[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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 (33)

Search Parameters:
Keywords = phototropism

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 2622 KiB  
Article
A Portable Photocollector for the Field Collection of Insects in Biodiversity Assessment
by Behnam Motamedinia, Sophie Cardinal, Scott Kelso, Carolyn Callaghan, Khorshid Ghahari, John F. Wilmshurst and Jeff Skevington
Insects 2024, 15(11), 896; https://doi.org/10.3390/insects15110896 - 16 Nov 2024
Viewed by 714
Abstract
Arthropod biodiversity research usually requires large sample collections. The efficient handling of these samples has always been a critical bottleneck. Sweep netting along transects is an effective and commonly used approach to sample diverse insects. However, sweep netting requires the time-consuming task of [...] Read more.
Arthropod biodiversity research usually requires large sample collections. The efficient handling of these samples has always been a critical bottleneck. Sweep netting along transects is an effective and commonly used approach to sample diverse insects. However, sweep netting requires the time-consuming task of sorting insects from the large amounts of debris and foliage that end up in the sweep net along with the insects. To address this, we introduce a robust, portable, and inexpensive photocollector device with an LED light source to extract insects from sweep net samples in a standardized way. Timed field trials tested the photocollector’s efficiency in extracting live insect samples from debris, focusing on Hymenoptera and Diptera. We found that 73% (±13%) of undamaged specimens moved toward the collection bottle within the first hour and 79% (±13%) after four hours. Of the insects failing to move after four hours, most (81%) were damaged and likely unable to move. Accounting only for undamaged specimens, 83% (±11%) moved after 1 h and 90% (±11%) moved after 4 h. We found significant differences in when families of Hymenoptera and Diptera moved. We suggest that the photocollector can be a useful tool in standardized biodiversity assessments. Full article
(This article belongs to the Section Insect Ecology, Diversity and Conservation)
Show Figures

Figure 1

Figure 1
<p>Photocollector: (<b>A1</b>) black plastic box (Ikea Uppsnofsasd™), (<b>A2</b>) wooden ramp, (<b>A3</b>) collector bottle entrance hole, (<b>A4</b>) front wall, (<b>B1</b>) collection bottle, (<b>B2</b>) threaded bottle cap (top of the cap has been cut off), (<b>B3</b>) calf feeder nipple, (<b>B4</b>) water bottle head (cut and painted). (<b>C</b>,<b>D</b>) Photocollector box cap (<b>C</b>) (dorsal view) and (<b>D</b>) (lateral view): (<b>D1</b>) wood screw, (<b>D2</b>) wall anchor, (<b>D3</b>) wooden disk, (<b>D4</b>) insulation foam stripe, (<b>D5</b>) foam board. (<b>E</b>) Gluing the threaded cap covered with calf feeder nipple to the collector bottle entrance hole: (<b>E1</b>) water bottle head (cut and painted), (<b>E2</b>) calf feeder nipple covering the threaded cap. (<b>F</b>) Interior with all components: (<b>F1</b>) guide wall foam. (<b>G</b>,<b>H</b>) Exterior: (<b>G1</b>) plastic box lid, (<b>G2</b>) collection bottle, (<b>G3</b>) LED light, (<b>G4</b>) sock; (<b>G5</b>) bungee cord, (<b>H</b>) assembled photocollector.</p>
Full article ">Figure 2
<p>Temporal distribution by order of the proportion of arthropod specimens moving into a lit collection bottle from a dark photocollector. Collections were made at hourly intervals for 4 h. Proportions include the specimens that did not move toward the light.</p>
Full article ">Figure 3
<p>Mean proportion of specimens that moved during the first hour of photocollector operation for each order. Fill patterns group insect orders into three general movement rates (red striped = fast, green solid = moderate, blue zigzag = slow). Bars with different lowercase letters are orders with statistically different proportions of individuals moving in the first hour. Analysis based on ANOVA with a post-hoc Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Proportion of specimens moving after the first hour by insect order. Black bars indicate movements in the second hour, gray bars indicate movements in the third hour, orange bars indicate movements in the fourth hour, and blue bars indicate the proportion of specimens that did not move after four hours. For most orders (Hemiptera was an exception), specimens that did not move in the first hour tended to not move at all (blue bars), but there were significant patterns among taxa in movement rates in hours 2, 3, and 4. Proportions of Hemiptera and Hymenoptera moving in hour 2 were far greater than those of Araneae and Coleoptera (Tukey’s post-hoc test: <span class="html-italic">p</span>&lt; 0.05), while a greater proportion of Orthoptera moved in hour 2 than Araneae (Tukey’s post-hoc test: <span class="html-italic">p</span> = 0.02). Orthoptera moved significantly less than Hemiptera (Tukey’s post-hoc test: <span class="html-italic">p</span> &lt; 0.002) and Hymenoptera (Tukey’s post-hoc test: <span class="html-italic">p</span> &lt; 0.02) in hour 3. During hour 4, more Hemiptera moved than Araneae, Coleoptera, Diptera, Lepidoptera, and Orthoptera (Tukey’s post-hoc test: <span class="html-italic">p</span> &lt; 0.02). Unmoved specimens were dominantly Lepidoptera (<span class="html-italic">p</span> &lt; 0.05), while a lower proportion of Coleoptera did not move compared to all orders, except for Araneae and Orthoptera (<span class="html-italic">p</span> = 0.01).</p>
Full article ">Figure 5
<p>Temporal distribution of families in the order Hymenoptera moving into a lit collection bottle from a dark photocollector. Collections were made at hourly intervals for 4 h. Proportions include the specimens that did not move toward the light after 4 h.</p>
Full article ">Figure 6
<p>Temporal distribution of Families in the Order Diptera moving into a lit collection bottle from a dark photocollector. Collections were made at hourly intervals for 4 h. Proportions include the specimens that did not move to the light after 4 h.</p>
Full article ">
26 pages, 4493 KiB  
Article
Trajectory Optimization to Enhance Observability for Bearing-Only Target Localization and Sensor Bias Calibration
by Jicheng Peng, Qianshuai Wang, Bingyu Jin, Yong Zhang and Kelin Lu
Biomimetics 2024, 9(9), 510; https://doi.org/10.3390/biomimetics9090510 - 23 Aug 2024
Viewed by 780
Abstract
This study addresses the challenge of bearing-only target localization with sensor bias contamination. To enhance the system’s observability, inspired by plant phototropism, we propose a control barrier function (CBF)-based method for UAV motion planning. The rank criterion provides only qualitative observability results. We [...] Read more.
This study addresses the challenge of bearing-only target localization with sensor bias contamination. To enhance the system’s observability, inspired by plant phototropism, we propose a control barrier function (CBF)-based method for UAV motion planning. The rank criterion provides only qualitative observability results. We employ the condition number for a quantitative analysis, identifying key influencing factors. After that, a multi-objective, nonlinear optimization problem for UAV trajectory planning is formulated and solved using the proposed Nonlinear Constrained Multi-Objective Gray Wolf Optimization Algorithm (NCMOGWOA). Simulations validate our approach, showing a threefold reduction in the condition number, significantly enhancing observability. The algorithm outperforms others in terms of localization accuracy and convergence, achieving the lowest Generational Distance (GD) (7.3442) and Inverted Generational Distance (IGD) (8.4577) metrics. Additionally, we explore the effects of the CBF attenuation rates and initial flight path angles. Full article
Show Figures

Figure 1

Figure 1
<p>Model of the relative motion of the UAV and the target.</p>
Full article ">Figure 2
<p>Phototropism in plants, where the bulb represents the light source.</p>
Full article ">Figure 3
<p>Effect of <math display="inline"> <semantics> <mi>δ</mi></semantics></math> (<math display="inline"> <semantics> <mrow> <mn>0</mn> <mo>&lt;</mo> <msub> <mi>δ</mi> <mn>2</mn></msub> <mo>&lt;</mo> <msub> <mi>δ</mi> <mn>1</mn></msub> <mo>&lt;</mo> <mn>1</mn></mrow></semantics></math>) on system observability: (<b>a</b>) <math display="inline"> <semantics> <mrow> <mi>δ</mi> <mo>=</mo> <msub> <mi>δ</mi> <mn>1</mn></msub></mrow></semantics></math>; (<b>b</b>) <math display="inline"> <semantics> <mrow> <mi>δ</mi> <mo>=</mo> <msub> <mi>δ</mi> <mn>2</mn></msub></mrow></semantics></math>. The gray oval with a solid outline denotes the unobservable area. The gray oval with a dashed outline and the white oval with a solid outline represent <math display="inline"> <semantics> <mrow> <msub> <mi>S</mi> <mrow> <mo>Δ</mo> <mi>t</mi></mrow></msub></mrow></semantics></math> and <math display="inline"> <semantics> <mrow> <msub> <mi>R</mi> <mrow> <mo>Δ</mo> <mi>t</mi></mrow></msub></mrow></semantics></math>, respectively. The black dot represents <math display="inline"> <semantics> <mrow> <mi>x</mi> <mo stretchy="false">(</mo> <mi>t</mi> <mo stretchy="false">)</mo></mrow></semantics></math>.</p>
Full article ">Figure 4
<p>Heatmap of the inverse of condition number. The horizontal coordinate represents the ratio of the relative distance between the UAV and the target to the speed of the UAV. The vertical coordinate represents the angle of separation.</p>
Full article ">Figure 5
<p>The Pareto solution of the two objective optimization problems. The black dots denote the Pareto solution while the blue dots do not. The red line denotes the Pareto front.</p>
Full article ">Figure 6
<p>Process of the TOPSIS and CRITIC methods (TCM).</p>
Full article ">Figure 7
<p>Flowchart of the algorithm for biased bearing information-only target localization and UAV trajectory optimization based on observability enhancement.</p>
Full article ">Figure 8
<p>UAV trajectory optimization results for a stationary target: (<b>a</b>) trajectory of UAV; (<b>b</b>) estimated position of target; (<b>c</b>) localization error; (<b>d</b>) estimation error of bias <math display="inline"> <semantics> <mi>b</mi></semantics></math>; (<b>e</b>) speed of UAV; (<b>f</b>) turn rate of UAV; (<b>g</b>) bearing and flight path angle; (<b>h</b>) separation angle; (<b>i</b>) attenuation rate of CBF; (<b>j</b>) relative distance; (<b>k</b>) rank of observability matrix; and (<b>l</b>) condition number.</p>
Full article ">Figure 8 Cont.
<p>UAV trajectory optimization results for a stationary target: (<b>a</b>) trajectory of UAV; (<b>b</b>) estimated position of target; (<b>c</b>) localization error; (<b>d</b>) estimation error of bias <math display="inline"> <semantics> <mi>b</mi></semantics></math>; (<b>e</b>) speed of UAV; (<b>f</b>) turn rate of UAV; (<b>g</b>) bearing and flight path angle; (<b>h</b>) separation angle; (<b>i</b>) attenuation rate of CBF; (<b>j</b>) relative distance; (<b>k</b>) rank of observability matrix; and (<b>l</b>) condition number.</p>
Full article ">Figure 9
<p>UAV trajectory optimization results between the open-loop and closed-loop mode: (<b>a</b>) estimated location of the target; (<b>b</b>) condition number; (<b>c</b>) localization error; and (<b>d</b>) estimation error of bias <math display="inline"> <semantics> <mi>b</mi></semantics></math>.</p>
Full article ">Figure 10
<p>Trajectory optimization results for different <math display="inline"> <semantics> <mi>δ</mi></semantics></math> values: (<b>a</b>) trajectory of UAV; (<b>b</b>) relative distance; (<b>c</b>) localization error; and (<b>d</b>) condition number.</p>
Full article ">Figure 10 Cont.
<p>Trajectory optimization results for different <math display="inline"> <semantics> <mi>δ</mi></semantics></math> values: (<b>a</b>) trajectory of UAV; (<b>b</b>) relative distance; (<b>c</b>) localization error; and (<b>d</b>) condition number.</p>
Full article ">Figure 11
<p>UAV trajectory optimization for different initial flight path angles: (<b>a</b>) trajectory of the UAV; (<b>b</b>) separation angle.</p>
Full article ">Figure 12
<p>State estimation errors between the NCMOGWOA, MOPSOA, MOAOA, and SQP Methods: (<b>a</b>) localization errors; (<b>b</b>) estimation error of <math display="inline"> <semantics> <mi>b</mi></semantics></math>.</p>
Full article ">Figure 13
<p>Comparison of the Pareto optimal solution sets over 50 runs: (<b>a</b>) t = 10 s; (<b>b</b>) t = 50 s.</p>
Full article ">
17 pages, 8343 KiB  
Article
Ultrastructure and Spectral Characteristics of the Compound Eye of Asiophrida xanthospilota (Baly, 1881) (Coleoptera, Chrysomelidae)
by Zu-Long Liang, Tian-Hao Zhang, Jacob Muinde, Wei-Li Fan, Ze-Qun Dong, Feng-Ming Wu, Zheng-Zhong Huang and Si-Qin Ge
Insects 2024, 15(7), 532; https://doi.org/10.3390/insects15070532 - 13 Jul 2024
Viewed by 1120
Abstract
In this study, the morphology and ultrastructure of the compound eye of Asi. xanthospilota were examined by using scanning electron microscopy (SEM), transmission electron microscopy (TEM), micro-computed tomography (μCT), and 3D reconstruction. Spectral sensitivity was investigated by electroretinogram (ERG) tests and phototropism experiments. [...] Read more.
In this study, the morphology and ultrastructure of the compound eye of Asi. xanthospilota were examined by using scanning electron microscopy (SEM), transmission electron microscopy (TEM), micro-computed tomography (μCT), and 3D reconstruction. Spectral sensitivity was investigated by electroretinogram (ERG) tests and phototropism experiments. The compound eye of Asi. xanthospilota is of the apposition type, consisting of 611.00 ± 17.53 ommatidia in males and 634.8 0 ± 24.73 ommatidia in females. Each ommatidium is composed of a subplano-convex cornea, an acone consisting of four cone cells, eight retinular cells along with the rhabdom, two primary pigment cells, and about 23 secondary pigment cells. The open type of rhabdom in Asi. xanthospilota consists of six peripheral rhabdomeres contributed by the six peripheral retinular cells (R1~R6) and two distally attached rhabdomeric segments generated solely by R7, while R8 do not contribute to the rhabdom. The orientation of microvilli indicates that Asi. xanthospilota is unlikely to be a polarization-sensitive species. ERG testing showed that both males and females reacted to stimuli from red, yellow, green, blue, and ultraviolet light. Both males and females exhibited strong responses to blue and green light but weak responses to red light. The phototropism experiments showed that both males and females exhibited positive phototaxis to all five lights, with blue light significantly stronger than the others. Full article
(This article belongs to the Section Insect Systematics, Phylogeny and Evolution)
Show Figures

