[go: up one dir, main page]

 
 
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
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (770)

Search Parameters:
Keywords = xylem

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 17756 KiB  
Article
PlLAC15 Facilitates Syringyl Lignin Deposition to Enhance Stem Strength in Herbaceous Peony
by Yuehan Yin, Shiqi Zuo, Minghao Zhao, Jun Tao, Daqiu Zhao and Yuhan Tang
Agriculture 2024, 14(9), 1609; https://doi.org/10.3390/agriculture14091609 (registering DOI) - 14 Sep 2024
Viewed by 209
Abstract
Stems are prone to bending or lodging due to inadequate stem strength, which seriously reduces the cut-flower ornamental quality of herbaceous peony (Paeonia lactiflora Pall.). Plant LACCASE (LAC), a copper-containing polyphenol oxidase, has been shown to participate in the polymerization process of [...] Read more.
Stems are prone to bending or lodging due to inadequate stem strength, which seriously reduces the cut-flower ornamental quality of herbaceous peony (Paeonia lactiflora Pall.). Plant LACCASE (LAC), a copper-containing polyphenol oxidase, has been shown to participate in the polymerization process of monolignols; however, the role of LAC in regulating the stem strength of P. lactiflora remains unclear. Here, the full-length cDNA of PlLAC15, which demonstrated a positive association with stem strength, was isolated. It consisted of 1790 nucleotides, encoding 565 amino acids that had four typical laccase copper ion-binding domains. Moreover, PlLAC15 was highly expressed in the stem, and its expression level gradually significantly increased during stem development. Furthermore, PlLAC15 was found to be localized specifically to the cell wall, and its recombinant protein exhibited laccase activity. Additionally, the role of PlLAC15 in regulating the stem strength of P. lactiflora was confirmed by transgenic studies. When PlLAC15 was overexpressed in tobacco, stem strength increased by more than 50%, S-lignin was significantly deposited, and the lignification degree of stem xylem fiber cells increased. These results suggested that PlLAC15 facilitated S-lignin deposition to enhance stem strength in P. lactiflora, which would provide precious information that benefits future exploration of stem bending or lodging resistance in plants. Full article
Show Figures

Figure 1

Figure 1
<p>Characterization of PlLAC15. (<b>a</b>) Venn diagram of differently expressed <span class="html-italic">LACs</span> in upright ‘Hong Feng’ and bending ‘Xixia Yingxue’ in <span class="html-italic">P. lactiflora</span> across three flower developmental stages from previous transcriptomic data [<a href="#B19-agriculture-14-01609" class="html-bibr">19</a>]; (<b>b</b>) expression patterns of four nonredundant <span class="html-italic">LACs</span> from (A), and the relationship between FPKM value of i2_HQ_PL_c36823 and stem strength; (<b>c</b>) phylogenetic analysis of LACs from <span class="html-italic">P. lactiflora</span> (orange) and <span class="html-italic">A. thaliana</span> (black); (<b>d</b>) sequence alignment of PlLAC15, AtLAC15, and AtLAC14. The laccase copper ion-binding domains are indicated by lines, respectively. ‘HF’, ‘Hong Feng’; ‘XX’, ‘Xixia Yingyue’; S1, flower-bud stage; S2, unfolded-petal stage; S3, full-bloom stage.</p>
Full article ">Figure 2
<p>Expression patterns of <span class="html-italic">PlLAC15</span>. (<b>a</b>) Expression heatmap of <span class="html-italic">PlLAC15</span> in different tissues of <span class="html-italic">P. lactiflora</span> at full-blooming period; (<b>b</b>) expression patterns of <span class="html-italic">PlLAC15</span> in stems at four different developmental periods (P1–P4) [<a href="#B20-agriculture-14-01609" class="html-bibr">20</a>]. Tukey’s HSD test was used for statistical analyses, means ± SD, letters indicated significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Subcellular localization of PlLAC15. Constructs of p35S:PlLAC15-eGFP or p35S-eGFP combined with p35S:AtEXPA1-mCherry (cell wall marker) were transferred to <span class="html-italic">N. benthamiana</span> leaves by agroinfiltration. eGFP signals, green; mCherry signals, red.</p>
Full article ">Figure 4
<p>Enzyme assays of the recombinant PlLAC15. (<b>a</b>) The purified recombinant PlLAC15 protein was analyzed by SDS-PAGE; (<b>b</b>) quantification of PlLAC15 activity in <span class="html-italic">E. coli</span>, with ABTS as the substrate. Student’s <span class="html-italic">t</span>-test was used for statistical analyses, means ± SD, ‘**’ indicated significant differences (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>Effects of changes on phenotypes and stem characteristics in <span class="html-italic">PlLAC15</span> overexpression lines. (<b>a</b>) Effects of changes on phenotypes in <span class="html-italic">PlLAC15</span> overexpression lines; (<b>b</b>) PCR identification of <span class="html-italic">PlLAC15</span> overexpression in tobacco; (<b>c</b>) qRT-PCR identification of <span class="html-italic">PlLAC15</span> overexpression in tobacco; (<b>d</b>) effects of changes on stem diameter in <span class="html-italic">PlLAC15</span> overexpression lines. (<b>e</b>) Effects of changes in stem strength in <span class="html-italic">PlLAC15</span> overexpression lines. WT, wild-type tobacco. OE-L1, overexpression of tobacco line 1. OE-L2, overexpression of tobacco line 2. Tukey’s HSD test was used for statistical analyses, means ± SD, and ‘**’ indicated significant differences (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Effects of changes in the lignin and monolignols deposition of stem xylem in <span class="html-italic">PlLAC15</span> overexpression lines. (<b>a</b>) Effects of changes on the lignin deposition of stem xylem in <span class="html-italic">PlLAC15</span> overexpression lines. The lignified cell walls were stained blue–green by toluidine blue staining; (<b>b</b>) effects of changes on the monolignols deposition of stem xylem in <span class="html-italic">PlLAC15</span> overexpression lines. The S-lignin was stained red, and the G-lignin was stained yellow by the maüle method staining. WT, wild-type tobacco. OE-L1, overexpression of tobacco line 1. OE-L2, overexpression of tobacco line 2. Xv, xylem vessel; Xf, xylem fibre.</p>
Full article ">Figure 7
<p>Effects of changes in the lignin content (<b>a</b>), S-lignin content (<b>b</b>), G-lignin content (<b>c</b>), and S/G ratio (<b>d</b>) of stems in <span class="html-italic">PlLAC15</span> overexpression lines. WT, wild-type tobacco. OE-L1, overexpression of tobacco line 1. OE-L2, overexpression of tobacco line 2. Tukey’s HSD test was used for statistical analyses, means ± SD, ‘*’ indicated significant differences (<span class="html-italic">p</span> &lt; 0.05), and ‘**’ indicated significant differences (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
12 pages, 1660 KiB  
Article
Detecting Glucose in the Phloem to Quickly Define Latent Post-Fire Mortality in Pinus Trees in Northern Italy
by Niccolò Frassinelli, Claudia Cocozza, Enrico Marchi, Cristiano Foderi, Eleftherios Touloupakis, Francesco Neri, Maria Laura Traversi and Alessio Giovannelli
Fire 2024, 7(9), 315; https://doi.org/10.3390/fire7090315 - 10 Sep 2024
Viewed by 411
Abstract
Background. Wildfires may cause serious injuries to the anatomical structure of trees that can lead to tree death or long-lasting injury recovery, limiting their growth and vitality for several years. Post-fire management involves a wide range of measures aimed at recovering and restoring [...] Read more.
Background. Wildfires may cause serious injuries to the anatomical structure of trees that can lead to tree death or long-lasting injury recovery, limiting their growth and vitality for several years. Post-fire management involves a wide range of measures aimed at recovering and restoring burnt areas. Usually, the first step is “salvage logging”, i.e., the removal of irremediably injured trees. The burn severity depends on several parameters and is variable within the burnt area. For this reason, in some areas, the death of apparently healthy individuals has often been observed even after several years. This study aims to assess delayed/latent mortality by analyzing glucose like a tracer in wood by using a blood glucometer and HPLC. Results. The glucose in the phloem, cambium, and last xylem rings was measured using a glucometer developed for measuring glucose in the blood. The adopted approach detected glucose concentrations that were recognizable for different functional levels of the trees. Conclusions. The glucometer was suitable to detect the glucose in wood and phloem in order to define the death or health of the disturbed and undisturbed trees post-fire. Further investigations are required to find new solutions for a rapid evaluation of the abiotic and biotic factors that influence tree functionality in the forest. This approach will be used to predict the probability of the death of the individuals injured, which would improve the efficiency and the economy of recovery operations. Full article
Show Figures

