutritional value. Biosensing technologies, such as electrochemical and optical biosensors, offer rapid, portable, and sensitive VB12 assessment. Carbon dot-based fluorescent nanosensors, nanocluster-based fluorescent probes, and electrochemical sensors show promise for precise detection, especially in pharmaceutical and biomedical applications. Full article
(This article belongs to the Special Issue Current Emerging Trends of Extraction and Encapsulation in Food)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of VB<sub>12.</sub> (Co<sup>+</sup> is central cobalt ion linked to the upper ligand (R).</p>
Full article ">Figure 2
<p>Scanning electron microscopy (SEM; magnification of 5000×, scale bar of 20 μm) images of electrospun and electrosprayed VB<sub>12</sub>-loaded zein microstructures (70% ethanol, 10–20% zein (Z), and 1–10% VB<sub>12</sub>) prepared by electrospinning and spray-drying techniques under different operating and formulation conditions: (<b>A</b>), 10 Z:1 VB<sub>12</sub>; (<b>B</b>), 10 Z:5 VB<sub>12</sub>; (<b>C</b>), 10 Z:10 VB<sub>12</sub>; 0.3 mL/h of flow rate, 7 cm distance; 20 Z:5 VB<sub>12</sub>; 0.2 mL/h of flow rate with 7 cm (<b>D</b>), 10 cm (<b>E</b>), and 15 cm (<b>F</b>) distances). Reprinted with permission from [<a href="#B90-molecules-28-07469" class="html-bibr">90</a>].</p>
Full article ">Figure 3
<p>SEM images (magnification of 5000×, scale bar of 20 μm) of 20% zein (Z)-based microcapsules loaded with VB<sub>12</sub> (1–10%) and prepared by spray-drying: (<b>A</b>), Z20:1 VB<sub>12</sub>; (<b>B</b>), Z20:5 VB<sub>12</sub>; (<b>C</b>), Z20:10 VB<sub>12</sub>). Reprinted with permission from [<a href="#B90-molecules-28-07469" class="html-bibr">90</a>].</p>
Full article ">Figure 4
<p>In vitro release profiles of EGCG/VB<sub>12</sub>/(EGCG + VB<sub>12</sub>) normalized by the total amount released, in water of the electrospun zein (30% <span class="html-italic">w</span>/<span class="html-italic">v</span>) microstructures loaded with active compounds (0.5, 1, and 5% <span class="html-italic">w</span>/<span class="html-italic">w</span>). Reprinted from [<a href="#B83-molecules-28-07469" class="html-bibr">83</a>].</p>
Full article ">Figure 5
<p>The release profile of VB<sub>12</sub> at pH 7.46 (<b>a</b>), doxorubicin at pH 7.46 (<b>b</b>), and doxorubicin at pH 4.5 (<b>c</b>). The pictorial representation of release of VB<sub>12</sub> (<b>d</b>) and doxorubicin (<b>e</b>). Reprinted with permission from [<a href="#B102-molecules-28-07469" class="html-bibr">102</a>].</p>
Full article ">Figure 6
<p>Representative images of lyophilized VB<sub>12</sub> lipid vesicles: (<b>a</b>) empty liposome with lactose, (<b>b</b>) empty liposome with sorbitol, (<b>c</b>) VB<sub>12</sub>-loaded liposome with lactose, (<b>d</b>) VB<sub>12</sub>-loaded liposome with sorbitol, (<b>e</b>) empty transfersome with lactose, (<b>f</b>) empty transfersome with sorbitol, (<b>g</b>) VB<sub>12</sub>-loaded transfersome with lactose, and (<b>h</b>) VB<sub>12</sub>-loaded transfersome with sorbitol. Reprinted from [<a href="#B107-molecules-28-07469" class="html-bibr">107</a>].</p>
Full article ">Figure 7
<p>Schematic images of some VB<sub>12</sub>-nanoencapsulating systems including niosome (<b>A</b>), liposome (<b>B</b>), O/W emulsion (<b>C</b>), W/O/W emulsion (<b>D</b>), and SLNs (<b>E</b>). Reprinted from [<a href="#B131-molecules-28-07469" class="html-bibr">131</a>,<a href="#B132-molecules-28-07469" class="html-bibr">132</a>].</p>
Full article ">Figure 8
<p>The limit of detection (LOD) and linear range of VB<sub>12</sub> detected by some electrochemical sensors (see <a href="#molecules-28-07469-t003" class="html-table">Table 3</a> for more information) [<a href="#B173-molecules-28-07469" class="html-bibr">173</a>,<a href="#B174-molecules-28-07469" class="html-bibr">174</a>,<a href="#B175-molecules-28-07469" class="html-bibr">175</a>,<a href="#B176-molecules-28-07469" class="html-bibr">176</a>,<a href="#B178-molecules-28-07469" class="html-bibr">178</a>,<a href="#B179-molecules-28-07469" class="html-bibr">179</a>,<a href="#B180-molecules-28-07469" class="html-bibr">180</a>,<a href="#B181-molecules-28-07469" class="html-bibr">181</a>].</p>
Full article ">Figure 9
<p>(<b>A</b>) Cyclic voltammograms recorded during the growth process of the PTH film on GCE in a nitrogen-saturated 0.1 M pH 6.0 buffer solution containing 5 mmol L<sup>−1</sup> TH and 0.1 mol L<sup>−1</sup> NaNO<sub>3</sub>, with a scan rate of 100 mV s<sup>−1</sup>. (<b>B</b>) SEM image of the GCE(ea)/PTH surface. (<b>C</b>) FTIR spectrum of the PTH film (The wavenumbers corresponding to the primary chemo-functional groups (−OH, −NH<sub>2</sub>, C=N, and −C=C−) are highlighted in red). (<b>D</b>) Cyclic voltammograms of GCE(ea)/PTH in 0.1 mol L<sup>−1</sup> pH 6.5 buffer at various scan rates (ν, color curves), ranging from a to i: from 10 to 600 mV s<sup>−1</sup>. Reprinted with permission from [<a href="#B179-molecules-28-07469" class="html-bibr">179</a>].</p>
Full article ">
Back to TopTop