Figure 1

Figure 1
<p>Adult <span class="html-italic">Asi. xanthospilota</span> are mating on host plant <span class="html-italic">Cotinus coggygria</span>. Photo taken by Dr. Zhengzhong Huang.</p>
Full article ">Figure 2
<p>Schematic drawing to illustrate how the local eye radius (<span class="html-italic">r</span>) and interommatidial angle (Δ<span class="html-italic">φ</span>) were determined. <span class="html-italic">s</span>: length of the baseline of the eye; <span class="html-italic">h</span>: the longest distance from the curvature to the baseline; and <span class="html-italic">d</span>: the facet diameter of an ommatidium.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>f</b>) Scanning electron microscopy (SEM) of <span class="html-italic">Asi. xanthospilota</span>. (<b>a</b>) head of female; (<b>b</b>) head of male; (<b>c</b>) compound eye of female; (<b>d</b>) compound eye of male; (<b>e</b>) hexagonal facets; (<b>f</b>) pentagonal facets. (<b>g</b>–<b>i</b>) Measurement of the external morphology of compound eyes. (<b>g</b>) Number of facets of male and female (n = 10); (<b>h</b>) Area of compound eye of male and female (n = 10); (<b>i</b>) Area of each hexagonal and pentagonal facet (n = 20). The red arrow in (<b>e</b>) indicates the interfacetal hair; the red areas in (<b>c</b>,<b>d</b>) indicate parts of the area of hexagonal facets; the green areas indicate parts of the pentagonal facets. A value of <span class="html-italic">p</span> &lt; 0.05 was considered statistically significant (n.s., not significant; *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Ultrastructure of ommatidia of <span class="html-italic">Asi. xanthospilota</span>. (<b>a</b>) Semi-schematic drawings of longitudinal and cross section of an ommatidium; (<b>b</b>) TEM micrograph of longitudinal section of ommatidia; (<b>c</b>) TEM micrograph of longitudinal section of cornea; (<b>d</b>) TEM micrograph of longitudinal section of crystalline cone and primary and secondary pigment cells; (<b>e</b>) TEM micrograph of longitudinal section of retinular cells, showing the arrangement of rhabdomeres. Reference figures of the semi-schematic drawings: I—(<b>b</b>–<b>e</b>); II—<a href="#insects-15-00532-f005" class="html-fig">Figure 5</a>a,b; III—<a href="#insects-15-00532-f005" class="html-fig">Figure 5</a>c; IV—<a href="#insects-15-00532-f005" class="html-fig">Figure 5</a>d; V—<a href="#insects-15-00532-f005" class="html-fig">Figure 5</a>e; VI—<a href="#insects-15-00532-f005" class="html-fig">Figure 5</a>f; VII—<a href="#insects-15-00532-f005" class="html-fig">Figure 5</a>g,h. Abbreviations: Co—cornea; CC—crystalline cone; PPC—primary pigment cell; SPC—secondary pigment cell; Rh—rhabdom; R1–R8—retinular cells; Rh1–Rh7—rhabdomeres; BM—basal membrane; CCN—nuclei of cone cells; PCN—nuclei of primary pigment cells; RCN—nuclei of retinular cells. Scale bar: (<b>b</b>) = 20 μm; (<b>c</b>,<b>d</b>) = 10 μm; (<b>e</b>) = 2 μm.</p>
Full article ">Figure 5
<p>TEM micrographs of cross section at different levels of the compound eye of <span class="html-italic">Asi. xanthospilota</span>. (<b>a</b>) cross section of cornea and crystalline cone; (<b>b</b>) cross section of crystalline cone; (<b>c</b>) cross section of crystalline cone and primary pigment cells; (<b>d</b>) cross section of distal central rhabdom and primary pigment cells; (<b>e</b>) cross section of retinular cells, showing the arrangement of rhabdomeres; (<b>f</b>) cross section of central retinular cells, showing the two segments of central rhabdomere attributing by R7 and orientation of microvilli; (<b>g</b>) cross section of retinular cells, showing the central rhabdom and nuclei of peripheral retinular cells; (<b>h</b>) cross section of central retinular cells, showing the two segments of rhabdomeres generated by R7; (<b>i</b>) cross section of the proximal region above and below the basal membrane, showing the arrangements of axons of retinular cells, red rectangular boxes indicate the axon bundles of retinular cells above the basal membrane, green rectangular boxes indicate the axon bundles of retinular cells below the basal membrane. Abbreviations: Co—cornea; CC—crystalline cone; PPC—primary pigment cell; SPC—secondary pigment cell; RCN—nuclei of retinular cells; Rh—rhabdom; R1–R8—retinular cells; Rh1–Rh7—rhabdomeres. Scale bar: (<b>a</b>) = 10 μm; (<b>b</b>), (<b>i</b>) = 5 μm; (<b>c</b>–<b>e</b>), (<b>g</b>) = 2 μm; (<b>f</b>,<b>h</b>) = 1 μm.</p>
Full article ">Figure 6
<p>3D reconstruction of the head of <span class="html-italic">Asi. xanthospilota</span>. (<b>a</b>) dorsal view of head; (<b>b</b>) frontal view of head; (<b>c</b>) lateral view of head; (<b>d</b>) ventral view of head; (<b>e</b>) posterior view of head; (<b>f</b>) posterior view of compound eyes and brain; (<b>g</b>) dorsal view of compound eyes and brain; (<b>h</b>) ventral view of compound eyes and brain. Scale bar = 500 μm.</p>
Full article ">Figure 7
<p>Electrophysiological waveforms of female and male <span class="html-italic">Asi. xanthospilota</span> to different light stimuli (red, yellow, green, blue, and ultraviolet from <b>top</b> to <b>bottom</b>). The waveforms have been denoised using R.</p>
Full article ">Figure 8
<p>Quantification of ERG voltage responses of female and male <span class="html-italic">Asi. xanthospilota</span> exposed to different light stimuli (n = 9). (<b>a</b>) female; (<b>b</b>) male. Boxplots not sharing the same Greek letter are significantly different at <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">Figure 9
<p>Phototaxis responses of female and male <span class="html-italic">Asi. xanthospilota</span> to different light stimuli (n = 9). (<b>a</b>) phototaxis response of female different light stimuli; (<b>b</b>) phototaxis response of male different light stimuli; (<b>c</b>) positive phototaxis rate of female different light stimuli; (<b>d</b>) positive phototaxis rate of male different light stimuli. A value of <span class="html-italic">p</span> &lt; 0.05 was considered statistically significant (n.s., not significant; ***, <span class="html-italic">p</span> &lt; 0.001). Boxplots not sharing the same Greek letter are significantly different at <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">
24 pages, 16539 KiB  
Review
What We Are Learning from the Diverse Structures of the Homodimeric Type I Reaction Center-Photosystems of Anoxygenic Phototropic Bacteria
by Robert A. Niederman
Biomolecules 2024, 14(3), 311; https://doi.org/10.3390/biom14030311 - 6 Mar 2024
Viewed by 1699
Abstract
A Type I reaction center (RC) (Fe-S type, ferredoxin reducing) is found in several phyla containing anoxygenic phototrophic bacteria. These include the heliobacteria (HB), the green sulfur bacteria (GSB), and the chloracidobacteria (CB), for which high-resolution homodimeric RC-photosystem (PS) structures have recently appeared. [...] Read more.
A Type I reaction center (RC) (Fe-S type, ferredoxin reducing) is found in several phyla containing anoxygenic phototrophic bacteria. These include the heliobacteria (HB), the green sulfur bacteria (GSB), and the chloracidobacteria (CB), for which high-resolution homodimeric RC-photosystem (PS) structures have recently appeared. The 2.2-Å X-ray structure of the RC-PS of Heliomicrobium modesticaldum revealed that the core PshA apoprotein (PshA-1 and PshA-2 homodimeric pair) exhibits a structurally conserved PSI arrangement comprising five C-terminal transmembrane α-helices (TMHs) forming the RC domain and six N-terminal TMHs coordinating the light-harvesting (LH) pigments. The Hmi. modesticaldum structure lacked quinone molecules, indicating that electrons were transferred directly from the A0 (81-OH-chlorophyll (Chl) a) acceptor to the FX [4Fe-4S] component, serving as the terminal RC acceptor. A pair of additional TMHs designated as Psh X were also found that function as a low-energy antenna. The 2.5-Å resolution cryo-electron microscopy (cryo-EM) structure for the RC-PS of the green sulfur bacterium Chlorobaculum tepidum included a pair of Fenna–Matthews–Olson protein (FMO) antennae, which transfer excitations from the chlorosomes to the RC-PS (PscA-1 and PscA-2) core. A pair of cytochromes cZ (PscC) molecules was also revealed, acting as electron donors to the RC bacteriochlorophyll (BChl) a’ special pair, as well as PscB, housing the [4Fe-4S] cluster FA and FB, and the associated PscD protein. While the FMO components were missing from the 2.6-Å cryo-EM structure of the Zn- (BChl) a’ special pair containing RC-PS of Chloracidobacterium thermophilum, a unique architecture was revealed that besides the (PscA)2 core, consisted of seven additional subunits including PscZ in place of PscD, the PscX and PscY cytochrome c serial electron donors and four low mol. wt. subunits of unknown function. Overall, these diverse structures have revealed that (i) the HB RC-PS is the simplest light–energy transducing complex yet isolated and represents the closest known homolog to a common homodimeric RC-PS ancestor; (ii) the symmetrically localized Ca2+-binding sites found in each of the Type I homodimeric RC-PS structures likely gave rise to the analogously positioned Mn4CaO5 cluster of the PSII RC and the TyrZ RC donor site; (iii) a close relationship between the GSB RC-PS and the PSII Chl proteins (CP)43 and CP47 was demonstrated by their strongly conserved LH-(B)Chl localizations; (iv) LH-BChls of the GSB-RC-PS are also localized in the conserved RC-associated positions of the PSII ChlZ-D1 and ChlZ-D2 sites; (v) glycosylated carotenoids of the GSB RC-PS are located in the homologous carotenoid-containing positions of PSII, reflecting an O2-tolerance mechanism capable of sustaining early stages in the evolution of oxygenic photosynthesis. In addition to the close relationships found between the homodimeric RC-PS and PSII, duplication of the gene encoding the ancestral Type I RC apoprotein, followed by genetic divergence, may well account for the appearance of the heterodimeric Type I and Type II RCs of the extant oxygenic phototrophs. Accordingly, the long-held view that PSII arose from the anoxygenic Type II RC is now found to be contrary to the new evidence provided by Type I RC-PS homodimer structures, indicating that the evolutionary origins of anoxygenic Type II RCs, along with their distinct antenna rings are likely to have been preceded by the events that gave rise to their oxygenic counterparts. Full article
(This article belongs to the Collection Feature Papers in Molecular Structure and Dynamics)
Show Figures