Figure 1

Figure 1
<p>Regression analysis of the extraction volume in water (mL) and glucose concentration (mg·dL<sup>−1</sup>) measured by glucometer.</p>
Full article ">Figure 2
<p>Regression analysis of the extraction time (h) and glucose concentration (mg·dL<sup>−1</sup>). Samples at room temperature (“unrefrig”, red) or in refrigerated conditions, 4 °C (“refrig”, blue).</p>
Full article ">Figure 3
<p>Regression analysis of glucometer (mg·dL<sup>−1</sup>) and HPLC measurements (mg·dL<sup>−1</sup>). Each point represents the glucose concentration of the same woody sample measured by glucometer and HPLC after the same extraction procedure.</p>
Full article ">Figure 4
<p>Mean glucose concentrations of core samples in “dead” (D), “living” (L), and “scorched” trees (X, both sides measured). Values are shown for HPLC (cyan) and human blood glucometer (red).</p>
Full article ">Figure 5
<p>Glucose values recorded with HPLC-differences between samples collected in fire front direction “F” and in the opposite direction “C”. In box plots, dots are outliers.</p>
Full article ">Figure 6
<p>Δ range and number of plants divided by injury groups (living, dead, and unknown, in green, red, and yellow, respectively) and condition F (at <b>left</b>) and condition C (at <b>right</b>) (Image: Maria Giulia Raeli); raw data are reported in the table.</p>
Full article ">
23 pages, 6157 KiB  
Article
Stomatal and Non-Stomatal Leaf Responses during Two Sequential Water Stress Cycles in Young Coffea canephora Plants
by Danilo F. Baroni, Guilherme A. R. de Souza, Wallace de P. Bernado, Anne R. Santos, Larissa C. de S. Barcellos, Letícia F. T. Barcelos, Laísa Z. Correia, Claudio M. de Almeida, Abraão C. Verdin Filho, Weverton P. Rodrigues, José C. Ramalho, Miroslava Rakočević and Eliemar Campostrini
Stresses 2024, 4(3), 575-597; https://doi.org/10.3390/stresses4030037 - 9 Sep 2024
Viewed by 440
Abstract
Understanding the dynamics of physiological changes involved in the acclimation responses of plants after their exposure to repeated cycles of water stress is crucial to selecting resilient genotypes for regions with recurrent drought episodes. Under such background, we tried to respond to questions [...] Read more.
Understanding the dynamics of physiological changes involved in the acclimation responses of plants after their exposure to repeated cycles of water stress is crucial to selecting resilient genotypes for regions with recurrent drought episodes. Under such background, we tried to respond to questions as: (1) Are there differences in the stomatal-related and non-stomatal responses during water stress cycles in different clones of Coffea canephora Pierre ex A. Froehner? (2) Do these C. canephora clones show a different response in each of the two sequential water stress events? (3) Is one previous drought stress event sufficient to induce a kind of “memory” in C. canephora? Seven-month-old plants of two clones (’3V’ and ‘A1’, previously characterized as deeper and lesser deep root growth, respectively) were maintained well-watered (WW) or fully withholding the irrigation, inducing soil water stress (WS) until the soil matric water potential (Ψmsoil) reached ≅ −0.5 MPa (−500 kPa) at a soil depth of 500 mm. Two sequential drought events (drought-1 and drought-2) attained this Ψmsoil after 19 days and were followed by soil rewatering until a complete recovery of leaf net CO2 assimilation rate (Anet) during the recovery-1 and recovery-2 events. The leaf gas exchange, chlorophyll a fluorescence, and leaf reflectance parameters were measured in six-day frequency, while the leaf anatomy was examined only at the end of the second drought cycle. In both drought events, the WS plants showed reduction in stomatal conductance and leaf transpiration. The reduction in internal CO2 diffusion was observed in the second drought cycle, expressed by increased thickness of spongy parenchyma in both clones. Those stomatal and anatomical traits impacted decreasing the Anet in both drought events. The ‘3V’ was less influenced by water stress than the ‘A1’ genotype in Anet, effective quantum yield in PSII photochemistry, photochemical quenching, linear electron transport rate, and photochemical reflectance index during the drought-1, but during the drought-2 event such an advantage disappeared. Such physiological genotype differences were supported by the medium xylem vessel area diminished only in ‘3V’ under WS. In both drought cycles, the recovery of all observed stomatal and non-stomatal responses was usually complete after 12 days of rewatering. The absence of photochemical impacts, namely in the maximum quantum yield of primary photochemical reactions, photosynthetic performance index, and density of reaction centers capable of QA reduction during the drought-2 event, might result from an acclimation response of the clones to WS. In the second drought cycle, the plants showed some improved responses to stress, suggesting “memory” effects as drought acclimation at a recurrent drought. Full article
(This article belongs to the Topic Plant Responses to Environmental Stress)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Soil matric water potential (Ψ<sub>msoil</sub>) at 100 mm cm and 500 mm from the soil surface in the pots of the <span class="html-italic">C. canephora</span> var. Robusta genotypes of (<b>A</b>) ‘3V’ and (<b>B</b>) ‘A1’ under well-watered (WW) and water stressed (WS) conditions. The water restriction was imposed during the drought-1 and drought-2 events, after which the soil was rewatered (and recovery-1 and recovery-2 events).</p>
Full article ">Figure 2
<p>Leaf gas exchanges of two genotypes (Gen) of <span class="html-italic">C. canephora</span> var. Robusta (‘3V’ and ‘A1’) grown under two water availability conditions [Wat, well-watered (WW) and water stress (WS)], over 12 time-points of six-day intervals (Day) during drought-1 and drought-2 and respective recovery events: (<b>A</b>) net CO<sub>2</sub> assimilation rate (<span class="html-italic">A</span><sub>net</sub>), (<b>B</b>) stomatal conductance to water (<span class="html-italic">g</span><sub>s</sub>), (<b>C</b>) transpiration rate (<span class="html-italic">E</span>), and (<b>D</b>) leaf-to-air vapor pressure deficit (VPD<sub>leaf-air</sub>). Inside the figures, the different lowercase letters indicate the significant difference among the time-points for each water regime (blue for WW and olive green for WS); different uppercase letters indicate the comparison between water availabilities for each time-point of observation (blue for WW and olive green for WS); and different superscript black ■ signs indicate that ‘3V’ was statistically superior to ‘A1’ at that time-point. Mean ± SE and ANOVA <span class="html-italic">p</span>-values (n = 7) for effects of three factors (water availability, genotype, and day of observation) and their interactions are shown. The significant <span class="html-italic">p</span>-values were marked in bold in the upper part of each graph.</p>
Full article ">Figure 3
<p>Instantaneous water-use efficiency (WUE, <span class="html-italic">A</span><sub>net</sub>/<span class="html-italic">E</span>) of two genotypes (Gen) of <span class="html-italic">C. canephora</span> var. Robusta (‘3V’ and ‘A1’) grown under two water availability conditions [Wat, well-watered (WW) and water stress (WS)], over 12 time-points of six-day intervals (Day) during drought-1 and drought-2 and respective recovery events. Inside the figure, different lowercase letters indicate the significant difference among the day-time points for each water regime (blue for WW and olive green for WS); different uppercase letters indicate the comparison between water availabilities for each time-point of observation (blue for WW and olive green for WS). Mean ± SE and ANOVA <span class="html-italic">p</span>-values (n = 7) for effects of three factors (water availability, genotype, and day of observation) and their interactions are shown. The significant <span class="html-italic">p</span>-values were marked in bold in the upper part of each graph.</p>
Full article ">Figure 4
<p>Variation of OJIP indexes of two genotypes (Gen) of <span class="html-italic">C. canephora</span> var. Robusta (‘3V’ and ‘A1’) grown under two water availability conditions [Wat, well-watered (WW) and water stress (WS)] over 12 time-points of six-day intervals (Day) during drought-1 and drought-2 and respective recovery events: (<b>A</b>) maximum quantum yield of primary photochemical reactions (ΦP<sub>0</sub>), (<b>B</b>) probability of electron transfer from Q<sub>A</sub>-to-electron transport chain beyond Q<sub>A</sub> (ΨE<sub>0</sub>), (<b>C</b>) photosynthetic performance index (PI<sub>ABS</sub>), and (<b>D</b>) density of reaction centers capable of Q<sub>A</sub> reduction (RC/CS<sub>0</sub>). Inside the figures, the different lowercase letters indicate the significant difference among the time-points for each water regime (blue for WW and olive green for WS); different uppercase letters indicate the comparison between water availabilities for each day of observation (blue for WW and olive green for WS); superscript black ■ signs indicate that ‘3V’ was statistically superior to ‘A1’, while superscript black ● signs indicate that ‘A1’ clone was statistically superior to ‘3V’ clone at that time-point. Mean ± SE and ANOVA <span class="html-italic">p</span>-values (n = 7) for effects of three factors (water availability, genotype, and day of observation) and their interactions are shown. The significant <span class="html-italic">p</span>-values were marked in bold.</p>
Full article ">Figure 5
<p>Variation of modulated chlorophyll <span class="html-italic">a</span> fluorescence indexes of two genotypes (Gen) of <span class="html-italic">C. canephora</span> var. Robusta (‘3V’ and ‘A1’) grown under two water availability conditions [Wat, well-watered (WW) and water stress (WS)] over 12 time-points of six-day intervals (Day) during drought-1 and drought-2 and respective recovery events: (<b>A</b>) effective quantum yield in PSII photochemistry (Φ<sub>PSII</sub>), (<b>B</b>) photochemical quenching (qP), (<b>C</b>) non-photochemical quenching (NPQ), and (<b>D</b>) linear electron transport rate (ETR). Inside the figures, the different lowercase letters indicate the significant difference among the time-points for each water regime (blue for WW and olive green for WS); different uppercase letters indicate the comparison between water availabilities for each day of observation (blue for WW and olive green for WS); different superscript black ■ signs indicate that ‘3V’ was statistically superior to ‘A1’, while superscript black ● signs indicate that ‘A1’ clone was statistically superior to ‘3V’ clone at that time-point. Mean ± SE and ANOVA <span class="html-italic">P</span>-values (n = 7) for effects of three factors (water availability, genotype, and day of observation) and their interactions are shown. The significant <span class="html-italic">P</span>-values were marked in bold.</p>
Full article ">Figure 6
<p>Variation of spectral reflectance indices of leaf adaxial surface of two genotypes (Gen) of C. canephora var. Robusta (‘3V’ and ‘A1’) grown under two water availability conditions [Wat, well-watered (WW) and water stress (WS)] over 12 time-points of six-day intervals (Day) during drought-1 and drought-2 and respective recovery events: (<b>A</b>) green chlorophyll index (GCI), (<b>B</b>) carotenoid reflectance index (CRI), (<b>C</b>) photochemical reflectance index (PRI), and (<b>D</b>) structure intensive reflectance index (SIPI). Inside the figures, the different lowercase letters indicate the significant difference among the time-points for each water regime (blue for WW and olive green for WS); different uppercase letters indicate the comparison between water availabilities for each day of observation (blue for WW and olive green for WS); different superscript black ■ signs indicate that ‘3V’ was statistically superior to ‘A1’, while superscript black ● signs indicate that ‘A1’ clone was statistically superior to ‘3V’ clone at that time-point. Mean ± SE and ANOVA <span class="html-italic">p</span>-values (n = 7) for effects of three factors (water availability, genotype, and day of observation) and their interactions are shown. The significant <span class="html-italic">p</span>-values were marked in bold.</p>
Full article ">Figure 7
<p>Representative area of leaf xylem vessel (µm<sup>2</sup>) measured in <span class="html-italic">C. canephora</span> var. Robusta clones (‘3V’ and ‘A1’) under well-watered (WW) and water stress (WS) conditions: (<b>A</b>) A1-WW, (<b>B</b>) 3V-WW, (<b>C</b>) A1-WS, and (<b>D</b>) 3V-WS, evaluated at the end of the second drought cycle. A scale of 100 µm is shown.</p>
Full article ">Figure 8
<p>Diagram of the two drought cycles. Transplant followed by drought-1 event last for 19 days (until −500 kPa of Ψ<sub>msoil</sub> was reached), followed by a 31-day period for a whole plant recovery (including 12-day period of recovery-1 event). The 2nd drought cycle was then applied, similarly to the 1st drought cycle, by withholding irrigation until the −500 kPa of Ψ<sub>msoil</sub> was reached (drought-2 event) and followed by another 12 days of recovery-2 event.</p>
Full article ">
18 pages, 3014 KiB  
Article
Zinc Enhances Cadmium Accumulation in Shoots of Hyperaccumulator Solanum nigrum by Improving ATP-Dependent Transport and Alleviating Toxicity