Figure 1

Figure 1
<p>Structure of the Type I homodimeric reaction center-photosystem (RC-PS) complex of <span class="html-italic">Hmi. modesticaldum</span> as determined by X-ray crystallography at 2.2 Å resolution [<a href="#B3-biomolecules-14-00311" class="html-bibr">3</a>] (Protein Data Bank (PDB) ID: 5V8K). View parallel (<b>A</b>) and perpendicular (<b>B</b>) to membrane plane, later showing the cytoplasmic surface: PshA-1, red; PshA-2, pink; PshX, orange; cofactors (electron transfer), teal; antenna, blue; carotenoids, lime; [4Fe-4S], red and yellow spheres. Bacteriochlorophyll (BChl) and chlorophyll (Chl) tails are not shown. (<b>C</b>,<b>D</b>) Respective cytoplasmic views of transmembrane α-helices (TMH) and cofactor arrangements. PshA helices are numbered 1–11, with transparent gray helices corresponding to heterodimeric PSI helical arrangement and with [4Fe-4S] F<sub>X</sub> component at the center. The PS pigments in panel D are superimposed and shown upon the gray PsaA-PsaB heterodimeric core-associated PSI cofactors. Heliobacterial (HB) electron transfer BChls and Chls, brown; bulk antenna pigments, blue; BChl <span class="html-italic">g</span> molecules flanking electron transport chain, teal.</p>
Full article ">Figure 2
<p>Cofactor arrangement of the HB RC electron transfer chain and modeled menaquinone (MQ) molecule inserted into a potential quinone (Q) binding site [<a href="#B3-biomolecules-14-00311" class="html-bibr">3</a>]. (<b>A</b>) Cofactors showing coordinating residues provided by the homodimeric (PshA)<sub>2</sub>. H<sub>2</sub>O molecules (small red balls) serve as axial ligands of the A<sub>0</sub> primary acceptor, while the purple spheres liganded to the A<sub>CC</sub> accessory BChls are possible chloride ions, and the magnesium atoms at the centers of the (B)chlorin rings are represented green spheres. Note that no carotenoid was found within the RC, and only two are seen in the (PshA)<sub>2</sub> complex, a reflection of the anaerobic environment in which <span class="html-italic">Hmi. modesticaldum</span> is found, lacking reactive oxygen species. (<b>B</b>) Placement of MQ molecules in potential binding sites located with isoprenyl tails in positions similar to unassigned electron densities in the crystal structure [<a href="#B21-biomolecules-14-00311" class="html-bibr">21</a>]. (<b>C</b>) Unassigned electron density map in the vicinity of A<sub>0</sub> of the HB RC [<a href="#B3-biomolecules-14-00311" class="html-bibr">3</a>] (Supplementary Material). (<b>D</b>) Residues (light gray carbons) provided by PsaA coordinate the phylloquinone (PQ) molecule of PSI (dark gray carbons) together with the residues of PshA (dark red carbons), forming an analogous binding site for MQ [<a href="#B3-biomolecules-14-00311" class="html-bibr">3</a>] (Supplementary Material).</p>
Full article ">Figure 3
<p>Overall asymmetric structure of the complete Type I homodimeric RC-PS complex of <span class="html-italic">Cba. tepidum</span> with associated trimeric Fenna-Matthews-Olson protein (FMO), as determined by cryo-electron microscopy (cryo-EM) at 2.5 Å resolution [<a href="#B4-biomolecules-14-00311" class="html-bibr">4</a>] (PDB ID: 7Z6Q). Views showing C<sub>2</sub> symmetry of PscA components parallel (<b>A</b>) and perpendicular (<b>B</b>) to membrane plane also revealing the asymmetric arrangement of FMO components at cytoplasmic surface. The (B)Chls, carotenoids, and [4Fe-4S] cluster cofactors and lipids are all depicted in wheat. The distinct FMO-1 to FMO-3 and FMO-4-to FMO-6 components forming the respective FMO trimers are shown. (<b>C</b>) Periplasmic surface view of the arrangement of PscA and PscC helices, together with the RC-PS pigments. Helices are numbered starting from the N-terminal position: (B)Chls, in stick representation, gray; central Mg atoms, pink. (<b>D</b>) Cofactor arrangement of the <span class="html-italic">Cba. tepidum</span> RC electron transfer chain [<a href="#B3-biomolecules-14-00311" class="html-bibr">3</a>]. Residues coordinating the (B)Chl cofactors are shown at right: A<sub>0</sub>, Chl <span class="html-italic">a</span>; A<sub>CC</sub>, Chl <span class="html-italic">a</span>; and P<sub>840</sub> special pair, BChl <span class="html-italic">a</span>’; PscA-1 residues, gray; PscA-2 residues, pink.</p>
Full article ">Figure 4
<p>The common arrangement of (B)Chls in Type I and II RC complexes. (<b>A</b>,<b>B</b>) Distributions and superpositioning of (B)Chls in the green sulfur bacterial (GSB) RC-PS complex (dark green) and in PSII (yellow) shown perpendicular to the membrane plane [<a href="#B5-biomolecules-14-00311" class="html-bibr">5</a>]. Dashed circles demarcate similar clustering of the antenna (B)Chls around the RC electron transferring cofactors. (<b>C</b>,<b>D</b>) Distributions and superpositioning of (B)Chls in GSB RC-PS (dark green) with those of the HB RC-PS (purple) and PSI (blue). Dashed ellipses demarcate the closed rings of (B)Chls distributed across the two RC-PS subunits. Note that the (B)Chl distribution in the GSB RC-PS more closely resembles that of PSII than PSI.</p>
Full article ">Figure 5
<p>Potential pathways of excitation energy transfer revealed by the structure of the GSB FMO-RC-PS complex [<a href="#B4-biomolecules-14-00311" class="html-bibr">4</a>]. (<b>A</b>) Interface region between FMO-1 and PscA-1 subunit at cytoplasmic pigment surface. Interpigment distances (Å) represent BChl edge-to-edge distances. (<b>B</b>) Interface region between FMO-4 and PscA-2 subunit showing BChl edge-to-edge distances. (<b>C</b>) Potential pathways of energy transfer from nearby light harvesting (LH)-BChls to P<sub>840</sub> special pair located in the periplasmic bilayer. BChl edge-to-edge distances are shown for PsaA-2-associated BChls, while their Mg-to-Mg distances are shown for their PsaA-1-associated counterparts. Phytyl tails of (B)Chls have been removed for clarity.</p>
Full article ">Figure 6
<p>(<b>A</b>) Potential <span class="html-italic">Cba. tepidum</span> cytochrome <span class="html-italic">c</span><sub>Z</sub> binding site as predicted from protein–protein docking analysis [<a href="#B4-biomolecules-14-00311" class="html-bibr">4</a>], using the crystal structure of the C-terminal electron carrier domain [<a href="#B41-biomolecules-14-00311" class="html-bibr">41</a>] with enlarged view at right. Red dotted lines in the right panel show respective distances from the heme Fe to the Mg ions of each of the P<sub>840</sub> special pair BChls. At right, the distance between the heme-associated propionate group and the neighboring PscA Trp601 residues is also shown. (<b>B</b>) Potential ferredoxin-PscB binding site [<a href="#B4-biomolecules-14-00311" class="html-bibr">4</a>], shown in the electrostatic surface representation of the <span class="html-italic">Cab. tepidum</span> FMO-RC-PS. Protein surface coloring represents the electrostatic surface potential (−10 kT, red; +10 kT, blue). An enlarged detailed view of the putative ferredoxin-PscB docking site is seen at right, along with a number of adjacent positively charged residues provided by both PscA-1 and PscB.</p>
Full article ">Figure 7
<p>Structure of the Type I homodimeric RC-PS complex of <span class="html-italic">Cac</span>. <span class="html-italic">thermophilum</span> as determined by cryo-EM at 2.6 Å resolution [<a href="#B6-biomolecules-14-00311" class="html-bibr">6</a>] (PDB ID: 7VZR). (<b>A</b>) View parallel to the membrane plane. (<b>B</b>) Side view along the membrane plane. Main tetrapyrroles of (B)Chls and Heme groups, green and hot pink, respectively; carotenoids, lipids, and unidentified molecules are shown as line models in light blue; [4Fe-4S] clusters F<sub>X</sub>, F<sub>A,</sub> and F<sub>B</sub>, are shown as red and yellow spheres. While PscA has a conserved Type I PS arrangement of 11 TMHs, a unique extramembrane loop region protrudes between helices 7 and 8. (<b>C</b>) Cofactor arrangement of the <span class="html-italic">Cab. themophilum</span> RC electron transfer chain as viewed from membrane plain (<b>left</b>), periplasmic surface (<b>center</b>), and cytoplasmic surface (<b>right</b>). Center-to-center distances between cofactors are presented on the ordinates. PME, phosphatidyl-N-methylethanolamine.</p>
Full article ">Figure 8
<p>Pigment distribution and proposed excitation energy transfer pathways in the <span class="html-italic">Cac. thermophilum</span> RC-PS [<a href="#B6-biomolecules-14-00311" class="html-bibr">6</a>]. (<b>A</b>) Layering of pigments in cytoplasmic and periplasmic membrane bilayer leaflets. Chl <span class="html-italic">a</span>, dark blue; BChl, green; lycopene, brown. (<b>B</b>) Mg<sup>2+</sup> → Mg<sup>2+</sup> distances (Å) for PsaA-1 (B)Chls in the cytoplasmic leaflet. (<b>C</b>) Mg<sup>2+</sup> → Mg<sup>2+</sup> distances (Å) for PsaA-1 (B)Chls in the periplasmic leaflet. (<b>D</b>) Proposed excitation energy pathways of PsaA-2 antenna (B)Chl <span class="html-italic">a</span> → RC as denoted by the dotted red arrows.</p>
Full article ">Figure 9
<p>Structural evidence supporting the role of the Ca<sup>2+</sup>-binding site of the HB RC as a possible evolutionary precursor of the PSII H<sub>2</sub>O oxidizing Mn<sub>4</sub>CaO<sub>5</sub> cluster region [<a href="#B2-biomolecules-14-00311" class="html-bibr">2</a>]. (<b>A</b>) PSII RC heterodimeric D1 and D2 RC core polypeptides (gray structures); Chl protein (CP)43 and CP47 antenna proteins (orange structures); cofactors are shown as stick structures. Boxed regions show an enhanced view of homology between the D1 Y<sub>Z</sub> donor site and Y<sub>D</sub> of the D2 polypeptide (<b>B</b>). The HB (PshA)<sub>2</sub> homodimeric antenna (orange structures comprising α-helices 1–6) and RC core regions (gray structure comprising α-helices 7–11). Spheres in panels A and B represent Fe (orange), S (yellow), Mn (purple), O (red), and Ca (green) atoms. (<b>C</b>) Superpositioning of homodimeric Type I RC Ca<sup>2+</sup>-binding sites within RC-PS structures [<a href="#B6-biomolecules-14-00311" class="html-bibr">6</a>]. (<b>D</b>) Enhanced view of conserved regions of TMH 3 and 4 of D1 polypeptide (orange) and helices 9 and 10 of the PshA RC domain (gray). (<b>E</b>) Enhanced view of matching regions of PsaB of cyanobacterial PSI RC (orange) and PshA RC (gray), demonstrating that helix 11 is two turns greater in length than PshA counterpart (as demarcated by the black and red arrowheads), as well as the lack of Ca<sup>2+</sup> binding site in PSI [<a href="#B2-biomolecules-14-00311" class="html-bibr">2</a>]. (<b>F</b>) Enhanced view of PshA Ca<sup>2+</sup>-binding site illustrating the connection of ligand of Asn263 of the antenna domain of the PshA protein. (<b>G</b>) Enhanced view of the Mn<sub>4</sub>CaO<sub>5</sub> cluster and Y<sub>Z</sub> regions of PSII illustrating how the C-terminal residues Glu354 of CP43 and Ala344 of D1 provide direct ligands to the Ca<sup>2+</sup> atom within the Mn<sub>4</sub>CaO<sub>5</sub> cluster.</p>
Full article ">Figure 10
<p>Overview of evolutionary steps leading to the extant Type I and I RCs (adapted from [<a href="#B2-biomolecules-14-00311" class="html-bibr">2</a>]). Step 1 is based upon sequence analyses and structural and functional changes initiated in the common ancestor containing both the Type II and Type I RC precursors. In steps 2 and 3, evolutionary precursors are further separated into distinct RC-PS precursors, ultimately forming the distinct monophyletic Type II and Type I RC clades. Subsequent heterodimerization occurs in steps 4, 5, and 6 to create the respective extant PSII, PSI, and anoxygenic Type II RCs, and then later acquiring the elliptical heterodimeric LH1 antenna, while PSII obtains the CP43 antenna on the D1 side and the CP47 antenna on the D2 side. The arrangement of antenna (B)Chls in PSI and in the homodimeric HB Type II RC-PS is likely to have already been established in their immediate homodimeric precursors. Notably, PSII shares more characteristics with homodimeric Type I RCs than with Type II anoxygenic counterparts, as detailed in the text. The structures shown at the top consist of RC-LH1 complex from <span class="html-italic">Thermochromatium tepidum</span> (3.0 Å resolution, Protein Data Bank (PDB) ID: 3WMM), PSII of <span class="html-italic">Thermosynechococcus vulcanus</span> (1.9 Å resolution, PDB ID: 3WU2), PSI of <span class="html-italic">Synecho-coccus elongatus</span> (2.5 Å resolution, PDB ID: 1JB0), Type I homodimeric RC-PS of <span class="html-italic">Hmi. modesticaldum</span> (2.2 Å resolution, PDB ID: 5V8K).</p>
Full article ">
16 pages, 24723 KiB  
Article
Sucrose and Mannans Affect Arabidopsis Shoot Gravitropism at the Cell Wall Level
by Gregory Pozhvanov and Dmitry Suslov
Plants 2024, 13(2), 209; https://doi.org/10.3390/plants13020209 - 11 Jan 2024
Viewed by 1422
Abstract
Gravitropism is the plant organ bending in response to gravity. Gravitropism, phototropism and sufficient mechanical strength define the optimal position of young shoots for photosynthesis. Etiolated wild-type Arabidopsis seedlings grown horizontally in the presence of sucrose had a lot more upright hypocotyls than [...] Read more.
Gravitropism is the plant organ bending in response to gravity. Gravitropism, phototropism and sufficient mechanical strength define the optimal position of young shoots for photosynthesis. Etiolated wild-type Arabidopsis seedlings grown horizontally in the presence of sucrose had a lot more upright hypocotyls than seedlings grown without sucrose. We studied the mechanism of this effect at the level of cell wall biomechanics and biochemistry. Sucrose strengthened the bases of hypocotyls and decreased the content of mannans in their cell walls. As sucrose is known to increase the gravitropic bending of hypocotyls, and mannans have recently been shown to interfere with this process, we examined if the effect of sucrose on shoot gravitropism could be partially mediated by mannans. We compared cell wall biomechanics and metabolomics of hypocotyls at the early steps of gravitropic bending in Col-0 plants grown with sucrose and mannan-deficient mutant seedlings. Sucrose and mannans affected gravitropic bending via different mechanisms. Sucrose exerted its effect through cell wall-loosening proteins, while mannans changed the walls’ viscoelasticity. Our data highlight the complexity of shoot gravitropism control at the cell wall level. Full article
(This article belongs to the Special Issue New Perspectives on the Plant Cell Wall)
Show Figures

Figure 1

Figure 1
<p>Sucrose in the growth medium increased the proportion of upright hypocotyls in etiolated Col-0 Arabidopsis seedlings. Five-day-old plants grown without sucrose (<b>a</b>,<b>b</b>) or in the presence of 1% <span class="html-italic">w</span>/<span class="html-italic">v</span> sucrose (<b>c</b>,<b>d</b>) on horizontal Petri plates.</p>
Full article ">Figure 2
<p>Sucrose strengthened the lower part of hypocotyls in five-day-old Arabidopsis seedlings grown on horizontal Petri plates. Five-millimeter-long basal segments of frozen/thawed heat-inactivated hypocotyls were extended at pH 5.0 under a 600 mg load, and their creep rate was measured. ‘Control’—plants grown without sucrose; ‘sucrose’—plants grown in the presence of sucrose (1% <span class="html-italic">w</span>/<span class="html-italic">v</span>). Data are means ± SE (<span class="html-italic">n</span> = 10). An asterisk denotes significant difference (<span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 3
<p>Biochemical composition of cell walls in Arabidopsis seedlings as affected by sucrose. The contents of uronic acids (<b>a</b>), crystalline cellulose (<b>b</b>) and monosaccharides derived from cell wall matrix polymers (<b>c</b>) were determined in five-day-old etiolated Col-0 seedlings grown on horizontal Petri plates without sucrose (control) or in the presence of sucrose (1% <span class="html-italic">w</span>/<span class="html-italic">v</span>). Data are means ± SE (<span class="html-italic">n</span> = 9). Asterisks denote significant differences (<span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 4
<p>Cell wall biomechanics of hypocotyls from gravistimulated and nongravistimulated Col-0 Arabidopsis seedlings grown without sucrose (control) or in the presence of sucrose (1% <span class="html-italic">w</span>/<span class="html-italic">v</span>). Etiolated seedlings grown on vertical Petri plates for 3 days were either gravistimulated via a 90° counterclockwise rotation of the plates or left unstimulated. Both groups of seedlings were frozen 4 h and 6 h after the moment of gravistimulation. Three-millimeter-long subapical segments of frozen/thawed hypocotyls were extended under a 600 mg load, and their creep rate was measured. Data are means ± SE (<span class="html-italic">n</span> = 8). Different letters denote the only significant difference between gravistimulated and nongravistimulated hypocotyls (<span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 5
<p>Cell wall biomechanics of hypocotyls from gravistimulated and nongravistimulated wild-type Col-0 Arabidopsis seedlings or mannan-deficient <span class="html-italic">csla2csla3csla9</span> triple mutants. Etiolated seedlings grown on vertical Petri plates for 3 days were either gravistimulated via a 90° counterclockwise rotation of the plates or left unstimulated. Both groups of seedlings were frozen 4 h and 6 h after the moment of gravistimulation. Three-millimeter-long subapical segments of frozen/thawed hypocotyls were extended under a 600 mg load, and their creep rate was measured. Data are means ± SE (<span class="html-italic">n</span> = 8). Different letters denote the only significant difference between gravistimulated and nongravistimulated hypocotyls (<span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Cellulose macrofibril arrangement in the outer epidermal cell wall from the subapical part of five-day-old Col-0 hypocotyls. Spinning-disc confocal microscopy on wall samples stained with Pontamine Fast Scarlet 4B. Subsequent optical sections of a representative Z-stack from the innermost to the outermost wall layer. Scale bars are 10 μm.</p>
Full article ">Figure 7
<p>Cellulose macrofibril arrangement in the outer epidermal cell wall from the subapical part of five-day-old <span class="html-italic">csla2csla3csla9</span> hypocotyls. Spinning-disc confocal microscopy on wall samples stained with Pontamine Fast Scarlet 4B. Subsequent optical sections of a representative Z-stack from the innermost to the outermost wall layer. Scale bars are 10 μm.</p>
Full article ">Figure 8
<p>Partial least squares discriminant analysis (PLS-DA) of metabolite profiles of Arabidopsis hypocotyls. Three-day-old etiolated Col-0 seedlings grown on vertical Petri plates without sucrose or in the presence of sucrose (1% <span class="html-italic">w</span>/<span class="html-italic">v</span>) were gravistimulated via a 90° counterclockwise rotation of the plates. PLS-DA scores are plotted along the X and Y axes, which are Component 1 and Component 2, respectively. Percentage in brackets indicates the explained variance of metabolite content. Abbreviations: C—nongravistimulated wild-type Col-0 seedlings, grown without sucrose; S—nongravistimulated wild-type Col-0 seedlings, grown in the presence of sucrose (1% <span class="html-italic">w</span>/<span class="html-italic">v</span>); CG—gravistimulated wild-type Col-0 seedlings, grown without sucrose; SG—gravistimulated wild-type Col-0 seedlings, grown in the presence of sucrose (1% <span class="html-italic">w</span>/<span class="html-italic">v</span>). Each circle shows a corresponding biological replicate. Large ellipses depict 95% confidence regions.</p>
Full article ">Figure 9
<p>Partial least squares discriminant analysis (PLS-DA) of metabolite profiles of wild-type Col-0 and <span class="html-italic">csla2csla3csla9</span> Arabidopsis hypocotyls. Three-day-old etiolated wild-type Col-0 and <span class="html-italic">csla2csla3csla9</span> seedlings grown on vertical Petri plates without sucrose were gravistimulated via a 90° counterclockwise rotation of the plates. PLS-DA scores are plotted along X and Y axes that are Component 1 and Component 2, respectively. Percentage in brackets indicate the explained variance for metabolite content. Abbreviations: WT—nongravistimulated wild-type Col-0 seedlings; M—nongravistimulated <span class="html-italic">csla2csla3csla9</span> triple mutant seedlings; WTG—gravistimulated wild-type Col-0 seedlings; MG—gravistimulated <span class="html-italic">csla2csla3csla9</span> triple mutant seedlings. Each circle shows a corresponding biological replicate. Large ellipses depict 95% confidence regions.</p>
Full article ">
19 pages, 3211 KiB  
Article
Phototactic Behavioral Responses of Mesozooplankton in the Barents Sea as an Indicator of Anthropogenic Impact
by Victor Dyomin, Yuri Morgalev, Sergey Morgalev, Alexandra Davydova, Oksana Kondratova, Tamara Morgaleva and Igor Polovtsev
Water 2023, 15(22), 3901; https://doi.org/10.3390/w15223901 - 8 Nov 2023
Cited by 1 | Viewed by 1437
Abstract
The behavioral responses of autochthonous organisms have recently been used for a system to monitor the state of fresh and sea waters for bioindication. The advantage of using the behavioral responses of mesozooplankton is determined by the higher sensitivity of such responses compared [...] Read more.
The behavioral responses of autochthonous organisms have recently been used for a system to monitor the state of fresh and sea waters for bioindication. The advantage of using the behavioral responses of mesozooplankton is determined by the higher sensitivity of such responses compared with changes in the composition of biota or the death of organisms. Earlier, we developed and tested in laboratory conditions and in freshwater reservoirs a submersible digital holographic camera as part of a hydrobiological probe, which allows one to determine the dimensions, shape and recognition of plankters in situ, as well as define the concentration of plankters in the working volume and perform photostimulation with attractive radiation with different levels of illuminance. This paper presents the data obtained during the expedition to the Barents Sea. The variability with regard to the immersion depth of the phototropic response and the interspecific and intraspecific diversity was determined. It was shown that within the framework of natural variability in natural factors (temperature, salinity, hydrostatic pressure, oxygen content, illumination) there are no reliable changes in the indicator response, unlike changes in the concentration of plankton associated with tidal currents. The anthropogenic distortion of water quality was modeled by introducing a saturated salt solution dropwise. There were no significant changes in the intraspecific and interspecific diversity index during the external impact, and the rhythms of tidal changes in the concentration of plankters were suppressed. The fact of increased phototropic sensitivity in crustaceans with a size of less than 120 μm was found. It was established that the most essential marker of the alternating factor was the suppression of the phototropic response. The identified patterns of behavioral responses of autochthonous zooplankton make it possible to create a network of continuous control over the environmental health of water bodies subject to increased anthropogenic impact (oil production zones beyond the Arctic Circle, estuaries and deltas of rivers carrying industrial waste). Full article
Show Figures

Figure 1

Figure 1
<p>Research site 69°07′05″ N, 36°03′30″ E (<b>a</b>); installation of DHC-probe at the bottom station (<b>b</b>,<b>d</b>); and the scheme of a digital holographic camera (<b>c</b>). In figure (<b>c</b>): semiconductor laser 1 (λ = 650 nm), semiconductor laser 2 (λ = 532 nm), optical multiplexer (mixer) 3, scattering lens 4, collimating and receiving objectives 5, windows 6, prisms 7, selective filter 8 and CMOS camera 9. In the figure, the lighting module is enclosed in hermetic housing I, and the recording module is enclosed in hermetic housing II. Red area—working volume formed by prisms.</p>
Full article ">Figure 2
<p>Averaged concentrations of crustaceans in successive profiles (<b>a</b>) and correlation of the DHC and Juday net catching data (<b>b</b>).</p>
Full article ">Figure 3
<p>Depth profiles of crustacean concentration (<b>a</b>), phototropic response (<b>b</b>), and mesozooplankton community entropy (<b>c</b>) by size (Hsize) and species (Hord).</p>
Full article ">Figure 4
<p>Phototropic response of zooplankton (total amount of Cladocera and Copepoda) at different durations of photostimulation stages (<b>a</b>) and changes in the concentration of small (&lt;120 μm) and large (&gt;120 μm) individuals in the working volume at 10 min stimulation stages (<b>b</b>). Green lines—the boundaries of photostimulation steps.</p>
Full article ">Figure 5
<p>Dynamics of mesozooplankton concentration (<b>a</b>) and phototropic response (<b>b</b>).</p>
Full article ">Figure 6
<p>Change in the phototropic response of mesozooplankton community (<b>a</b>), Copepoda (<b>b</b>) and Cladocera (<b>c</b>) in the experiment with salt solution.</p>
Full article ">Figure 7
<p>Histograms of the plankters’ size distribution in the zooplankton community in the DHC working volume before (<b>a</b>) and during (<b>b</b>) the introduction of salt solution.</p>
Full article ">Figure 8
<p>pPhR spectrograms before (<b>a</b>), during (<b>b</b>) and after (<b>c</b>) introduction of salt solution.</p>
Full article ">
16 pages, 3995 KiB  
Article
Phototropic Behavioral Responses of Zooplankton in Lake Baikal In Situ and during the Anthropogenic Impact Modeling
by Victor Dyomin, Yuri Morgalev, Igor Polovtsev, Sergey Morgalev, Tamara Morgaleva, Alexandra Davydova and Oksana Kondratova
Water 2023, 15(16), 2957; https://doi.org/10.3390/w15162957 - 16 Aug 2023
Cited by 3 | Viewed by 1182
Abstract
Earlier, we showed that the registration of the behavioral responses of autochthonous mesozooplankton communities in situ is a more dynamic methodological approach in the biological assessment of the environmental well-being of aquatic ecosystems, as well as an alternative method to generally accepted tests [...] Read more.
Earlier, we showed that the registration of the behavioral responses of autochthonous mesozooplankton communities in situ is a more dynamic methodological approach in the biological assessment of the environmental well-being of aquatic ecosystems, as well as an alternative method to generally accepted tests on mortality and immobilization. The change in behavioral responses, including phototropic responses, may occur at lower concentrations of pollutants, leading to the inhibition of the risk-avoidance response of predatory fish attack and, ultimately, to the change in zooplankton abundance and biodiversity. The biological significance of such changes is quite high since zooplankters form the basis of food chains. This work studies the possibility of biomonitoring the quality of fresh water in Lake Baikal according to the state of the autochthonous mesozooplankton community in summer and winter using a digital holographic camera developed and tested by us in laboratory conditions. This method makes it possible to determine the concentration of plankters in the controlled volume of the DHC and perform photostimulation with different levels of illuminance. The depth profilometry of the phototropic response was compared with the profilometry of plankton concentration, intraspecific diversity of crustaceans according to the Pielou index, and the results of catching using the Juday net in the natural environment of the lake and during the modeling of the anthropogenic impact (introduction of table salt solution into the local area close to the registration probe). The circadian rhythm parameters were determined by the spectral analysis of the long-term registration of the phototropic response dynamics. It was noted that the inhibition of the phototropic response was the most adequate marker of the exogenous impact and the appearance of an alternating factor among the studied indicators of the state of the plankton community, namely, intraspecific diversity, synchronism of circadian rhythms, and response to paired photostimulation. The revealed patterns of behavioral responses of autochthonous zooplankton in natural and artificially modified conditions will allow for the implementation of long-term continuous control over the environmental well-being of water areas, including the collection ponds of treatment facilities, cooling ponds of nuclear power plants, and other water areas in contact with potentially hazardous facilities. The comparison of the identified patterns with the behavioral responses of euryhaline mesozooplankton will expand this method to assess the well-being of salt-water and marine reservoirs under the anthropogenic impact and will make it possible to create a continuous monitoring system. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the DHC probe during summer (<b>a</b>) and winter (<b>b</b>) expeditions, digital holographic camera (<b>c</b>), probe installation at the bottom station (<b>d</b>), and from ice (<b>e</b>). The arrow in Figure (<b>a</b>) indicates the place of the buoy above the camera (50 m offshore). The arrow in Figure (<b>b</b>) shows the place of the hole (100 m offshore). In Figure (<b>c</b>): semiconductor laser 1 (λ = 650 nm); semiconductor laser 2 (λ = 532 nm); optical multiplexer (mixer) 3; scattering lens 4; collimating and receiving objectives 5; windows 6; prisms 7; selective filter 8; and CMOS camera 9. In the figure, the lighting module is enclosed in hermetic housing I, and the recording module is enclosed in hermetic housing II. α—working volume formed by prisms.</p>
Full article ">Figure 2
<p>Average total (<b>a</b>), “day” (<b>b</b>), and “night” (<b>c</b>) concentration profile of Copepoda by depth.</p>
Full article ">Figure 3
<p>Cross-correlation functions of the relationship between the daily dynamics of the average number of crustaceans in the profile measured by the DHC probe and the Juday net during summer (<b>a</b>) and winter (<b>b</b>). Along the Y-axis—shift of the cross-correlation function in hours, along the X-axis—value of the correlation coefficient. Red lines—95% reliability boundaries.</p>
Full article ">Figure 4
<p>Increase in crustacean concentration in the controlled volume of the DHC under different schemes of the paired photostimulation. Δ1—increase in C when the first stage of photostimulation is activated; Δ2—increase in C when the second stage of photostimulation is activated.</p>
Full article ">Figure 5
<p>pPhR profiles and concentrations of crustaceans in the DHC controlled volume at different levels of photostimulation (<b>a</b>) and correlation coefficients of profiles (<b>b</b>).</p>
Full article ">Figure 6
<p>Histograms of the size distribution of crustaceans at a depth of 5 m (<b>a</b>), 40 m (<b>b</b>), and Shannon index profiles (<b>c</b>) without (H<sub>0</sub>) at first (H<sub>1</sub>) and second (H<sub>2</sub>) photostimulation stages. (<b>a</b>,<b>b</b>) Y-axis—number of the size class; X-axis—share of individuals of this size class.</p>
Full article ">Figure 7
<p>Fragment of the time series of pPhR values (<b>a</b>) and spectrograms of rhythmic processes in the background (<b>b</b>) and during the modeling of anthropogenic impact (<b>c</b>). Impact—period of anthropogenic impact modeling.</p>
Full article ">
16 pages, 3123 KiB  
Article
Top and Side Lighting Induce Morphophysiological Improvements in Korean Ginseng Sprouts (Panax ginseng C.A. Meyer) Grown from One-Year-Old Roots
by Jingli Yang, Jinnan Song, Jayabalan Shilpha and Byoung Ryong Jeong
Plants 2023, 12(15), 2849; https://doi.org/10.3390/plants12152849 - 2 Aug 2023
Cited by 5 | Viewed by 1540
Abstract
Nowadays, not only the roots, but also leaves and flowers of ginseng are increasingly popular ingredients in supplements for healthcare products and traditional medicine. The cultivation of the shade-loving crop, ginseng, is very demanding in terms of the light environment. Along with the [...] Read more.
Nowadays, not only the roots, but also leaves and flowers of ginseng are increasingly popular ingredients in supplements for healthcare products and traditional medicine. The cultivation of the shade-loving crop, ginseng, is very demanding in terms of the light environment. Along with the intensity and duration, light direction is another important factor in regulating plant morphophysiology. In the current study, three lighting directions—top (T), side (S), or top + side (TS)—with an intensity of 30 ± 5 μmol·m−2·s−1 photosynthetic photon flux density (PPFD) were employed. Generally, compared with the single T lighting, the composite lighting direction, TS, was more effective in shaping the ginseng with improved characteristics, including shortened, thick shoots; enlarged, thick leaves; more leaf trichomes; earlier flower bud formation; and enhanced photosynthesis. The single S light resulted in the worst growth parameters and strongly inhibited the flower bud formation, leading to the latest flower bud observation. Additionally, the S lighting acted as a positive factor in increasing the leaf thickness and number of trichomes on the leaf adaxial surface. However, the participation of the T lighting weakened these traits. Overall, the TS lighting was the optimal direction for improving the growth and development traits in ginseng. This preliminary research may provide new ideas and orientations in ginseng cultivation lodging resistance and improving the supply of ginseng roots, leaves, and flowers to the market. Full article
(This article belongs to the Special Issue Plants towards the Light: The Phototropic Growth)
Show Figures