by Jia Zheng, Yukang Yue, Yuting Zhu, Yufeng Wang, Wenwen Zheng, Linfeng Hu, Dianyun Hou, Fayuan Wang, Liming Yang and Hongxiao Zhang
Plants 2024, 13(17), 2528; https://doi.org/10.3390/plants13172528 - 9 Sep 2024
Viewed by 319
Abstract
Solanum nigrum is a cadmium (Cd) and zinc (Zn) accumulator with potential for phytoextraction of soil contaminated with heavy metals. However, how Zn affects Cd accumulation in S. nigrum remains unclear. In this study, S. nigrum seedlings were treated with 100 μmol·L−1 [...] Read more.
Solanum nigrum is a cadmium (Cd) and zinc (Zn) accumulator with potential for phytoextraction of soil contaminated with heavy metals. However, how Zn affects Cd accumulation in S. nigrum remains unclear. In this study, S. nigrum seedlings were treated with 100 μmol·L−1 Zn (Zn100), 100 μmol·L−1 Cd (Cd100), and the Zn and Cd combination (Zn100+Cd100) for 10 days under hydroponic culture. Compared with Cd100, the Cd content in stems, leaves, and xylem saps was 1.8, 1.6, and 1.3 times more than that in Zn100+Cd100, respectively. In addition, the production of reactive oxygen species in leaves was significantly upregulated in Cd100 compared with the control, and it was downregulated in Zn100. Comparative analyses of transcriptomes and proteomes were conducted with S. nigrum leaves. Differentially expressed genes (DEGs) were involved in Cd uptake, transport, and sequestration, and the upregulation of some transporter genes of Zn transporters (ZIPs), a natural resistance associated macrophage protein (Nramp1), a metal–nicotianamine transporter (YSL2), ATP-binding cassette transporters (ABCs), oligopeptide transporters (OPTs), and metallothionein (MTs) and glutathione S-transferase (GSTs) genes was higher in Zn100+Cd100 than in Cd100. In addition, differentially expressed proteins (DEPs) involved in electron transport chain, ATP, and chlorophyll biosynthesis, such as malate dehydrogenases (MDHs), ATPases, and chlorophyll a/b binding proteins, were mostly upregulated in Zn100. The results indicate that Zn supplement increases Cd accumulation and tolerance in S. nigrum by upregulating ATP-dependent Cd transport and sequestration pathways. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Zn (<b>a</b>,<b>b</b>) and Cd (<b>c</b>,<b>d</b>) content in stems, leaves, xylem, and phloem saps of <span class="html-italic">S. nigrum</span>. Plants were exposed to a complete Hoagland solution (CK) or with 100 μmol·L<sup>−1</sup> Zn (Zn100), 100 μmol·L<sup>−1</sup> Cd (Cd100) and 100 μmol·L<sup>−1</sup> Zn+100 μmol·L<sup>−1</sup> Cd (Zn100+Cd100) for 10 days. Values are means ± SE (<span class="html-italic">n</span> = 3) of three different experiments. Means denoted by different letters refer to the significant differences (<span class="html-italic">p</span> &lt; 0.05, Duncan’s test).</p>
Full article ">Figure 2
<p>Production of O<sub>2</sub><sup>−</sup> (<b>a</b>,<b>b</b>) and H<sub>2</sub>O<sub>2</sub> (<b>c</b>,<b>d</b>) in leaves of <span class="html-italic">S. nigrum</span> under Zn and Cd treatment. Histochemical location of O<sub>2</sub><sup>−</sup> by NBT staining (<b>a</b>) and H<sub>2</sub>O<sub>2</sub> by DAB staining (<b>c</b>), with bar = 1 cm; O<sub>2</sub><sup>−</sup> producing rate (<b>b</b>) and H<sub>2</sub>O<sub>2</sub> content (<b>d</b>) in leaves of <span class="html-italic">S. nigrum</span>. Samples from the second youngest leaf of plants, which were exposed to a complete Hoagland solution (CK) or with 100 μmol·L<sup>−1</sup> Zn (Zn100), 100 μmol·L<sup>−1</sup> Cd (Cd100), and 100 μmol·L<sup>−1</sup> Zn+100 μmol·L<sup>−1</sup> Cd (Zn100+Cd100) for 10 days. Staining experiments were repeated at least three times, with similar results. Values are means ± SE (<span class="html-italic">n</span> = 3) of three different experiments. Means denoted by different letters refer to the significant differences (<span class="html-italic">p</span> &lt; 0.05, Duncan’s test).</p>
Full article ">Figure 3
<p>The numbers of differentially expressed genes (<b>a</b>,<b>b</b>) and differentially expressed proteins (<b>c</b>,<b>d</b>) in leaves of <span class="html-italic">S. nigrum</span> by transcriptome and proteome. Plants were exposed to a complete Hoagland solution (CK) or with 100 μmol·L<sup>−1</sup> Zn (Zn), 100 μmol·L<sup>−1</sup> Cd (Cd), and 100 μmol·L<sup>−1</sup> Zn +100 μmol·L<sup>−1</sup> Cd (ZnCd) for 10 days. Rising green arrow shows increase, and falling red arrow shows decrease in significant differential expression between sample set (Zn vs. CK, Cd vs. CK, and ZnCd vs. CK).</p>
Full article ">Figure 4
<p>Identification and gene expression levels of significantly differentially expressed transporters in leaves of <span class="html-italic">S. nigrum</span> by transcriptome. Proportions of the identified transporters (<b>a</b>). Gene expression level of metal transporters (<b>b</b>); ABC transporters (<b>c</b>); peptide transporters (<b>d</b>); nitrate, phosphate, and boron transporters (<b>e</b>); and sulfate and amino acid transporters (<b>f</b>). The boxed transporter genes were then verified by qRT-PCR. Plant was exposed to a complete Hoagland solution (CK) or with 100 μmol·L<sup>−1</sup> Zn (Zn), 100 μmol·L<sup>−1</sup> Cd (Cd), and 100 μmol·L<sup>−1</sup> Zn+100 μmol·L<sup>−1</sup> Cd (ZnCd) for 10 days. Expression levels of transporters shown use Log<sub>2</sub> (fold change) between sample sets (Zn vs. CK, Cd vs. CK, and ZnCd vs. CK). <span class="html-italic">ABC</span> (<span class="html-italic">A</span>, <span class="html-italic">B</span>, <span class="html-italic">C</span>, <span class="html-italic">F</span>, <span class="html-italic">G</span>, <span class="html-italic">I</span>): ATP-biding cassette transporter six subfamilies; <span class="html-italic">Sultr</span>, sulfate transporter; <span class="html-italic">AAT</span>, amino acid transporter; <span class="html-italic">ZIP</span>, zinc transporter; <span class="html-italic">COP</span>, copper transporter; <span class="html-italic">Nramp</span>, natural resistance associated macrophage protein; <span class="html-italic">YSL</span>, metal–nicotianamine transporter; <span class="html-italic">VIT</span>, vacuolar iron transporter; <span class="html-italic">MGT</span>, magnesium transporter; <span class="html-italic">PTR</span>, peptide transporter; <span class="html-italic">OPT</span>, oligopeptide transporter; <span class="html-italic">NRT</span>, nitrate transporter; <span class="html-italic">PNT</span>, peptide/nitrate transporter; <span class="html-italic">BOR</span>, boron transporter; <span class="html-italic">KT</span>, potassium transporter; <span class="html-italic">PHT</span>, phosphate transporter; <span class="html-italic">SWEET</span>, bidirectional sugar transporter.</p>
Full article ">Figure 5
<p>Relative gene expression level of transporters in leaves of <span class="html-italic">S. nigrum</span> by qRT-PCR. Plant was exposed to a complete Hoagland solution (CK) or with 100 μmol·L<sup>−1</sup> Zn (Zn100), 100 μmol·L<sup>−1</sup> Cd (Cd100), and 100 μmol·L<sup>−1</sup> Zn+100 μmol·L<sup>−1</sup> Cd (Zn100+Cd100) for 10 days. Relative expression level of genes denoted by different letters refer to the significant differences (<span class="html-italic">p</span> &lt; 0.05, Duncan’s test).</p>
Full article ">Figure 6
<p>Expression levels of DEGs and DEPs involved in glutathione (<b>a</b>,<b>b</b>) and malate (<b>c</b>,<b>d</b>) metabolism in leaves of <span class="html-italic">S. nigrum</span> by transcriptome and proteome. The boxes with the same color are the same genes. Plant was exposed to a complete Hoagland solution (CK) or with 100 μmol·L<sup>−1</sup> Zn (Zn), 100 μmol·L<sup>−1</sup> Cd (Cd), and 100 μmol·L<sup>−1</sup> Zn+100 μmol·L<sup>−1</sup> Cd (ZnCd) for 10 days. Expression level of gene by transcriptome was shown using Log<sub>2</sub> (fold change) between sample sets (Zn vs. CK, Cd vs. CK, and ZnCd vs. CK). Expression level of protein by proteome was shown using a fold change (<span class="html-italic">p</span> &lt; 0.05, Student’s <span class="html-italic">t</span>-test) between sample sets (Zn/CK, Cd/CK, and ZnCd/CK). GRK, cysteine-rich receptor-like protein kinase; CysS, cysteine synthase; GPX, glutathione peroxidase; GR, glutathione reductase; GST, glutathione S-transferase; CysP, cysteine proteinase precursor; Lgl, lactoylglutathione lyase; MDH, malate dehydrogenase; DTC, dicarboxylate/tricarboxylate transporter; CS, ATP-citrate synthase.</p>
Full article ">Figure 7
<p>Expression levels of DEPs involved in chlorophyll (<b>a</b>) and ATP metabolism (<b>b</b>), chlorophyll content (<b>c</b>), and cytochemical characteristics (<b>d</b>) in leaves of <span class="html-italic">S. nigrum</span>. Plant was exposed to a complete Hoagland solution (CK) or with 100 μmol·L<sup>−1</sup> Zn (Zn), 100 μmol·L<sup>−1</sup> Cd (Cd), and 100 μmol·L<sup>−1</sup> Zn+100 μmol·L<sup>−1</sup> Cd (ZnCd) for 10 days. Expression level of protein was shown using a fold change (<span class="html-italic">p</span> &lt; 0.05, Student’s <span class="html-italic">t</span>-test) between sample sets (Zn/CK, Cd/CK, and ZnCd/CK). Chlorophyll (Chl) contents denoted by different letters refer to the significant differences (<span class="html-italic">p</span> &lt; 0.05, Duncan’s test). Paraffin-section experiments were repeated at least three times with similar results; bar, 20 μm. psbA, photosystem I P700 chlorophyll apoprotein; psbC, photosystem II CP43 chlorophyll apoprotein; RCCR, red chlorophyll catabolite reductase; POR, protochlorophyllide reductase; CAB, chlorophyll <span class="html-italic">a</span>/<span class="html-italic">b</span> binding protein; H<sup>+</sup>-ATPase, plasma membrane H<sup>+</sup>-ATPase; Zmp, ATP-dependent zinc metalloprotease; PFK, ATP-dependent 6-phosphofructokinase; ANT, ADP/ATP translocator; V-ATPase, vacuolar-type ATPase; ClpP, ATP-dependent Clp protease; ASP, ATP sulfurylase; ADRH, ATP-dependent RNA helicase.</p>
Full article ">Figure 8
<p>Molecular mechanism involved in transport and accumulation of Cd in leaves of <span class="html-italic">S. nigrum</span> exposed to Zn and Cd. Magenta and green pellets indicate Cd and Zn, respectively; and the genes or proteins in red font represent those upregulated by Cd or Zn in leaves of <span class="html-italic">S. nigrum</span>. Cd or Zn enters into leaf cells by plasma membrane transporters of <span class="html-italic">Nramp1</span>, <span class="html-italic">YSLs</span>, <span class="html-italic">ZIPs</span>, etc.; <span class="html-italic">MTs</span> and <span class="html-italic">GSTs</span> in cells are induced for antioxidant protection or chelation with excess metal ions; and then Cd-GSH complexes are transported to vacuoles for sequestration, or to cell walls for xylem transport by ABCs and OPTs. In addition, Zn promoted electron transport chain (ETC) activities and ATP biosynthesis via increased expression levels of MDHs and ATPases.</p>
Full article ">
15 pages, 10476 KiB  
Article
Effects of Cold Acclimation on Morpho-Anatomical Traits of Heteroblastic Foliage in Pinus massoniana (Lamb.) Seedlings
by Yingying Xu, Haoyun Wang, Hongyang He and Feng Wu
Forests 2024, 15(9), 1560; https://doi.org/10.3390/f15091560 - 5 Sep 2024
Viewed by 344
Abstract
Cold acclimation before winter has been shown to enhance the cold tolerance of evergreen conifers, including Pinus massoniana Lamb., a characteristic heteroblastic foliage tree in the conifer. In the initial growing season of P. massoniana, both primary needle seedlings (PNSs) and secondary [...] Read more.
Cold acclimation before winter has been shown to enhance the cold tolerance of evergreen conifers, including Pinus massoniana Lamb., a characteristic heteroblastic foliage tree in the conifer. In the initial growing season of P. massoniana, both primary needle seedlings (PNSs) and secondary needle seedlings (SNSs) are generated. While previous research has highlighted differences in the morphological structure and photosynthetic physiological functions of primary and secondary needles, their response to cold acclimation remains poorly understood. This study aimed to investigate the changes in morpho-anatomical structure, starch grain accumulation, and lignin deposition in the roots, stems, and leaves of PNSs and SNSs during cold acclimation using solid potassium iodide and hydrochloric acid phloroglucinol double-staining techniques. The results revealed that, during cold acclimation, the leaves and stems of PNSs exhibited sensitivity to low-temperature stress, resulting in noticeable shrinkage and fracture of mesophyll and cortical parenchyma cells. Furthermore, the early stages of cold acclimation promoted the accumulation of starch grains and lignin in the seedling tissues. In contrast to PNSs, the leaves and stems of SNSs exhibited a shorter cold acclimation period, attributed to the hydrolysis of starch grains in the epidermal cell walls and the transformation of xylem lignin, which supports cell structure stability and enhances cold resistance. In conclusion, these findings suggest that SNSs displayed a superior cold resistance potential compared to PNSs following cold acclimation, providing a significant theoretical basis for the further screening of cold-tolerant germplasm resources of P. massoniana and the analysis of cold resistance traits in heteroblastic foliage. Full article
(This article belongs to the Section Forest Ecophysiology and Biology)
Show Figures

Figure 1

Figure 1
<p>Morphological observation of the primary needle seedlings (PNSs) and secondary needle seedlings (SNSs). Primary needles are short and flat, singly attached to the stem; secondary needles are significantly longer and thicker, clustered in groups of two needles.</p>
Full article ">Figure 2
<p>Morphological structure and starch grain accumulation of leaf in <span class="html-italic">P. massoniana</span> during cold acclimation. Ep, epidermis; En, endodermis; Mc, mesophyll cell; Xy, xylem. The longitudinal (<b>a</b>–<b>e</b>) and cross-cutting (<b>a1</b>–<b>e1</b>) structure of PNSs at 0 h, 24 h, 72 h, 168 h, and 336 h, respectively. The longitudinal (<b>f</b>–<b>j</b>) and cross-cutting (<b>f1</b>–<b>j1</b>) structure of SNSs. The green represents starch grains.</p>
Full article ">Figure 3
<p>Morphological structure and starch grain accumulation of stem in <span class="html-italic">P. massoniana</span> during cold acclimation. Ep, epidermis; Cp, cortex; Xy, xylem. The longitudinal (<b>a</b>–<b>e</b>) and cross-cutting (<b>a1</b>–<b>e1</b>) structure of PNSs at 0 h, 24 h, 72 h, 168 h, and 336 h, respectively. The longitudinal (<b>f</b>–<b>j</b>) and cross-cutting (<b>f1</b>–<b>j1</b>) structure of SNSs. The green represents starch grains.</p>
Full article ">Figure 4
<p>Morphological structure and starch grain accumulation of root in <span class="html-italic">P. massoniana</span> during cold acclimation. Ep, epidermis; Cp, cortex; St, stele. The longitudinal (<b>a</b>–<b>e</b>) and cross-cutting (<b>a1</b>–<b>e1</b>) structure of PNSs at 0 h, 24 h, 72 h, 168 h, and 336 h respectively; the longitudinal (<b>f</b>–<b>j</b>) and cross-cutting (<b>f1</b>–<b>j1</b>) structure of SNSs. The green represents starch grains.</p>
Full article ">Figure 5
<p>Morphological structure and lignin content of leaf in <span class="html-italic">P. massoniana</span> during cold acclimation. Ep, epidermis; En, endodermis; Mc, mesophyll cell; Xy, xylem. The longitudinal (<b>a</b>–<b>e</b>) and cross-cutting (<b>a1</b>–<b>e1</b>) structure of PNSs at 0 h, 24 h, 72 h, 168 h, and 336 h, respectively. The longitudinal (<b>f</b>–<b>j</b>) and cross-cutting (<b>f1</b>–<b>j1</b>) structure of SNSs. The red represents lignin accumulation.</p>
Full article ">Figure 6
<p>Morphological structure and lignin content of stem in <span class="html-italic">P. massoniana</span> during cold acclimation. Ep, epidermis; Cp, cortex; Xy, xylem. The longitudinal (<b>a</b>–<b>e</b>) and cross-cutting (<b>a1</b>–<b>e1</b>) structure of PNSs at 0 h, 24 h, 72 h, 168 h, and 336 h, respectively. The longitudinal (<b>f</b>–<b>j</b>) and cross-cutting (<b>f1</b>–<b>j1</b>) structure of SNSs. The red represents lignin accumulation.</p>
Full article ">Figure 7
<p>Morphological structure and lignin content of root in <span class="html-italic">P. massoniana</span> during cold acclimation. Ep, epidermis; Cp, cortex; St, stele. The longitudinal (<b>a</b>–<b>e</b>) and cross-cutting (<b>a1</b>–<b>e1</b>) structure of PNSs at 0 h, 24 h, 72 h, 168 h, and 336 h, respectively. The longitudinal (<b>f</b>–<b>j</b>) and cross-cutting (<b>f1</b>–<b>j1</b>) structure of SNSs. The red represents lignin accumulation.</p>
Full article ">
15 pages, 1881 KiB  
Article
Variations in Root Characteristics and Cadmium Accumulation of Different Rice Varieties under Dry Cultivation Conditions
by Chaoping Shan, Can Shi, Xinran Liang, Yanqun Zu, Jixiu Wang, Bo Li and Jianjun Chen
Plants 2024, 13(17), 2457; https://doi.org/10.3390/plants13172457 - 2 Sep 2024
Viewed by 446
Abstract
Variations in the cadmium (Cd) accumulation and root characteristics of different genotypes of rice during three developmental periods of dry cultivation were investigated in pot experiments in which two levels of Cd were added to the soil (0 and 10 mg kg−1 [...] Read more.
Variations in the cadmium (Cd) accumulation and root characteristics of different genotypes of rice during three developmental periods of dry cultivation were investigated in pot experiments in which two levels of Cd were added to the soil (0 and 10 mg kg−1). The results show that the Cd concentration in each organ of the different rice genotypes decreased in both the order of roots > shoots > grains and during the three developmental periods in the order of the maturity stage > booting stage > tillering stage. The lowest bioaccumulation factor (BCF) and translocation factor (TF) were found in Yunjing37 (YJ37) under Cd stress. At maturity, Cd stress inhibited the root length of Dianheyou34 (DHY34) the most and that of Dianheyou 918 (DHY918) the least, also affecting the root volume of DHY34 and Dianheyou615 (DHY615) the most and that of YJ37 and Yiyou 673 (YY673) the least; the inhibition rates were 41.80, 5.09, 40.95, and 10.51%, respectively. The exodermis showed the greatest thickening in YY673 and the lowest thickening in DHY615, while the endodermis showed the opposite result. The rates of change were 16.48, 2.45, 5.10, and 8.49%, respectively. The stele diameter of DHY615 decreased the most, and that of YY673 decreased the least, while the secondary xylem area showed the opposite result; the rates of change were −21.50, −14.29, −5.86, and −26.35%, respectively. Under Cd stress treatment at maturity, iron plaque was extracted using the dithionite–citrate–bicarbonate (DCB) method. The concentration of iron (DCB-Fe) was highest in YJ37, and the concentration of cadmium (DCB-Cd) was lowest in DHY34. YJ37 was screened as a low Cd-accumulating variety. The concentration of available Cd in the rhizosphere soil, iron plaque, root morphology, and anatomy affect Cd accumulation in rice with genotypic differences. Our screening of Cd-accumulating rice varieties provides a basis for the dry cultivation of rice in areas with high background values of Cd in order to avoid the health risks of Cd intake. Full article
(This article belongs to the Special Issue Crop Plants and Heavy Metals)
Show Figures