Figure 1

Figure 1
<p>Morphology (<b>a</b>,<b>c</b>), morphologic parameters (<b>b</b>,<b>d</b>), and the relative growth rate (<b>e</b>) of ginseng shoots, leaves, and roots as affected by the different lighting directions for 21 days. T, top; TS, top + side; S, side. Leaves in (<b>c</b>) are the intermediate compound leaves (<b>f</b>) of ginseng plants. Leaf length and width were measured according to the intermediate simple leaf of an intermediate compound leaf (<b>g</b>). The lowercase letters indicate significant separation within treatments by the Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05 in the same cultivar. Vertical bars indicate the means ± standard error (<span class="html-italic">n</span> = 12).</p>
Full article ">Figure 2
<p>Epidermal hair micro-observation (<b>a</b>–<b>c</b>) and thickness (<b>d</b>) of ginseng leaves as affected by the different lighting directions for 21 days. T, top; TS, top + side; S, side. Leaf epidermal hair micro-observations and thicknesses were based on the intermediate simple leaf of an intermediate compound leaf (as shown in <a href="#plants-12-02849-f001" class="html-fig">Figure 1</a>g). The lowercase letters indicate significant separation within treatments by the Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05 in the same cultivar. Vertical bars indicate the means ± standard error (<span class="html-italic">n</span> = 12).</p>
Full article ">Figure 3
<p>Flower bud state observation (<b>a</b>–<b>c</b>) and days (<b>d</b>) to the first visible flower bud in ginseng plants, as affected by the different lighting directions for 21 days. T, top; TS, top + side; S, side. The lowercase letters indicate significant separation within treatments by the Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05 in the same cultivar. Vertical bars indicate the means ± standard error (<span class="html-italic">n</span> = 12).</p>
Full article ">Figure 4
<p>The photosynthesis-related pigments in ginseng leaves as affected by the different lighting directions for 21 days. Chlorophyll a (<b>a</b>), chlorophyll b (<b>b</b>), chlorophyll a + b (<b>c</b>), carotenoid (<b>d</b>), chlorophyll a/chlorophyll b (<b>e</b>), and carotenoid/chlorophyll (<b>f</b>). T, top; TS, top + side; S, side. The lowercase letters indicate significant separation within treatments by the Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05 in the same cultivar. Vertical bars indicate the means ± standard error (<span class="html-italic">n</span> = 12).</p>
Full article ">Figure 5
<p>One-year-old roots of Korean ginseng (<b>a</b>); top and side views of the rectangular planting container (length 52.0 cm × width 36.0 cm × height 8.5 cm) (<b>b</b>); planting pattern (<b>c</b>).</p>
Full article ">Figure 6
<p>The spectral distribution of the experimental light treatments (<b>a</b>): the daily white light (~400–720 nm, peaked at 435 nm) provided by white LEDs; the experimental layout and lighting direction design employed in this study (<b>b</b>). Side, S; top + side, TS; and top, T.</p>
Full article ">
26 pages, 7219 KiB  
Article
Phylogenomic Analysis of micro-RNA Involved in Juvenile to Flowering-Stage Transition in Photophilic Rice and Its Sister Species
by Prasanta K. Dash, Payal Gupta, Rohini Sreevathsa, Sharat Kumar Pradhan, Tenkabailu Dharmanna Sanjay, Mihir Ranjan Mohanty, Pravat K. Roul, Nagendra K. Singh and Rhitu Rai
Cells 2023, 12(10), 1370; https://doi.org/10.3390/cells12101370 - 12 May 2023
Cited by 3 | Viewed by 1636
Abstract
Vegetative to reproductive phase transition in phototropic plants is an important developmental process and is sequentially mediated by the expression of micro-RNA MIR172. To obtain insight into the evolution, adaptation, and function of MIR172 in photophilic rice and its wild relatives, we [...] Read more.
Vegetative to reproductive phase transition in phototropic plants is an important developmental process and is sequentially mediated by the expression of micro-RNA MIR172. To obtain insight into the evolution, adaptation, and function of MIR172 in photophilic rice and its wild relatives, we analyzed the genescape of a 100 kb segment harboring MIR172 homologs from 11 genomes. The expression analysis of MIR172 revealed its incremental accumulation from the 2-leaf to 10-leaf stage, with maximum expression coinciding with the flag-leaf stage in rice. Nonetheless, the microsynteny analysis of MIR172s revealed collinearity within the genus Oryza, but a loss of synteny was observed in (i) MIR172A in O. barthii (AA) and O. glaberima (AA); (ii) MIR172B in O. brachyantha (FF); and (iii) MIR172C in O. punctata (BB). Phylogenetic analysis of precursor sequences/region of MIR172 revealed a distinct tri-modal clade of evolution. The genomic information generated in this investigation through comparative analysis of MIRNA, suggests mature MIR172s to have evolved in a disruptive and conservative mode amongst all Oryza species with a common origin of descent. Further, the phylogenomic delineation provided an insight into the adaptation and molecular evolution of MIR172 to changing environmental conditions (biotic and abiotic) of phototropic rice through natural selection and the opportunity to harness untapped genomic regions from rice wild relatives (RWR). Full article
Show Figures