Figure 1

Figure 1
<p>The Cd concentrations in the various organs and rhizosphere soil of different genotypes of rice under dry cultivation conditions. (<b>a</b>) Cd concentrations in roots of different genotypes of rice in three developmental periods (<span class="html-italic">n</span> = 3); (<b>b</b>) Cd concentrations in shoots of different genotypes of rice in three developmental periods (<span class="html-italic">n</span> = 3); (<b>c</b>) Cd concentrations in grains of different genotypes of rice at maturity stage (<span class="html-italic">n</span> = 3); (<b>d</b>) available Cd concentrations in the rhizosphere soil of different genotypes of rice in three developmental periods. The values are the mean ± standard deviation (<span class="html-italic">n</span> = 3). Different letters represent significance at <span class="html-italic">p</span> &lt; 0.05 (Duncan) among the different genotypes of rice under the same treatment during the same stage of dry cultivation. CK represents a concentration of 0 mg kg<sup>−1</sup> Cd added externally to the soil. Cd treatment represents a concentration of 10 mg kg<sup>−1</sup> Cd added externally to the soil. Abbreviations: DHY34, Dianheyou 34; DHY615, Dianheyou 615; DHY918, Dianheyou 918; YY673, Yiyou 673; YJ37, Yunjing 37.</p>
Full article ">Figure 2
<p>The effect of Cd stress treatments on the BCF (<b>a</b>) and TF (<b>b</b>) values of different genotypes of rice under dry cultivation conditions. The BCF values use the date at maturity, and the TF values use the date in three developmental periods. The BCF was calculated using Equation (1) to evaluate the accumulation of Cd in soils and grains. The TF was calculated using Equation (2) to evaluate the upward conduction ability of Cd in roots and shoots. The values are the mean ± standard deviation (<span class="html-italic">n</span> = 3). Different letters represent significance at <span class="html-italic">p</span> &lt; 0.05 (Duncan) among the different genotypes of rice under the same treatment during the same stage of dry cultivation. CK represents a concentration of 0 mg kg<sup>−1</sup> Cd added externally to the soil. Cd treatment represents a concentration of 10 mg kg<sup>−1</sup> Cd added externally to the soil. Abbreviations: DHY34, Dianheyou 34; DHY615, Dianheyou 615; DHY918, Dianheyou 918; YY673, Yiyou 673; YJ37, Yunjing 37.</p>
Full article ">Figure 3
<p>Cluster analyses of Cd accumulation characteristic phenotypes. The grain Cd concentration, Cd accumulation, BCFs, and TFs of five different genotypes with Cd stress under dry cultivation conditions (<span class="html-italic">n</span> = 3). The Cd concentrations in grains and BCFs use the date at the maturity stage, while the Cd accumulation and TFs use the dates for the three developmental periods. Abbreviations: DHY34, Dianheyou 34; DHY615, Dianheyou 615; DHY918, Dianheyou 918; YY673, Yiyou 673; YJ37, Yunjing 37.</p>
Full article ">Figure 4
<p>The effect of the Cd stress treatment on the root morphology of different genotypes of rice under dry cultivation. (<b>a</b>) Inhibition rate of Cd stress treatment on the root lengths of 5 genotypes of rice under dry cultivation conditions in three developmental periods. (<b>b</b>) Inhibition rate of Cd stress treatment on the root volume of 5 genotypes of rice in three developmental periods. Fresh and clean whole roots were used to analyze root morphology. The inhibition rate was obtained using Equation (3) and represents the degree of inhibition of rice root morphology under the Cd stress treatments. The values are the mean ± standard deviation (<span class="html-italic">n</span> = 3). Different letters indicate significance at <span class="html-italic">p</span> &lt; 0.05 (Duncan) between the inhibition rates of the Cd stress treatments on the root length and root volume of different genotypes of rice at the same stage under dry cultivation conditions. Abbreviations: DHY34, Dianheyou 34; DHY615, Dianheyou 615; DHY918, Dianheyou 918; YY673, Yiyou 673; YJ37, Yunjing 37.</p>
Full article ">Figure 5
<p>The anatomy of the root systems of different rice genotypes under dry cultivation conditions. The scale bar is 250 μm. The number of rice root sections used for anatomical observation was 30 (<span class="html-italic">n</span> = 3). Abbreviations: DHY34, Dianheyou 34; DHY615, Dianheyou 615; DHY918, Dianheyou 918; YY673, Yiyou 673; YJ37, Yunjing 37.</p>
Full article ">Figure 6
<p>Rates of change in the anatomical parameters of the root systems of different rice genotypes under dry cultivation conditions. The rates of change were calculated using Equation (3) to represent the extent of the effect of the Cd stress treatment on rice. The values are the mean ± standard deviation (n = 3). Different letters indicate significance at p &lt; 0.05 (Duncan) among the change rates. Abbreviations: DHY34, Dianheyou 34; DHY615, Dianheyou 615; DHY918, Dianheyou 918; YY673, Yiyou 673; YJ37, Yunjing 37.</p>
Full article ">Figure 7
<p>Effects of Cd stress treatments on DCB-Fe and DCB-Cd concentrations in different rice genotypes at maturity under dry cultivation. DCB-Fe indicates the Fe concentration in the iron plaque on the root surface, and DCB-Cd indicates the Cd concentration in the iron plaque on the root surface. The values are the mean ± standard deviation (<span class="html-italic">n</span> = 3). Different letters represent the variability of DCB-Fe and DCB-Cd concentrations among different rice genotypes in the same treatment at maturity. Abbreviations: DHY34, Dianheyou 34; DHY615, Dianheyou 615; DHY918, Dianheyou 918; YY673, Yiyou 673; YJ37, Yunjing 37.</p>
Full article ">Figure 8
<p>Correlation analysis of Cd accumulation in different genotypes of rice at maturity under dry cultivation. Correlation analyses (Pearson, <span class="html-italic">p</span> ≤ 0.05) between Cd concentrations in various rice organs and available Cd concentrations in the rhizosphere soil, root morphology, root anatomy, DCB-Fe, and DCB-Cd at maturity under Cd stress treatments. Asterisks indicate significance at * <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 9
<p>An SEM showing the direct and indirect effects of Cd treatment, the available Cd concentrations in the rhizosphere soil, root morphology, root anatomy, root surface iron plaques, root Cd concentration, and shoot Cd concentration on grain Cd concentration (<b>a</b>). Standardized direct, indirect, and total effects of Cd treatment, available Cd concentrations in the rhizosphere soil, root morphology, root anatomy, root surface iron plaques, root Cd concentration, and shoot Cd (<b>b</b>). An SEM plotted using maturity data. Root morphology was analyzed using principal components for root length and root volume, and the first principal component was selected with a contribution of 78.60%. Root anatomy was analyzed using principal components for exodermis and endodermis thickness, and the first principal component was selected with a contribution of 71.13%. Iron plaque was analyzed using principal components for DCB-Fe and DCB-Cd concentration, and the first principal component was selected with a contribution of 81.73%. Numbers adjacent to arrows are standardized path coefficients, analogous to relative regression weights, and are indicative of the magnitude of the correlation. Continuous and dashed arrows indicate positive and negative relationships, respectively. The arrow width is proportional to the strength of the relationship. The proportion of variance explained (<span class="html-italic">R</span><sup>2</sup>) appears alongside every response variable in the model. Goodness-of-fit statistics for the model are shown at the bottom of the image. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
62 pages, 2191 KiB  
Review
The Role of Low-Molecular-Weight Organic Acids in Metal Homeostasis in Plants
by Ilya V. Seregin and Anna D. Kozhevnikova
Int. J. Mol. Sci. 2024, 25(17), 9542; https://doi.org/10.3390/ijms25179542 - 2 Sep 2024
Viewed by 417
Abstract
Low-molecular-weight organic acids (LMWOAs) are essential O-containing metal-binding ligands involved in maintaining metal homeostasis, various metabolic processes, and plant responses to biotic and abiotic stress. Malate, citrate, and oxalate play a crucial role in metal detoxification and transport throughout the plant. This review [...] Read more.
Low-molecular-weight organic acids (LMWOAs) are essential O-containing metal-binding ligands involved in maintaining metal homeostasis, various metabolic processes, and plant responses to biotic and abiotic stress. Malate, citrate, and oxalate play a crucial role in metal detoxification and transport throughout the plant. This review provides a comparative analysis of the accumulation of LMWOAs in excluders, which store metals mainly in roots, and hyperaccumulators, which accumulate metals mainly in shoots. Modern concepts of the mechanisms of LMWOA secretion by the roots of excluders and hyperaccumulators are summarized, and the formation of various metal complexes with LMWOAs in the vacuole and conducting tissues, playing an important role in the mechanisms of metal detoxification and transport, is discussed. Molecular mechanisms of transport of LMWOAs and their complexes with metals across cell membranes are reviewed. It is discussed whether different endogenous levels of LMWOAs in plants determine their metal tolerance. While playing an important role in maintaining metal homeostasis, LMWOAs apparently make a minor contribution to the mechanisms of metal hyperaccumulation, which is associated mainly with root exudates increasing metal bioavailability and enhanced xylem loading of LMWOAs. The studies of metal-binding compounds may also contribute to the development of approaches used in biofortification, phytoremediation, and phytomining. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Participation of organic acids in metal transport and detoxification in plants. A generalized scheme is presented, without taking tissue specificity into account. The release of root exudates containing various organic acids (OAs) is the basis of the exclusion tolerance mechanism. Malate is secreted by aluminum-activated malate transporter 1 (ALMT1), whereas citrate is secreted by some members of the multidrug and toxic compound extrusion transporter (MATE) family (e.g., AtMATE, BoMATE, EcMATE1, FeMATE1, GmMATE13/47, GsMATE, HvAACT1, MtMATE66, OsFRDL2/4, PtrMATE1, SbMATE, ScFRDL2, TaMATE1B, VuMATE1/2, and ZmMATE1), which are located at the plasma membrane of the root rhizodermal cells. In the rhizosphere, OAs bind metal ions (Me<sup>n+</sup>) with the formation of complexes of various structures (Me-OA complexes), which affects metal entry into the plant. Aluminum ions (Al<sup>3+</sup>) in the cell walls of root rhizodermal cells may bind to malate, which is transported there by ALMT1, and the resulting complexes (Al-malate) are transported across the plasma membrane into the cytosol with the involvement of nodulin 26-like intrinsic protein (NIP1;2). In the cell, the major site of OA biosynthesis is the mitochondria and the glyoxisomes, where the reactions of the Krebs cycle and the glyoxylate cycle, respectively, take place. Pyruvate transport into the mitochondria is mediated by the mitochondrial pyruvate carrier (MPC). The transport of other OAs across the inner membrane of the mitochondria is carried out by the exchange mechanism with the involvement of dicarboxylate carriers (DICs), dicarboxylate/tricarboxylate carrier (DTC), and succinate/fumarate carrier (SFC). Having entered the cytosol via the plasma membrane transporters, metal ions bind to different low-molecular-weight ligands, including OAs, though the stability of Me-OA complexes at neutral pH values is lower than that of metal complexes with nicotianamine and histidine. The possibility of translocation of Me-OA complexes across the tonoplast cannot be excluded, though the mechanism of such transport is unknown yet (this pathway is designated by a dotted line and the unknown transporter by a question mark). The entry of metal ions into the vacuole is carried out by different vacuolar transporters, whereas the transport of OAs across the tonoplast is mediated by the tonoplast dicarboxylate transporter (tDT) and the vacuolar citrate/H+ symporter Cit1 (the latter is not shown in the figure). In <span class="html-italic">A. thaliana</span>, OA transport across the tonoplast is also carried out by ALMT4 and ALMT6, whose gene expression was shown in stomatal guard cells, and for ALMT4—also in leaf mesophyll. Potential substrates for the mitochondrial carriers (DICs, DTC, and SFC) and the vacuolar OA-transporting proteins (tDT and ALMT4) are shown next to the corresponding transporters/carriers/channels. Metal binding to OAs in the vacuole is an important internal metal detoxification mechanism. The transport of malate across the plasma membrane in stomatal guard cells in <span class="html-italic">A. thaliana</span> may be carried out by ALMT12 and ABCB14 (ATP-binding cassette subfamily B protein), being directly involved in the stomatal movement and indirectly involved in maintaining metal homeostasis. Metals are translocated from roots to shoots via the xylem mainly as complexes with OAs. Citrate is transported into the xylem vessels by the FRD3 transporter (ferric chelate reductase defective 3) located at the plasma membrane of root central cylinder cells, as well as other proteins of the MATE family (e.g., OsFRDL1, PtrMATE1, MtMATE66, and ScFRDL1). Plastidic transport of OAs in the cells of photosynthetic tissues is carried out via a double-transporter system at the inner chloroplast membrane involving the plastidic 2-oxoglutarate/malate transporter (OMT1) and the general dicarboxylate transporter (DCT1). The arrows show the direction of transport. Designations: Acetyl-CoA, acetyl coenzyme A; ACO, aconitase; CoA-SH, coenzyme A; CS, citrate synthase; FAD-SDH, FAD-succinic dehydrogenase; FUM, fumarase; GABA, gamma-aminobutyric acid; Glu, glutamic acid; ICL, isocitrate lyase; MS, malate synthase; NAD-IDH, NAD-isocitrate dehydrogenase; NAD-MDH, NAD-malate dehydrogenase; NAD-ME, NAD-malic enzyme; NAD-OGDH, NAD-2-oxoglutarate dehydrogenase; NADP-ME, NADP-malic enzyme; NADP-IDH, NADP-isocitrate dehydrogenase; PDC, pyruvate dehydrogenase complex; PEP, phosphoenolpyruvate; PEPC, phosphoenolpyruvate carboxylase; PEPCK, phosphoenolpyruvate carboxykinase; PPDK, pyruvate orthophosphate dikinase; STK, succinate thiokinase (succinyl-CoA synthetase).</p>
Full article ">Figure 2
<p>Protonation/deprotonation equilibria of citric (<b>A</b>) and malic (<b>B</b>) acids. p<span class="html-italic">K</span><sub>α</sub> values for citric and malic acids are presented according to [<a href="#B116-ijms-25-09542" class="html-bibr">116</a>,<a href="#B117-ijms-25-09542" class="html-bibr">117</a>], respectively.</p>
Full article ">
13 pages, 4556 KiB  
Article
Phosphate Turnover in Various Parts of Nodulated Soybean (Glycine max (L.) Merr.) Plants and the Relation to the Xylem Transport
by Yoshiaki Yamamura, Kyoko Higuchi, Akihiro Saito and Takuji Ohyama
Crops 2024, 4(3), 413-425; https://doi.org/10.3390/crops4030029 - 2 Sep 2024
Viewed by 235
Abstract
Phosphorus is a major essential element in plants, and the absorption and transport of P are related to crop growth and productivity. Phosphate (Pi) is absorbed in the roots and transported to the shoot. Plants store surplus Pi in the vacuoles. The characteristics [...] Read more.
Phosphorus is a major essential element in plants, and the absorption and transport of P are related to crop growth and productivity. Phosphate (Pi) is absorbed in the roots and transported to the shoot. Plants store surplus Pi in the vacuoles. The characteristics of Pi storage and turnover in various parts of the nodulated soybeans might be related to plant growth and P-use efficiency. This research focused on the changes in the Pi concentrations and Pi contents in each part of young soybean plants grown in Pi-sufficient (50 μM Pi) or Pi-deficient (0 μM Pi) conditions. Also, the Pi flux rate in xylem sap from roots to shoot was determined. The growth of the plants was the same after 7 days of Pi-sufficient and Pi-deficient treatments. During the Pi-deficient period, the Pi concentrations in the roots, leaves, and stems decreased significantly but did not deplete. The decrease in Pi concentration in nodules was much slower than the other parts. After the re-supply of 50 μM Pi in the solution, the Pi concentration increased only a little in each part of the Pi-deficient plants. The Pi concentration and Pi flux in the xylem sap quickly responded to the changes in the Pi concentration in the culture solution. Full article
Show Figures