Figure 1

Figure 1
<p>Identification and location of <span class="html-italic">MIR172</span> and its homologs amongst analyzed poaceae members. Genomic location of (<b>a</b>) <span class="html-italic">MIR172</span>A (Scales: 0–30 Mb and 0–210 Mb); Osa9-<span class="html-italic">Oryza sativa</span> chr 9, Ogl2-<span class="html-italic">Oryza glaberrima</span> chr 2, Oba2-<span class="html-italic">Oryza barthii</span> chr 2, Oglu9-<span class="html-italic">Oryza glumaepatula</span> chr 9, Obr9-<span class="html-italic">Oryza brachyantha</span> chr 9, Oru9-<span class="html-italic">Oryza rufipogon</span> chr 9, Opu9-<span class="html-italic">Oryza punctata</span> chr 9, Zma7-<span class="html-italic">Zea mays</span> chr 7, Sbi9- <span class="html-italic">Sorghum bicolor</span> chr 9, Ath2-<span class="html-italic">Arabidopsis thaliana</span> chr 2; (<b>b</b>) <span class="html-italic">MIR172</span>B (Scales: 0–60 Mb and 0–240 Mb); Osa1-<span class="html-italic">Oryza sativa</span> chr 1, Ogl1-<span class="html-italic">Oryza glaberrima</span> chr 1, Oba1-<span class="html-italic">Oryza barthii</span> chr 1, Oglu1-<span class="html-italic">Oryza glumaepatula</span> chr 1, Obr1-<span class="html-italic">Oryza brachyantha</span> chr 1, Oru1-<span class="html-italic">Oryza rufipogon</span> chr 1, Opu1-<span class="html-italic">Oryza punctata</span> chr 1, Zma5-<span class="html-italic">Zea mays</span> chr 5, Sbi3-<span class="html-italic">Sorghum bicolor</span> chr 3, Ath5-<span class="html-italic">Arabidopsis thaliana</span> chr 5; (<b>c</b>) <span class="html-italic">MIR172</span>C (Scales:0–40 Mb and 0–300 Mb); Osa7-<span class="html-italic">Oryza sativa</span> chr 7, Ogl7-<span class="html-italic">Oryza glaberrima</span> chr 7, Oba7-<span class="html-italic">Oryza barthii</span> chr 7, Oru7-<span class="html-italic">Oryza rufipogon</span> chr 7, Oglu7-<span class="html-italic">Oryza glumaepatula</span> chr 7, Opu2<span class="html-italic">-Oryza punctata</span> chr 2, Obr7-<span class="html-italic">Oryza brachyantha</span> chr 7, Zma4-<span class="html-italic">Zea mays</span> chr 4, Sbi4-<span class="html-italic">Sorghum bicolor</span> chr 4, Ath3-<span class="html-italic">Arabidopsis thaliana</span> chr 3; (<b>d</b>) <span class="html-italic">MIR172</span>D (Scales: 0–40 Mb and 0–210 Mb) Osai2-<span class="html-italic">Oryza sativa</span> homeolog i chr 2, Osaii2-<span class="html-italic">Oryza sativa</span> homeolog ii chr 2, Ogl2-<span class="html-italic">Oryza glaberrima</span> chr 2, Oba2-<span class="html-italic">Oryza barthii</span> chr 2, Oglu2-<span class="html-italic">Oryza glumaepatula</span> chr 2, Obr2-<span class="html-italic">Oryza brachyantha</span> chr 2, Oru2-<span class="html-italic">Oryza rufipogon</span> chr 2, Opu2-<span class="html-italic">Oryza punctata</span> chr 2, Zma6-<span class="html-italic">Zea mays</span> chr 6, Sbi2-<span class="html-italic">Sorghum bicolor</span> chr 2, Ath3-<span class="html-italic">Arabidopsis thaliana</span> chr 3; on chromosomes of different poaceae members.</p>
Full article ">Figure 2
<p>Expression profiling of <span class="html-italic">MIR172</span> in rice, other cereals, and A. thaliana. Expression pattern obtained by qRT-PCR of each block shows log<sub>2</sub>-fold expression in different stage-specific developmental tissue. Details of specific stages used for each plant species: <span class="html-italic">Oryza</span> spp. (2LS-2 leaf-shoot, 2LR-2 leaf-root, 4LS-4 leaf-shoot, 4LR-4 leaf-root, 10LS-10 leaf-shoot apical, 10LR-10 leaf-root, FL-flag leaf, BP-booting panicle, p (&lt;0.5 cm)-panicle (&lt;0.5 cm), p (0.5–1 cm)-panicle (0.5–1 cm), p (1–2 cm)-panicle (1–2 cm) and p (2–4 cm)-panicle (2–4 cm)); <span class="html-italic">Sorghum bicolor</span> (3LS-3 leaf-shoot, 3LR-3 leaf-root, 5LS-5 leaf-shoot, 5LR-5 leaf-root, GPD-growing point differentiation, FL-flag leaf, BP-booting panicle, F-flower not yet bloomed, F-blooming-flower-blooming, G-blooming complete); <span class="html-italic">Zea mays</span> (2LS-2 leaf-shoot, 2LR-2 leaf-root, 4LS-4 leaf-shoot, 10LS-10 leaf-shoot, 10LR-10 leaf-root, FL-flag leaf, T-Tassel and S-silk) and <span class="html-italic">Arabidopsis thaliana</span> (2RLS-2 rosette leaf-shoot, 2RLR-2 rosette leaf-root, 4RLS-4 rosette leaf -shoot, 10RLS-10 leaf rosette-shoot, 10RLR-10 rosette leaf -root, CRG-complete rosette growth, I-Inflorescence, FB-floral bud, and OF-open flower).</p>
Full article ">Figure 3
<p>Multiple sequence alignment of mature <span class="html-italic">MIR172</span> exhibiting conserved and diverged regions. (<b>a</b>) Highly conserved mature <span class="html-italic">MIR172</span>A sequences amongst poaceae members and Arabidopsis, (<b>b</b>) Prevalence of SNPs and InDels (insertions/deletions) detected at multiple positions in Zma-172B and At-172B by multiple sequence alignment of mature <span class="html-italic">MIR172</span>B. Conserved mature sequence of <span class="html-italic">MIR172</span>B in seven <span class="html-italic">Oryza</span> spp. * Denotes position of single nucleotide polymorphism. (<b>c</b>) Sequence alignment of mature <span class="html-italic">MIR172</span>C of all poaceae members vis-a-vis Arabidopsis shows the prevalence of substitutions. Single nucleotide substitution where ‘T’ in rice is replaced by ‘A’ in <span class="html-italic">S. bicolor</span>, <span class="html-italic">Z. mays,</span> and Arabidopsis while ‘C’ in rice is replaced by ‘T’ in <span class="html-italic">S. bicolor</span>, <span class="html-italic">Z. mays,</span> and Arabidopsis; in highly conserved mature <span class="html-italic">MIR172</span>C of all poaceae members and Arabidopsis and (<b>d</b>) Sequence alignment of mature <span class="html-italic">MIR172</span>D of all poaceae members vis-a-vis Arabidopsis. Osa-<span class="html-italic">Oryza sativa</span>, Oba-<span class="html-italic">Oryza barthii</span>, Ogl-<span class="html-italic">Oryza glaberrima</span>, Oglu-<span class="html-italic">Oryza glumaepatula</span>, Oru-<span class="html-italic">Oryza rufipogon</span>, Obr-<span class="html-italic">Oryza brachyantha</span>, Opu-<span class="html-italic">Oryza punctata</span>, Sbi-<span class="html-italic">Sorghum bicolor</span>, Zma-<span class="html-italic">Zea mays</span>, Tae-<span class="html-italic">Triticum aestivum</span>, Ath-<span class="html-italic">Arabidopsis thaliana</span>.</p>
Full article ">Figure 4
<p>Multiple sequence alignment for elucidation of conservation and divergence in <span class="html-italic">MIR172</span>A and B precursor region (<b>a</b>) Precursor sequence alignment of <span class="html-italic">MIR172</span>A across poaceae; and (<b>b</b>) Precursor sequence alignment of <span class="html-italic">MIR172</span>B across poaceae. The sequences in the red box are regions of the sequences that are highly conserved among all the species included in the study. * denotes the sites for single nucleotide polymorphism. Osa-<span class="html-italic">Oryza sativa</span>, Ogl-<span class="html-italic">Oryza glaberrima</span>, Oba-<span class="html-italic">Oryza barthii</span>, Oru-<span class="html-italic">Oryza rufipogon</span>, Oglu-<span class="html-italic">Oryza glumaepatula</span>, Opu-<span class="html-italic">Oryza punctata</span>, Obr-<span class="html-italic">Oryza brachyantha</span>, Zma-<span class="html-italic">Zea mays</span>, Sbi-<span class="html-italic">Sorghum bicolor</span>, Ath-<span class="html-italic">Arabidopsis thaliana</span>.</p>
Full article ">Figure 5
<p>Multiple sequence alignment discerns conservation and divergence in the precursor region of <span class="html-italic">MIR172</span>C and D. (<b>a</b>) <span class="html-italic">MIR172</span>C precursor sequence alignment in poaceae; (<b>b</b>) <span class="html-italic">MIR172</span>D precursor sequence alignment in poaceae. The sequences in the red box are regions of the sequences that are highly conserved among all the species included in the study. * denotes the sites of single nucleotide polymorphism. Osa-<span class="html-italic">Oryza sativa</span>, Ogl-<span class="html-italic">Oryza glaberrima</span>, Oba-<span class="html-italic">Oryza barthii</span>, Oru-<span class="html-italic">Oryza rufipogon</span>, Oglu-<span class="html-italic">Oryza glumaepatula</span>, Opu-<span class="html-italic">Oryza punctata</span>, Obr-<span class="html-italic">Oryza brachyantha</span>, Zma-<span class="html-italic">Zea mays</span>, Sbi-<span class="html-italic">Sorghum bicolor</span>, Ath-<span class="html-italic">Arabidopsis thaliana</span>.</p>
Full article ">Figure 6
<p>Cartoon representation of (<b>a</b>) 100 kb genomic segments containing <span class="html-italic">MIR172</span>A/B/C/D conserved between <span class="html-italic">O. sativa</span> and other poaceae members; and (<b>b</b>) the gene density in 100 kb genomic segments containing <span class="html-italic">MIR172</span>A/B/C/D between <span class="html-italic">O. sativa</span> and other poaceae members.</p>
Full article ">Figure 7
<p>Evaluation of 100 kb regions of <span class="html-italic">MIR172</span> amongst poaceae family. Collinearity blocks of (<b>a</b>) <span class="html-italic">MIR172</span>A, (<b>b</b>) <span class="html-italic">MIR172</span>B, (<b>c</b>) <span class="html-italic">MIR172</span>C, and (<b>d</b>) <span class="html-italic">MIR172</span>D with <span class="html-italic">Oryza sativa</span> as reference. The first column indicates the depth of duplication at each gene locus; the second column indicates the genes in reference chromosomes, and the subsequent columns represent aligned collinear blocks with the matched genes. Alignment among non-anchor genes is removed in the output and represented by ‘||’ in the multi-alignment of gene ordering.</p>
Full article ">Figure 8
<p>Analysis of 100 kb region flanking <span class="html-italic">MIR172</span> amongst Poaceae members. (<b>a</b>) Circular plot of <span class="html-italic">MIR172</span>A displaying synteny and collinearity patterns. Os9-<span class="html-italic">Oryza sativa</span> chr 9, og2-<span class="html-italic">Oryza glaberrima</span> chr 2, ob2-<span class="html-italic">Oryza barthii</span> chr 2, ou9-<span class="html-italic">Oryza glumaepatula</span> chr 9, oa9-<span class="html-italic">Oryza brachyantha</span> chr 9, or9-<span class="html-italic">Oryza rufipogon</span> chr 9, op9-<span class="html-italic">Oryza punctata</span> chr 9, zm7-<span class="html-italic">Zea mays</span> chr 7, sb9-<span class="html-italic">Sorghum bicolor</span> chr 9, at2-<span class="html-italic">Arabidopsis thaliana</span> chr 2; (<b>b</b>) Circular plot displaying synteny and collinearity patterns in <span class="html-italic">MIR172</span>B. os1-<span class="html-italic">Oryza sativa</span> chr 1, og1-<span class="html-italic">Oryza glaberrima</span> chr 1, ob1-<span class="html-italic">Oryza barthii</span> chr 1, ou1-<span class="html-italic">Oryza glumaepatula</span> chr 1, oa1-<span class="html-italic">Oryza brachyantha</span> chr 1, or1-<span class="html-italic">Oryza rufipogon</span> chr 1, op1-<span class="html-italic">Oryza punctata</span> chr 1, zm5-<span class="html-italic">Zea mays</span> chr 5, sb3-<span class="html-italic">Sorghum bicolor</span> chr 3, at5-<span class="html-italic">Arabidopsis thaliana</span> chr 5; (<b>c</b>) Circular plot displaying synteny and collinearity patterns in <span class="html-italic">MIR172</span>C. os7-<span class="html-italic">Oryza sativa</span> chr 7, og7-<span class="html-italic">Oryza glaberrima</span> chr 7, ob7-<span class="html-italic">Oryza barthii</span> chr 7, or7-<span class="html-italic">Oryza rufipogon</span> chr 7, ou7-<span class="html-italic">Oryza glumaepatula</span> chr 7, op2<span class="html-italic">-Oryza punctata</span> chr 2, oa7-<span class="html-italic">Oryza brachyantha</span> chr 7, zm4-<span class="html-italic">Zea mays</span> chr 4, sb4-<span class="html-italic">Sorghum bicolor</span> chr 4, at3-<span class="html-italic">Arabidopsis thaliana</span> chr 3; and (<b>d</b>) Circular plot displaying synteny and collinearity patterns in <span class="html-italic">MIR172</span>D. osi2-<span class="html-italic">Oryza sativa</span> homeolog i chr 2, osii2-<span class="html-italic">Oryza sativa</span> homeolog ii chr 2, og2-<span class="html-italic">Oryza glaberrima</span> chr 2, ob2-<span class="html-italic">Oryza barthii</span> chr 2, ou2-<span class="html-italic">Oryza glumaepatula</span> chr 2, oa2-<span class="html-italic">Oryza brachyantha</span> chr 2, or2-<span class="html-italic">Oryza rufipogon</span> chr 2, op2-<span class="html-italic">Oryza punctata</span> chr 2, zm6-<span class="html-italic">Zea mays</span> chr 6, sb2-<span class="html-italic">Sorghum bicolor</span> chr 2, at3-<span class="html-italic">Arabidopsis thaliana</span> chr 3.</p>
Full article ">Figure 9
<p>Phylogenetic analysis of <span class="html-italic">MIR172</span>. The ML phylogenetic tree shows 40 <span class="html-italic">MIR172</span> precursors and promoter sequences (500 bp) from <span class="html-italic">Oryza sativa</span> and its six wild cousins, <span class="html-italic">Sorghum bicolor</span> and <span class="html-italic">Zea mays</span>. Arabidopsis was included as an outlier. Bootstrap values (0.53 to 1.000) are represented by triangles. Maximum-likelihood tree revealed that <span class="html-italic">MIR172</span> sequences clustered into three clades viz. Clade-I, Clade-II, and Clade-III. Osa-<span class="html-italic">Oryza sativa</span>, Ogl-<span class="html-italic">Oryza glaberrima</span>, Oba-<span class="html-italic">Oryza barthii</span>, Oru-<span class="html-italic">Oryza rufipogon</span>, Oglu-<span class="html-italic">Oryza glumaepatula</span>, Opu-<span class="html-italic">Oryza punctata</span>, Obr-<span class="html-italic">Oryza brachyantha</span>, Zma-<span class="html-italic">Zea mays</span>, Sbi-<span class="html-italic">Sorghum bicolor</span>, Ath-<span class="html-italic">Arabidopsis thaliana</span>.</p>
Full article ">
12 pages, 1407 KiB  
Article
Determination of Hourly Distribution of Tuta absoluta (Meyrick) (Lepidoptera: Gelechiidae) Using Sex Pheromone and Ultraviolet Light Traps in Protected Tomato Crops
by Gui-Fen Zhang, Yi-Bo Zhang, Lin Zhao, Yu-Sheng Wang, Cong Huang, Zhi-Chuang Lü, Ping Li, Wan-Cai Liu, Xiao-Qing Xian, Jing-Na Zhao, Ya-Hong Li, Fang-Hao Wan, Wan-Xue Liu and Fu-Lian Wang
Horticulturae 2023, 9(3), 402; https://doi.org/10.3390/horticulturae9030402 - 20 Mar 2023
Cited by 2 | Viewed by 2215
Abstract
Tuta absoluta (Meyrick), a leafminer that damages tomato leaves, terminal buds, flowers, and fruits, is a destructive tomato pest and is responsible for 80–100% of tomato yield losses globally. Different insect species have different courtship responses and phototropic flight rhythms. Improving the trapping [...] Read more.
Tuta absoluta (Meyrick), a leafminer that damages tomato leaves, terminal buds, flowers, and fruits, is a destructive tomato pest and is responsible for 80–100% of tomato yield losses globally. Different insect species have different courtship responses and phototropic flight rhythms. Improving the trapping effects of the sex pheromone and light traps is important for constructing an IPM system for T. absoluta. The present study explored the hourly distribution of T. absoluta adults caught by the sex pheromone (on the ground) and UV light (380 nm) traps in greenhouses over 24 h. The responses of males to sex pheromone (false female) lures were detected at dawn and early morning. The responses lasted for 3 h, from 05:30 (1 h before sunrise) to 08:30 (2 h after sunrise), and 95.8% of the males were caught during this period. The peak of the male responses to the sex pheromone was detected at 07:30 (from 06:30 to 07:30, 1 h after sunrise), and 80.8% of the males were caught during this period. The flight of male (proportion of 54.3%) and female (45.7%) adults toward the UV light traps occurred from 19:30 (time of sunset) to 06:30 (time of sunrise), lasted for 11 h, and exhibited a scotophase rhythm; 97.4% of the adults were caught during this period. The peak of adults flying toward the UV light traps occurred at 21:30 (from 20:30 to 21:30, 2 h after sunset). The rhythms of males’ responses to the sex pheromone and of the adults’ flight toward the UV lights can help to reveal the mechanisms of chemotactic and phototactic responses and may play a significant role in constructing an IPM system for this pest. Full article
Show Figures

Figure 1

Figure 1
<p>Hourly distribution of <span class="html-italic">Tuta absoluta</span> male adults responding to sex pheromone lures. Different lowercase letters indicate significant differences among different hourly periods at the <span class="html-italic">p</span> &lt; 0.05 significance level (one-way ANOVA and LSD test). The up and down arrows indicate sunrise and sunset, respectively. (<b>A</b>), hourly distribution of numbers; (<b>B</b>), hourly distribution of proportions. The same applies in <a href="#horticulturae-09-00402-f002" class="html-fig">Figure 2</a> below.</p>
Full article ">Figure 2
<p>Hourly distribution of total <span class="html-italic">Tuta absoluta</span> adults flying toward UV light traps. Different lowercase letters indicate significant differences among different hourly periods at the <span class="html-italic">p</span> &lt; 0.05 significance level (one-way ANOVA and LSD test). The up and down arrows indicate sunrise and sunset, respectively. (<b>A</b>), hourly distribution of numbers; (<b>B</b>), hourly distribution of proportions.</p>
Full article ">Figure 3
<p>Hourly distributions of male and female <span class="html-italic">Tuta absoluta</span> adults flying toward UV light traps. Different lowercase Latin and Greek letters indicate significant differences among hourly periods for male and female adults, respectively, at the <span class="html-italic">p</span> &lt; 0.05 significance level (one-way ANOVA and LSD test). The up and down arrows indicate sunrise and sunset, respectively. (<b>A</b>), hourly distribution of numbers; (<b>B</b>), hourly distribution of proportions. The asterisk above the horizontal line indicates a significant difference between males and females during the same period at the <span class="html-italic">p</span> &lt; 0.05 significance level (paired-sample <span class="html-italic">t</span>-test); ns, not significant.</p>
Full article ">
15 pages, 640 KiB  
Review
The Role of Light-Regulated Auxin Signaling in Root Development
by Fahong Yun, Huwei Liu, Yuzheng Deng, Xuemei Hou and Weibiao Liao
Int. J. Mol. Sci. 2023, 24(6), 5253; https://doi.org/10.3390/ijms24065253 - 9 Mar 2023
Cited by 8 | Viewed by 5180
Abstract
The root is an important organ for obtaining nutrients and absorbing water and carbohydrates, and it depends on various endogenous and external environmental stimulations such as light, temperature, water, plant hormones, and metabolic constituents. Auxin, as an essential plant hormone, can mediate rooting [...] Read more.
The root is an important organ for obtaining nutrients and absorbing water and carbohydrates, and it depends on various endogenous and external environmental stimulations such as light, temperature, water, plant hormones, and metabolic constituents. Auxin, as an essential plant hormone, can mediate rooting under different light treatments. Therefore, this review focuses on summarizing the functions and mechanisms of light-regulated auxin signaling in root development. Some light-response components such as phytochromes (PHYs), cryptochromes (CRYs), phototropins (PHOTs), phytochrome-interacting factors (PIFs) and constitutive photo-morphorgenic 1 (COP1) regulate root development. Moreover, light mediates the primary root, lateral root, adventitious root, root hair, rhizoid, and seminal and crown root development via the auxin signaling transduction pathway. Additionally, the effect of light through the auxin signal on root negative phototropism, gravitropism, root greening and the root branching of plants is also illustrated. The review also summarizes diverse light target genes in response to auxin signaling during rooting. We conclude that the mechanism of light-mediated root development via auxin signaling is complex, and it mainly concerns in the differences in plant species, such as barley (Hordeum vulgare L.) and wheat (Triticum aestivum L.), changes of transcript levels and endogenous IAA content. Hence, the effect of light-involved auxin signaling on root growth and development is definitely a hot issue to explore in the horticultural studies now and in the future. Full article
(This article belongs to the Special Issue Cellular and Molecular Mechanisms of Plant Responses to Light)
Show Figures

Figure 1

Figure 1
<p>(HY5), UV resistance locus 8 (UVR8), phototropins (PHOTs), phytochromes (PHYs) and cryptochromes (CRYs) are regarded as photoreceptors with a response to light, and they act as the upstream of auxin-signaling-related genes or proteins. PHOTs could interact with PHYTOCHROME KINASE SUBSTRATE 1 (PKS1), ROOT PHOTOTROPISM 2 (RPT2) and NON-PHOTOTROPIC HYPOCOTYL 3 (NPH3), and act as the upstream of the auxin-signaling-related genes PINs. The MYB domain protein 73/77 (MYB73/MYB77) acts as the downstream of UVR8 to regulate auxin responses and lateral root growth. In the auxin-signaling transduction pathway, PIN-formed (<span class="html-italic">PINs</span>), <span class="html-italic">YUCs</span>, <span class="html-italic">auxin response factors</span> (<span class="html-italic">ARFs</span>) and LIKE AUX3 (<span class="html-italic">LAX3</span>) are key genes, and act as the downstream of light-response-related genes or proteins. When a plant is exposed to light, miR775 could induce the root growth and development by regulating root-hair-related genes such as ROOTHAIR DEFECTIVE SIX-LIKE (<span class="html-italic">RSLs</span>) and <span class="html-italic">PROTEIN PHOSPHATASE 2A</span> (<span class="html-italic">PP2A</span>). MEDIATOR18 (MED18) could promote root growth with continuous light via directly regulating auxin-signaling transduction-related genes. Moreover, MED18 triggered the transcript level of <span class="html-italic">HY5</span>, which led to chlorophyll accumulation. When a plant is exposed to light, HY1 induces the up-regulation of HY5 and HYH, then HY5 triggers the expression of auxin-signaling-related genes IAAs, which finally promotes root branching. Furthermore, HY1 can directly regulate root branching via enhancing the transcript level of PINs and AUX1. The arrows mean that one protein or stimuli could directly and positively act as the upstream to regulate another protein. The double arrows represent interactions between proteins. The “T” shape indicates that one stimuli could negatively regulate the protein.</p>
Full article ">
23 pages, 3620 KiB  
Article
Evaluation of the Cleaning Effect of Natural-Based Biocides: Application on Different Phototropic Biofilms Colonizing the Same Granite Wall
by Chiara Genova, Elsa Fuentes, Gabriele Favero and Beatriz Prieto
Coatings 2023, 13(3), 520; https://doi.org/10.3390/coatings13030520 - 26 Feb 2023
Cited by 8 | Viewed by 2373
Abstract
Natural derivatives, such as essential oils, are presented as an alternative to classical biocides to the treatment of biocolonization. Thus, in this work, the cleaning and biocidal potential of some natural derivatives towards two natural biofilms’ growth on the same granite wall, with [...] Read more.
Natural derivatives, such as essential oils, are presented as an alternative to classical biocides to the treatment of biocolonization. Thus, in this work, the cleaning and biocidal potential of some natural derivatives towards two natural biofilms’ growth on the same granite wall, with different microbial composition, was evaluated. For this purpose, three essential oils (EOs) (from Origanum vulgare, Thymus vulgaris and Calamintha nepeta) and their main active principles (APs) (carvacrol, thymol and R-(+)-pulegone, respectively) were embedded in a hydrogel matrix, with different combinations of EOs and APs, in order to evaluate the synergistic action of different actives. For comparative purposes, pure hydrogel and a mechanical method (brushing) were also used. Colorimetric measurements and chlorophyll a fluorescence analyses were performed to evaluate the cleaning action of the treatments on the biofilms. Overall, the EOs and APs present in the hydrogel proved to be reliable treatments to limit natural biocolonization, with O. vulgare being one of the most effective treatments in combination with other compounds, due to the majority presence of carvacrol. Moreover, the effect of the different treatments strictly depended on the biofilm in question, as well as its ability to adhere to the substrate. Full article
(This article belongs to the Special Issue Coatings on Built Heritage and New Build)
Show Figures

Figure 1

Figure 1
<p>Experimental surfaces (details). In the pictures, the two surfaces are compared. In (<b>a</b>) the dark green biopatina is shown (Surface a), and in (<b>b</b>) the green-orange biopatina is shown (Surface b).</p>
Full article ">Figure 2
<p>(<b>a</b>) Surface a; (<b>b</b>) Surface b. Composition of each treatment: T1 = <span class="html-italic">O. vulgare</span>; T2 = <span class="html-italic">T. vulgaris</span>; T3 = <span class="html-italic">C. nepeta</span>; T4 = Carvacrol; T5 = Thymol; T6 = Pulegone; T7 = <span class="html-italic">O. vulgare</span> + <span class="html-italic">T. vulgaris</span>; T8 = <span class="html-italic">O. vulgare</span> + <span class="html-italic">C. nepeta</span>; T9 = <span class="html-italic">C. nepeta</span> + <span class="html-italic">T. vulgaris</span>; T10 = <span class="html-italic">O. vulgare</span> + <span class="html-italic">T. vulgaris</span> + <span class="html-italic">C. nepeta</span>; T11 = Carvacrol + Thymol; T12 = Carvacrol + Pulegone; T13 = Pulegone + Thymol; T14 = Carvacrol + Thymol + Pulegone; T15 = Hydrogel; T16 = Brushing. The orange rectangles individuate the zones where the biological material was collected.</p>
Full article ">Figure 3
<p>Schematic representation of the treatments employed and their composition.</p>
Full article ">Figure 4
<p>Taxa identified in sampled subaerial biofilms: (<b>A</b>,<b>B</b>) protonema of bryophyta; (<b>C</b>,<b>D</b>) <span class="html-italic">Trentepohlia aurea</span> (Linnaeus) C. Martius; (<b>E</b>,<b>F</b>) <span class="html-italic">Mesotaenium macrococcum</span> (Kützing ex Kützing) J. Roy &amp; Bisset; (<b>G</b>–<b>I</b>) <span class="html-italic">Apatococcus lobatus</span> (Chodat) J.B.Petersen: diagnostic detail of the cells with bilobate chloroplast without pyrenoid indicate by arrow (<b>I</b>);(<b>J</b>,<b>K</b>) <span class="html-italic">Desmococcus olivaceus</span> (Persoon ex Acharius) J. R. Laundon; (<b>L</b>) <span class="html-italic">Klebsormidium flaccidum</span> (Kützing) P. C. Silva, K. R. Mattox &amp; W. H. Blackwell; (<b>N</b>) <span class="html-italic">Oscillatoria formosa</span> Bory ex Gomont; (<b>O</b>) <span class="html-italic">Nostoc</span> sp. Vaucher ex Bornet &amp; Flahault trichomes without mucilage formed by subspherical cells of 2.0-4.5 µm in diameter and heterocytes indicated by an arrow; (<b>P</b>) <span class="html-italic">Gloeocapsa punctata</span> Nägeli; (<b>Q</b>–<b>S</b>) <span class="html-italic">Hantzschia amphioxys</span> (Ehrenberg) Grunow. Scale bar = 50 µm (<b>A</b>,<b>C</b>,<b>G</b>); 20 µm (<b>E</b>,<b>H</b>); 10 µm (<b>B</b>,<b>D</b>); 5 µm (<b>F</b>,<b>I</b>,<b>K</b>,<b>L</b>–<b>S</b>).</p>
Full article ">Figure 5
<p>LB-carbohydrates present in biofilm patinas from Surface a and Surface b.</p>
Full article ">Figure 6
<p>Appearance of the experimental surfaces after the cleaning procedures. (<b>a</b>) Peeling of the treatment that provided the employment of the hydrogel; (<b>b</b>) Surface a after the removal of the treatments (t<sub>1</sub>) and (<b>c</b>) Surface b after the removal of the treatments (t<sub>1</sub>).</p>
Full article ">Figure 7
<p>Color changes of the samples from Surface a and Surface b, before (t<sub>0</sub>) and after (t<sub>1</sub>) the treatment’s removal, are reported in the colorimetric plane of a* (x-axis, changes in redness-greenness) and b* (y-axis, changes in yellowness-blueness). (<b>a</b>) Color data from Surface a at t<sub>0</sub>; (<b>b</b>) color data from Surface b at t<sub>0</sub>; (<b>c</b>) color data form Surface a at t<sub>1</sub>; (<b>d</b>) color data from Surface b at t<sub>1</sub>. In each graph, the color of the reference value for the uncolonized granite was also represented (red spot). Each number represents the corresponding treatment (codes are shown in <a href="#coatings-13-00520-f003" class="html-fig">Figure 3</a>), and the letters (a and b) correspond to the surface where the treatments have been applied.</p>
Full article ">Figure 8
<p>Relative differences (ΔY) of maximum quantum yield (Y) calculated between t<sub>0</sub> and t<sub>1</sub> for (<b>a</b>) Surface a, and (<b>b</b>) Surface b. ∆Y &gt; 0 establishes the increase in the vital activity after the application of the treatments; ∆Y &lt; 0 is the decrease in the vital activity.</p>
Full article ">
13 pages, 3887 KiB  
Communication
Evaluation of Abnormal Hypocotyl Growth of Mutant Capsicum annuum Plants
by Bánk Pápai, Zsófia Kovács, Kitti Andrea Tóth-Lencsés, Janka Bedő, Gábor Csilléry, Anikó Veres and Antal Szőke
Agriculture 2023, 13(2), 481; https://doi.org/10.3390/agriculture13020481 - 17 Feb 2023
Cited by 1 | Viewed by 2093
Abstract
Horticulture is a dynamically evolving and an ever-changing sector which needs new ideas, plant materials, and cultivating methods to produce more. Involving different mutants in breeding lines may lead to new opportunities to create new cultivating methods. pcx (procumbent plant) and tti (tortuosa [...] Read more.
Horticulture is a dynamically evolving and an ever-changing sector which needs new ideas, plant materials, and cultivating methods to produce more. Involving different mutants in breeding lines may lead to new opportunities to create new cultivating methods. pcx (procumbent plant) and tti (tortuosa internodi) Capsicum annuum mutant plants, which present abnormal stem growth, were investigated in various in vitro experiments. The pcx breeding line presents highly diverse hypocotyl growth even in the early phenophase, such as normally growing plants and the ‘laying’ habit. On the other hand, tti plants only present their elongated slender stem trait in a more mature phase. In our experiment of reorientation, we used one-sided illumination, where each of the phenotypes sensed and reacted to light, and only the pcx plants exhibited a negative gravitropic response. It was also the result that the tti plants sensed gravity, but the weak structure of the hypocotyls made them incapable of following its direction. Since the pcx plants were the only ones with an ‘antigravitropic’ growth, we used them to evaluate the time course they needed to adapt and follow the gravity vector after reorientation. The pcx plants sensing gravity adapted similarly to controls and started bending after 120 min, but those which presented as ‘anti-gravitropic’ did not respond even after 420 min. Full article
Show Figures

Figure 1

Figure 1
<p>The <span class="html-italic">pcx</span> (<b>right</b>) and <span class="html-italic">tti</span> (<b>left</b>) plants grown in greenhouse in Szentes, Hungary.</p>
Full article ">Figure 2
<p>Various stem growth habit of the <span class="html-italic">pcx</span> plants.</p>
Full article ">Figure 3
<p>Method used to evaluate graviresponse in seedlings Straight seedlings were placed on solid plant media (<b>A</b>), which was then rotated 90° (<b>B</b>). Responding seedlings reoriented to resume vertical growth (<b>C</b>). Measurements taken included hypocotyl curvature. α, hypocotyl curvature; θ, root curvature; g, direction of the gravity vector. After 24 h, the hypocotyl curvatures were checked (°) (<b>D</b>) [<a href="#B49-agriculture-13-00481" class="html-bibr">49</a>].</p>
Full article ">Figure 4
<p>Two- (<b>A</b>) and three-week-old (<b>B</b>) in vitro seedlings of <span class="html-italic">pcx</span>, <span class="html-italic">tti</span> and control plants.</p>
Full article ">Figure 5
<p>Curvature of the hypocotyl base of <span class="html-italic">pcx</span>, <span class="html-italic">tti</span> and control <span class="html-italic">Capsicum annuum</span> seedlings after being grown for two (<b>A</b>) and three weeks (<b>B</b>) under in vitro conditions. Plant height is measured in cm and the curvature was measured every 0.1 cm from the hypocotyl base. n = 30.</p>
Full article ">Figure 6
<p>Curvature of the hypocotyls of <span class="html-italic">pcx</span>, <span class="html-italic">tti</span> and control <span class="html-italic">Capsicum annuum</span> three-week-old seedlings 24 h after reorientation by 90° under in vitro conditions. Plant length is measured in cm and the curvature was measured every 0.1 cm from the hypocotyl base. n = 30.</p>
Full article ">Figure 7
<p>Photo- and gravitropic response of three-week-old <span class="html-italic">pcx, control</span> and <span class="html-italic">tti</span> plants reoriented by 90 degrees and documented after 24 h under in vitro conditions.</p>
Full article ">Figure 8
<p>Gravitropic responses of the hypocotyls of three-week-old seedlings, documented every hour after a horizontal reorientation in the case of control and <span class="html-italic">pcx</span> plants.</p>
Full article ">Figure 9
<p>Time course of the gravitropic response of the hypocotyls of <span class="html-italic">pcx</span> and control <span class="html-italic">Capsicum annuum</span> three-week-old seedlings reoriented by 90°, documented every hour under in vitro conditions. Hypocotyl curvature was measured in °. n = 11.</p>
Full article ">
19 pages, 5685 KiB  
Article
Environmental Contamination with Micro- and Nanoplastics Changes the Phototaxis of Euryhaline Zooplankton to Paired Photostimulation
by Yuri Morgalev, Victor Dyomin, Sergey Morgalev, Alexandra Davydova, Tamara Morgaleva, Oksana Kondratova, Igor Polovtsev, Nikolay Kirillov and Alexey Olshukov
Water 2022, 14(23), 3918; https://doi.org/10.3390/w14233918 - 1 Dec 2022
Cited by 8 | Viewed by 1876
Abstract
Our earlier studies showed that paired photostimulation allows the detection of pollutants in an aqueous medium according to the behavioral responses of freshwater Crustacea. The first stimulus initiated and stabilized the behavioral response. The increase in response to the second stimulus made [...] Read more.
Our earlier studies showed that paired photostimulation allows the detection of pollutants in an aqueous medium according to the behavioral responses of freshwater Crustacea. The first stimulus initiated and stabilized the behavioral response. The increase in response to the second stimulus made it possible to assess the responsiveness of the zooplankton community. This paper studies the validity of this method for the detection of micro- and nanoplastic contamination of saltwater reservoirs according to the behavioral response of Artemia salina and Moina salina crustaceans. The studies were conducted in laboratory conditions using a submersible holographic camera developed by us, which ensures the in situ detection of the concentration and speed of crustaceans in a volume of up to 1 dm3, as well as makes it possible to change the intensity and duration of the attracting light. It was established that the phototropic response of crustaceans decreases in seawater at the cumulative dose of exposure to microplastics—0.15 mg∙dm−3∙h and nanoplastics—0.3 mg∙dm−3∙h. The paired photostimulation reveals the altering effect of micro- and nanoplastics in the saltwater medium no later than 3 h after their appearance, which indicates the promising potential of this method for the alarm response in monitoring the environmental well-being of water bodies. Full article
Show Figures