Figure 1

Figure 1
<p>The effect of Pi-sufficient (<b>A</b>) and Pi-deficient (<b>B</b>) treatments on the dry weight of each part. Pi-deficient treatment was carried out from day 0 to day 7. The blue background in (<b>B</b>) indicates the period of Pi-deficient conditions. Average ± standard error. There was no significant difference in the dry weight of each part between Pi-sufficient and Pi-deficient treatments using Student’s <span class="html-italic">T</span>-test (<span class="html-italic">p</span> &lt; 0.05). <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 2
<p>The effect of Pi-sufficient (<b>A</b>) and Pi-deficient (<b>B</b>) treatments on the Pi content in each part. Average ± standard error. * and ** indicate the significant difference in the total values at 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background in (<b>B</b>) indicates the period of Pi-deficient conditions. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 3
<p>Effect of Pi-sufficient and Pi-deficient treatments on the xylem sap exudation rate (<b>A</b>) and the transpiration rate (<b>B</b>). The xylem sap was collected for 1 h after the decapitation of the shoot. The transpiration rate was measured by the decrease in the weight of the culture solution for one day just before sampling and divided by a 16 h light period per day. Average ± standard error. Different alphabet indicates the significant difference in the values among treatments times in Pi-sufficient (blue) and Pi-deficient (red) by Tukey’s Test. * indicates the significant difference at <span class="html-italic">p</span> &lt; 0.05 in the values between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background in (<b>B</b>) indicates the period of Pi-deficient conditions. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 4
<p>Comparison of Pi-sufficient and Pi-deficient treatments on Pi concentration (<b>A</b>) and Pi content (<b>B</b>) in the roots. Average ± standard error. Different letters indicate the significant difference in the values among treatments times in Pi-sufficient (blue) and Pi-deficient (red) by Tukey’s Test. * and ** indicate the significant difference in the values at 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background in (<b>B</b>) indicates the period of Pi-deficient conditions. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 5
<p>Comparison of Pi-sufficient and Pi-deficient treatments on Pi concentration (<b>A</b>) and Pi content (<b>B</b>) in the leaves. Average ± standard error. Different letters indicate the significant difference in the values among treatments times in Pi-sufficient (blue) and Pi-deficient (red) by Tukey’s Test. * and ** indicate the significant difference in the values at 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background indicates the period of P-deficient treatment. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 6
<p>Comparison of Pi-sufficient and Pi-deficient treatments on Pi concentration (<b>A</b>) and Pi content (<b>B</b>) in the stems and petioles. Average ± standard error. Different letters indicate the significant difference in the values among treatment times in Pi-sufficient (blue) and Pi-deficient (red) by Tukey’s Test. * and ** indicate the significant difference in the values at 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background indicates the period of Pi-deficient treatment. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 7
<p>Comparison of Pi-sufficient and Pi-deficient treatments on Pi concentration (<b>A</b>) and Pi content (<b>B</b>) in the nodules. Average ± standard error. Different letters indicate the significant difference in the values among treatment times in Pi-sufficient (blue) and Pi-deficient (red) by Tukey’s Test. * and ** indicate the significant difference in the values at 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background indicates the period of Pi-deficient treatment. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 8
<p>Comparison of Pi-sufficient and Pi-deficient treatments on Pi concentration (<b>A</b>) and Pi content (<b>B</b>) in the buds and young developing leaves. Average ± standard error. Different letters indicate the significant difference in the values among treatment times in Pi-sufficient (blue) and Pi-deficient (red) by Tukey’s Test. * and ** indicate the significant difference in the values at 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background indicates the period of P-deficient treatment. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 9
<p>Comparison of Pi-sufficient and Pi-deficient treatments on Pi concentration (<b>A</b>) in the xylem sap and Pi flux (<b>B</b>). Average ± standard error. Different letters indicate the significant difference in the values among treatment times in Pi-sufficient (blue) and Pi-deficient (red) by Tukey’s Test. * and ** indicate the significant difference in the values at 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 between Pi-sufficient and Pi-deficient treatments by Student’s <span class="html-italic">T</span>-test. The blue background indicates the period of Pi-deficient treatment. <span class="html-italic">n</span> = 3.</p>
Full article ">
24 pages, 7903 KiB  
Article
Populus trichocarpa EXPA6 Facilitates Radial and Longitudinal Transport of Na+ under Salt Stress
by Zhe Liu, Kexin Yin, Ying Zhang, Caixia Yan, Ziyan Zhao, Jing Li, Yi Liu, Bing Feng, Rui Zhao, Jian Liu, Kaiyue Dong, Jun Yao, Nan Zhao, Xiaoyang Zhou and Shaoliang Chen
Int. J. Mol. Sci. 2024, 25(17), 9354; https://doi.org/10.3390/ijms25179354 - 29 Aug 2024
Viewed by 299
Abstract
Expansins are cell wall (CW) proteins that mediate the CW loosening and regulate salt tolerance in a positive or negative way. However, the role of Populus trichocarpa expansin A6 (PtEXPA6) in salt tolerance and the relevance to cell wall loosening is still unclear [...] Read more.
Expansins are cell wall (CW) proteins that mediate the CW loosening and regulate salt tolerance in a positive or negative way. However, the role of Populus trichocarpa expansin A6 (PtEXPA6) in salt tolerance and the relevance to cell wall loosening is still unclear in poplars. PtEXPA6 gene was transferred into the hybrid species, Populus alba × P. tremula var. glandulosa (84K) and Populus tremula × P. alba INRA ‘717-1B4’ (717-1B4). Under salt stress, the stem growth, gas exchange, chlorophyll fluorescence, activity and transcription of antioxidant enzymes, Na+ content, and Na+ flux of root xylem and petiole vascular bundle were investigated in wild-type and transgenic poplars. The correlation analysis and principal component analysis (PCA) were used to analyze the correlations among the characteristics and principal components. Our results show that the transcription of PtEXPA6 was downregulated upon a prolonged duration of salt stress (48 h) after a transient increase induced by NaCl (100 mM). The PtEXPA6-transgenic poplars of 84K and 717-1B4 showed a greater reduction (42–65%) in stem height and diameter growth after 15 days of NaCl treatment compared with wild-type (WT) poplars (11–41%). The Na+ accumulation in roots, stems, and leaves was 14–83% higher in the transgenic lines than in the WT. The Na+ buildup in the transgenic poplars affects photosynthesis; the activity of superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT); and the transcription of PODa2, SOD [Cu-Zn], and CAT1. Transient flux kinetics showed that the Na+ efflux of root xylem and leaf petiole vascular bundle were 1.9–3.5-fold greater in the PtEXPA6-transgenic poplars than in the WT poplars. PtEXPA6 overexpression increased root contractility and extensibility by 33% and 32%, indicating that PtEXPA6 increased the CW loosening in the transgenic poplars of 84K and 717-1B4. Noteworthily, the PtEXPA6-promoted CW loosening was shown to facilitate Na+ efflux of root xylem and petiole vascular bundle in the transgenic poplars. We conclude that the overexpression of PtEXPA6 leads to CW loosening that facilitates the radial translocation of Na+ into the root xylem and the subsequent Na+ translocation from roots to leaves, resulting in an excessive Na+ accumulation and consequently, reducing salt tolerance in transgenic poplars. Therefore, the downregulation of PtEXPA6 in NaCl-treated Populus trichocarpa favors the maintenance of ionic and reactive oxygen species (ROS) homeostasis under long-term salt stress. Full article
(This article belongs to the Special Issue Plant Response to Abiotic Stress—3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Transcription profile of <span class="html-italic">PtEXPA6</span> in the leaves, stems, and roots of <span class="html-italic">Populus trichocarpa</span> during the period of salt stress. Uniform plants of <span class="html-italic">P. trichocarpa</span> were treated with NaCl saline (0 or 100 mM) for 48 h. The fine roots, stems, and upper leaves (3rd to 8th from shoot tip) were sampled at 0, 3, 6, 12, 24, and 48 h, respectively. For the RT-qPCR analysis, the primer sequences for <span class="html-italic">PtEXPA6</span> and the reference gene, <span class="html-italic">PtUBQ</span>, are shown in <a href="#app1-ijms-25-09354" class="html-app">Supplementary Table S1</a>. The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with asterisks indicate significant differences, **: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Sequence and phylogenetic analysis of <span class="html-italic">Populus trichocarpa PtEXPA6</span>. (<b>A</b>) The multiple sequence alignment of EXPA and expansin family from <span class="html-italic">Populus</span> and other species. The black shading indicates identical amino acid residues, and the blue and pink shadings indicate conserved amino acids, respectively. The lower-case letters represent the same amino acids in different species. (<b>B</b>) The phylogenetic analysis of expansin from various species. <span class="html-italic">Populus euphratica</span> (Pe), <span class="html-italic">Populus trichocarpa</span> (Pt), <span class="html-italic">Populus tremula × Populus tremuloides</span> (Ptt), <span class="html-italic">Populus alba</span> (Pa), <span class="html-italic">Arabidopsis thaliana</span> (At), <span class="html-italic">Zea mays</span> (Zm), <span class="html-italic">Nicotiana tabacum</span> (Nt), and <span class="html-italic">Oryza sativa</span> (Os), <span class="html-italic">Glycine max</span> (Gm), <span class="html-italic">Salix viminalis</span> (Sv), <span class="html-italic">Morus notabilis</span> (Mn), <span class="html-italic">Rosa rugosa</span> (Rr), <span class="html-italic">Prunus persica</span> (Pp), <span class="html-italic">Pistacia vera</span> (Pv), <span class="html-italic">Ziziphus jujuba</span> (Zj), <span class="html-italic">Cucumis melo</span> (Cm), <span class="html-italic">Gossypium arboreum</span> (Ga), <span class="html-italic">Nicotiana sylvestris</span> (Ns), <span class="html-italic">Syzygium oleosum</span> (So), <span class="html-italic">Hibiscus syriacus</span> (Hs), <span class="html-italic">Pistacia vera</span> (Pv), and <span class="html-italic">Salix purpurea</span> (Sp). <a href="#app1-ijms-25-09354" class="html-app">Supplementary Table S2</a> lists the accession numbers of the EXPA orthologs. The blue, yellow, pink and green shadings indicate expansin orthologs of EXPA, EXLA, EXLB, and EXPB, respectively. The PtEXPA6 is labelled with star symbols in (<b>A</b>,<b>B</b>).</p>
Full article ">Figure 3
<p>Molecular verification of the transgenic lines overexpressing <span class="html-italic">P. trichocarpa PtEXPA6</span> in 84K and 717-1B4. (<b>A</b>) The PCR assay of the transgenic poplars. The primer sequences for <span class="html-italic">PtEXPA6</span> are shown in <a href="#app1-ijms-25-09354" class="html-app">Supplementary Table S1</a>. WT: negative control (wild type); CK: blank control; P: positive control. (<b>B</b>) The Western blot of the transgenic lines. The Western blot analysis performed with an anti-MYC-specific antibody for PtEXPA6.</p>
Full article ">Figure 4
<p>Phenotypic tests of the wild-type (WT) and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4 under long-term salt stress. The <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to NaCl with 0 or 100 mM for 15 days. The stem height and diameter of the no-salt control and salinized plants were measured after 15 days of the salt treatment. The relative growth of the stem height and diameter during the observation period are shown. (<b>A</b>) Representative images showing plant performance after the salt treatment. Scale bars = 5 cm. (<b>B</b>) Relative growth of the stem height. (<b>C</b>) Relative growth of the stem diameter. The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of NaCl on leaf gas exchange in the wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4. The <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to NaCl with 0 or 100 mM for 15 days. Leaf gas exchange, i.e., net photosynthetic rate, transpiration rate, and stomatal conductance were measured in the leaves of the no-salt control and salinized plants after 15 days of the salt treatment. (<b>A</b>) Net photosynthetic rate (Pn). (<b>B</b>) Transpiration rate (E). (<b>C</b>) Stomatal conductance (Cleaf). The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effect of NaCl on chlorophyll fluorescence in the wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4. The <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to NaCl with 0 or 100 mM for 15 days. Chlorophyll fluorescence, i.e., the relative electron transport rate, the actual photosynthetic quantum yield, and the maximum photochemical efficiency of PSII were measured in the leaves of the no-salt control and salinized plants after 15 days of the salt treatment. (<b>A</b>) The relative electron transport rate (ETR). (<b>B</b>) The actual photosynthetic quantum yield (YII). (<b>C</b>) The maximum photochemical efficiency of PSII (Fv/Fm). The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Effect of NaCl on antioxidant enzyme activity and relative electrolyte leakage in the wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4. The <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to NaCl with 0 or 100 mM for 15 days. The activity of superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT), and relative electrolyte leakage were measured in the leaves of the no-salt control and salinized plants after 15 days of the salt treatment. (<b>A</b>) POD activity. (<b>B</b>) SOD activity. (<b>C</b>) CAT activity. (<b>D</b>) Relative electrolyte leakage. The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Effect of NaCl on the transcription levels of antioxidant enzymes in the wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4. The <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K (L11, L12, and L13), 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to NaCl with 0 or 100 mM for 15 days. The relative expression of the antioxidant enzyme genes such as peroxidase a2 (<span class="html-italic">PODa2</span>), superoxide dismutase [Cu-Zn] (<span class="html-italic">SOD [Cu-Zn]</span>), and catalase 1 (<span class="html-italic">CAT1</span>) were examined in the WT and <span class="html-italic">PtEXPA6</span>-overexpressing poplars after 15 days of the salt treatment. (<b>A</b>) <span class="html-italic">PODa2</span>. (<b>B</b>) <span class="html-italic">SOD [Cu-Zn]</span>. (<b>C</b>) <span class="html-italic">CAT1</span>. The primer sequences of <span class="html-italic">PODa2</span>, <span class="html-italic">SOD [Cu-Zn]</span>, and <span class="html-italic">CAT1</span> and the reference actin gene, <span class="html-italic">PtUBQ</span>, are shown in <a href="#app1-ijms-25-09354" class="html-app">Supplementary Table S1</a>. The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Na<sup>+</sup> content in roots, stems, and leaves of wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4 under long-term salt stress. <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to NaCl with 0 or 100 mM for 15 days. Data are means ± SD (<span class="html-italic">n</span> = 3), and bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Comparative contractability and comparative extensibility of the intact root tip sites in the wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4. The comparative contractability of the <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) was measured after the intact root tips were exposed to 300 mOsmol kg<sup>−1</sup> mannitol (−0.75 MPa). Then, comparative extensibility of the intact root tip was measured after a 0.10 MPa osmotic jump. The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>Na<sup>+</sup> flux of the root xylem and the response to an osmotic jump in the wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4. After exposure to 200 mM NaCl for 4 h, the intact root tips of the <span class="html-italic">PtEXPA6</span>-trangenic lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to 300 mOsmol kg<sup>−1</sup> mannitol (−0.75 MPa), followed by a 0.1 MPa osmotic jump. The net Na<sup>+</sup> flux of the root xylem was measured before and after the addition of mannitol and the subsequent 0.1 MPa osmotic jump. The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 12
<p>Na<sup>+</sup> flux of the petiole vascular bundle and the response to an osmotic jump in the wild-type and <span class="html-italic">PtEXPA6</span>-overexpressing lines of 84K and 717-1B4. After exposure to 200 mM NaCl for 4 h, the intact root tips of the <span class="html-italic">PtEXPA6</span>-trangenic lines of 84K (L11, L12, and L13) and 717-1B4 (L9, L15, and L16), and wild-type (WT) were exposed to 300 mOsmol kg<sup>−1</sup> mannitol (−0.75 MPa), followed by a 0.1 MPa osmotic jump. The net Na<sup>+</sup> flux of the petiole vascular bundle was measured before and after the addition of mannitol, and the subsequent 0.1 MPa osmotic jump. The data are means ± SD (<span class="html-italic">n</span> = 3), and the bars with different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 1744 KiB  
Article
Early Growth and Physiological Acclimation to Shade and Water Restriction of Seven Sclerophyllous Species of the Mediterranean Forests of Central Chile
by Marco A. Yáñez, Sergio E. Espinoza, Carlos R. Magni and Eduardo Martínez-Herrera
Plants 2024, 13(17), 2410; https://doi.org/10.3390/plants13172410 - 29 Aug 2024
Viewed by 365
Abstract
The success of using active restoration in Mediterranean-type climate zones mostly depends on an appropriate matching of plant species and specific management prescriptions upon establishment. In this study, we assessed the early growth and short-term physiological acclimation of seven common species found in [...] Read more.
The success of using active restoration in Mediterranean-type climate zones mostly depends on an appropriate matching of plant species and specific management prescriptions upon establishment. In this study, we assessed the early growth and short-term physiological acclimation of seven common species found in the sclerophyllous forests in central Chile to water restriction and shading. We established a nursery experiment that included three treatments (T0: sun-exposed and water-restricted, T1: sun-exposed and fully irrigated, and T2: shaded and fully irrigated) and seven tree species differing in their shade and drought tolerance (Quillaja saponaria Molina, Aristotelia chilensis (Mol.) Stuntz, Peumus boldus Molina, Lithraea caustica (Mol.) Hook. and Arn, Luma apiculata (DC.) Burret, Colliguaja odorifera Molina, and Escallonia pulverulenta (Ruiz and Prav.) Pers). We measured the increment in seedling height and different leaf morpho-physiological traits during two months in the dry season. Based on the measured traits, none of the species took advantage of the higher water availability in T1 relative to T0, but most of the species responded to the shade in T2, regardless of their shade or drought tolerance. Height increments due to shade varied from 0% in P. boldus to 203% in L. apiculata. Overall, all the species responded similarly to the treatments in specific leaf area, chlorophyll content index, photosynthetic rate, stomatal conductance, and intrinsic water use efficiency. This suggests that the species exhibited similar acclimation patterns of these parameters to shade and drought, even regarding the variation in midday xylem water potential found in the water-restricted treatment T0 (from −1.5 MPa in P. boldus to −3.1 MPa in E. pulverulenta). In this study, shading had a higher positive effect on the seedling performance of sclerophyllous species than watering, which at operational level highlights the need for investing in tree shelters when using these species in restoration programs. Full article
(This article belongs to the Special Issue Development of Woody Plants)
Show Figures