Figure 1

Figure 1
<p>Picture of microplastic particles received using a confocal microscope and (<b>a</b>) content of ≤10 px and &gt;10 px particles in samples (<b>b</b>): (<b>a</b>) Microplastic (green) particles are marked with numbers.</p>
Full article ">Figure 2
<p>DHC-based laboratory unit and (<b>a</b>) photo of light columns with hydrobionts (<b>b</b>): (<b>a</b>): 1—DHC, 2—DHC recording unit, 3—DHC lighting module, 4—laboratory water tank, 5—test (working) volume formed by recording (red) and attracting (green) light beams, 6—mirror-prism system for working volume forming, 7—semiconductor laser diode (λ = 650 nm), 8—semiconductor laser diode (λ = 532 nm), 9—fiber-optic multiplexer (mixer), 10—beam expander, 11—windows, 12—selective filter, 13—receiving lens, 14—CMOS camera.</p>
Full article ">Figure 3
<p>Biosystem response scheme (<b>a</b>) and difference in accuracy of system dynamics prediction (<b>b</b>) during paired photostimulation. Cv—coefficient of variation.</p>
Full article ">Figure 4
<p>Response of <span class="html-italic">A. salina</span> to paired photostimulation in the presence of K<sub>2</sub>Cr<sub>2</sub>O<sub>7</sub> at the concentration of 4.0 mg/dm<sup>3</sup> (<b>a</b>), 8.0 mg/dm<sup>3</sup> (<b>b</b>), 16.0 mg/dm<sup>3</sup> (<b>c</b>). C<sub>0</sub>—average concentration of crustaceans before photostimulation; C<sub>1</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>1</sub> (1150 lx); C<sub>2</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>2</sub> (3450 lx); Bv—background value; Green arrow—pollution.</p>
Full article ">Figure 5
<p>Approximation of ΔC/C<sub>2</sub> dynamics and (<b>a</b>) the slope of fitting lines (<b>b</b>) under the influence of various concentrations of K<sub>2</sub>Cr<sub>2</sub>O<sub>7</sub> on <span class="html-italic">A. salina</span>.</p>
Full article ">Figure 6
<p>Response of <span class="html-italic">A. salina</span> to paired photostimulation in the presence of nanoplastics at the concentration of 0.1 mg/dm<sup>3</sup> (<b>a</b>), 1.0 mg/dm<sup>3</sup> (<b>b</b>). C<sub>0</sub>—average concentration of crustaceans before photostimulation; C<sub>1</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>1</sub> (1150 lx); C<sub>2</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>2</sub> (3450 lx); Bv—background value; Green arrow—pollution.</p>
Full article ">Figure 7
<p>Approximation of ΔC/C<sub>2</sub> dynamics and (<b>a</b>) the slope of fitting lines (<b>b</b>) under the influence of various concentrations of nanoplastics on <span class="html-italic">A. salina</span>.</p>
Full article ">Figure 8
<p>Response of <span class="html-italic">M. salina</span> to paired photostimulation in the presence of nanoplastics at the concentration of 0.1 mg/dm<sup>3</sup> (<b>a</b>), 1.0 mg/dm<sup>3</sup> (<b>b</b>). C<sub>0</sub>—average concentration of crustaceans before photostimulation; C<sub>1</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>1</sub> (1150 lx); C<sub>2</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>2</sub> (3450 lx); Bv—background value; Green arrow—pollution.</p>
Full article ">Figure 9
<p>Approximation of ΔC/C<sub>2</sub> dynamics and (<b>a</b>) the slope of fitting lines (<b>b</b>) under the influence of various concentrations of nanoplastics on <span class="html-italic">M. salina</span>.</p>
Full article ">Figure 10
<p>Response of <span class="html-italic">A. salina</span> to paired photostimulation in the presence of microplastics at the concentration of 0.05 mg/dm<sup>3</sup> (<b>a</b>), 0.1 mg/dm<sup>3</sup> (<b>b</b>), 1.0 mg/dm<sup>3</sup> (<b>c</b>), C<sub>0</sub>—average concentration of crustaceans before photostimulation; C<sub>1</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>1</sub> (1150 lx); C<sub>2</sub>—average concentration of crustaceans at photostimulation with intensity I<sub>2</sub> (3450 lx) ; Bv—background value; Green arrow—pollution.</p>
Full article ">Figure 11
<p>Approximation of the dynamics of the ΔC/C<sub>2</sub> ratio (<b>a</b>) and the slope of fitting lines (<b>b</b>) under the influence of various concentrations of microplastics on <span class="html-italic">A. salina</span>.</p>
Full article ">
21 pages, 4136 KiB  
Article
Aux/IAA11 Is Required for UV-AB Tolerance and Auxin Sensing in Arabidopsis thaliana
by Jakub Mielecki, Piotr Gawroński and Stanisław Karpiński
Int. J. Mol. Sci. 2022, 23(21), 13386; https://doi.org/10.3390/ijms232113386 - 2 Nov 2022
Cited by 6 | Viewed by 2717
Abstract
In order to survive, plants have, over the course of their evolution, developed sophisticated acclimation and defense strategies governed by complex molecular and physiological, and cellular and extracellular, signaling pathways. They are also able to respond to various stimuli in the form of [...] Read more.
In order to survive, plants have, over the course of their evolution, developed sophisticated acclimation and defense strategies governed by complex molecular and physiological, and cellular and extracellular, signaling pathways. They are also able to respond to various stimuli in the form of tropisms; for example, phototropism or gravitropism. All of these retrograde and anterograde signaling pathways are controlled and regulated by waves of reactive oxygen species (ROS), electrical signals, calcium, and hormones, e.g., auxins. Auxins are key phytohormones involved in the regulation of plant growth and development. Acclimation responses, which include programmed cell death induction, require precise auxin perception. However, our knowledge of these pathways is limited. The Aux/IAA family of transcriptional corepressors inhibits the growth of the plant under stress conditions, in order to maintain the balance between development and acclimation responses. In this work, we demonstrate the Aux/IAA11 involvement in auxin sensing, survival, and acclimation to UV-AB, and in carrying out photosynthesis under inhibitory conditions. The tested iaa11 mutants were more susceptible to UV-AB, photosynthetic electron transport (PET) inhibitor, and synthetic endogenous auxin. Among the tested conditions, Aux/IAA11 was not repressed by excess light stress, exclusively among its phylogenetic clade. Repression of transcription by Aux/IAA11 could be important for the inhibition of ROS formation or efficiency of ROS scavenging. We also hypothesize that the demonstrated differences in the subcellular localization of the two Aux/IAA11 protein variants might indicate their regulation by alternative splicing. Our results suggest that Aux/IAA11 plays a specific role in chloroplast retrograde signaling, since it is not repressed by high (excess) light stress, exclusively among its phylogenetic clade. Full article
(This article belongs to the Special Issue New Insight into Signaling and Autophagy in Plants 2.0)
Show Figures

Figure 1

Figure 1
<p>Identification of T-DNA mutants in <span class="html-italic">Aux/IAA11</span> gene, and phylogenetic analysis of Aux/IAAs. (<b>A</b>) Splicing variants of <span class="html-italic">Aux/IAA11</span> gene. Black vertical lines represent the insertion of T-DNA in <span class="html-italic">iaa11-1</span>, <span class="html-italic">iaa11-2</span>, <span class="html-italic">iaa11-3</span>, and <span class="html-italic">iaa11-4</span> mutant lines. The red rectangle indicates the localization of the transmembrane (TM) domain in <span class="html-italic">Aux/IAA11.2.</span> Black half arrows illustrate binding sites of the primers used for qPCR. (<b>B</b>) Cladogram of protein similarities. <span class="html-italic">Aux/IAA11</span> clade was prepared using the ClustalW alignment algorithm and the maximum likelihood statistical method in Molecular Evolutionary Genetics Analysis (MEGA) 11 software [<a href="#B60-ijms-23-13386" class="html-bibr">60</a>]. (<b>C</b>) Relative expression of <span class="html-italic">Aux/IAA11</span> in tested mutant lines. (<b>D</b>) Relative expression of <span class="html-italic">Aux/IAA10</span>, <span class="html-italic">Aux/IAA12</span>, and <span class="html-italic">Aux/IAA13</span> in <span class="html-italic">iaa11-1</span> mutant background. Relative gene expressions were calculated and normalized with <span class="html-italic">UBIQUITIN-PROTEIN LIGASE 7 (UPL7</span>) and <span class="html-italic">YELLOW-LEAF-SPECIFIC GENE 8</span> (<span class="html-italic">YLS8</span>) as references, using the Relative Expression Software Tool (REST 2009) (n = 6, *** <span class="html-italic">p</span> &lt; 0.001). Error bars represent the standard deviation.</p>
Full article ">Figure 2
<p>Promoter analysis and light induction of the <span class="html-italic">Aux/IAA11</span> genetic clade. (<b>A</b>) Location of light responsive elements (LREs), auxin responsive elements (AuxREs), and abscisic acid responsive elements (ABREs) in promoter sequences 2000 bp upstream from the translation start site (TSS) of <span class="html-italic">Aux/IAA10</span>, <span class="html-italic">Aux/IAA11</span>, <span class="html-italic">Aux/IAA12</span>, and <span class="html-italic">Aux/IAA13</span> genes. (<b>B</b>) Expression of the <span class="html-italic">Aux/IAA11</span> phylogenetic clade in response to high light stress. Relative gene expressions were calculated and normalized with <span class="html-italic">UBIQUITIN-PROTEIN LIGASE 7</span> (<span class="html-italic">UPL7</span>) and <span class="html-italic">YELLOW-LEAF-SPECIFIC GENE 8</span> (<span class="html-italic">YLS8</span>) as references, using Relative Expression Software Tool (REST 2009) (n = 6, * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001). Error bars represent the standard deviation.</p>
Full article ">Figure 3
<p>UV-AB susceptibility of <span class="html-italic">Aux/IAA11</span> mutants. (<b>A</b>) Photographs of tested <span class="html-italic">Arabidopsis thaliana</span> plants before, and 3 and 7 days after, the UV stress exposure. (<b>B</b>) Relative ion leakage measured 3 days after UV stress exposure. n = 9, Dunnett’s test (*** <span class="html-italic">p</span> &lt; 0.001). Error bars represent the standard deviation.</p>
Full article ">Figure 4
<p>3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) susceptibility of <span class="html-italic">Aux/IAA11</span> mutants. (<b>A</b>) Maximum quantum efficiency of photosystem II (<span class="html-italic">F</span><sub>v</sub>/<span class="html-italic">F</span><sub>m</sub>) after 0, 1, 2, and 3 h of 60 µM DCMU treatment (n = 9). (<b>B</b>) <span class="html-italic">F</span><sub>v</sub>/<span class="html-italic">F</span><sub>m</sub> and (<b>C</b>) H<sub>2</sub>O<sub>2</sub> content measured after 15 min of 200 µM DCMU treatment. n ≥ 6, Dunnett’s test (* <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001). Error bars represent the standard deviation.</p>
Full article ">Figure 5
<p>Visualization of singlet oxygen sensor green (SOSG). Images were taken 2 and 5 h after a 15 min treatment with 200 µM DCMU and SOSG staining. Green and red colors represent SOSG and chlorophyll fluorescence, respectively. White scale bars represent 50 µm.</p>
Full article ">Figure 6
<p><span class="html-italic">Aux/IAA11</span> is involved in auxin sensing. (<b>A</b>) Phenotype of 7-day-old <span class="html-italic">Aux/IAA11</span> mutants and <span class="html-italic">auxinresistant2</span> (<span class="html-italic">axr2</span>) after treatment with 2,4-Dichlorophenoxyacetic acid (2,4-D). (<b>B</b>) Main root length of plants grown on ½ MS and on ½ MS supplemented with 10 nM of 2,4-D. (<b>C</b>) Reduction of the main root length of plants grown on 2,4-D compared to control conditions. n ≥ 6, Dunnet’s test (* <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001). Root lengths were measured using ImageJ software. Error bars represent standard deviation.</p>
Full article ">Figure 7
<p>Subcellular localization of IAA11 protein variants in <span class="html-italic">Allium cepa</span>. Transient expression of fusion proteins: YFP::IAA11.1 and YFP:IAA11.2 in <span class="html-italic">Allium cepa</span> adaxial epidermis under 35S constitutive promoter. Pictures were taken 72 h post agrobacterium infiltration using confocal microscopy. Magenta and cyan colors represent Yellow fluorescent protein (YFP) and 4′,6-Diamidino-2-Phenylindole, Dihydrochloride (DAPI) fluorescence, respectively. Red scale bars represent 50 µm.</p>
Full article ">
Back to TopTop