Figure 1

Figure 1
<p>Daily mean moisture (<b>A</b>) and temperature (<b>B</b>) of the substrate for the full (T2) and restricted (T0) water treatments during the experimental period.</p>
Full article ">Figure 2
<p>Mean height increment during the experimental period per plant species and treatment (T0, T1, and T2). Different letters denote significant differences among treatments for a specific plant species according to Tukey’s test, with a significance level of 0.05.</p>
Full article ">Figure 3
<p>Mean specific leaf area (SLA) (<b>A</b>) and chlorophyll content index (<b>B</b>) per plant species. Different letters denote significant differences among plant species according to Tukey’s test, with a significance level of 0.05.</p>
Full article ">Figure 4
<p>Means for net saturated photosynthetic rate (<span class="html-italic">A<sub>sat</sub></span>), stomatal conductance (<span class="html-italic">g<sub>s</sub></span>), and intrinsic water use efficiency (WUE<sub>int</sub>) over time per plant species (left panels) and treatment (T0, T1, and T2) (right panels). * indicates significant differences between species and treatments at a specific date, respectively, with a significance level of 0.05.</p>
Full article ">Figure 5
<p>Means for midday xylematic hydric potential per treatment (T0, T1, and T2) over time and separated by species. * indicates significant differences between treatments at a specific date, respectively, with a significance level of 0.05.</p>
Full article ">
18 pages, 8138 KiB  
Article
Genome-Wide Characterization of the BTB Gene Family in Poplar and Expression Analysis in Response to Hormones and Biotic/Abiotic Stresses
by Jing Yue, Xinren Dai, Quanzi Li and Mingke Wei
Int. J. Mol. Sci. 2024, 25(16), 9048; https://doi.org/10.3390/ijms25169048 - 21 Aug 2024
Viewed by 532
Abstract
The BTB (Broad-complex, tramtrack, and bric-a-brac) gene family, characterized by a highly conserved BTB domain, is implicated in a spectrum of biological processes, encompassing growth and development, as well as stress responses. Characterization and functional studies of BTB genes in poplar are still [...] Read more.
The BTB (Broad-complex, tramtrack, and bric-a-brac) gene family, characterized by a highly conserved BTB domain, is implicated in a spectrum of biological processes, encompassing growth and development, as well as stress responses. Characterization and functional studies of BTB genes in poplar are still limited, especially regarding their response to hormones and biotic/abiotic stresses. In this study, we conducted an HMMER search in conjunction with BLASTp and identified 95 BTB gene models in Populus trichocarpa. Through domain motif and phylogenetic relationship analyses, these proteins were classified into eight families, NPH3, TAZ, Ankyrin, only BTB, BACK, Armadillo, TPR, and MATH. Collinearity analysis of poplar BTB genes with homologs in six other species elucidated evolutionary relationships and functional conservations. RNA-seq analysis of five tissues of poplar identified BTB genes as playing a pivotal role during developmental processes. Comprehensive RT-qPCR analysis of 11 BTB genes across leaves, roots, and xylem tissues revealed their responsive expression patterns under diverse hormonal and biotic/abiotic stress conditions, with varying degrees of regulation observed in the results. This study marks the first in-depth exploration of the BTB gene family in poplar, providing insights into the potential roles of BTB genes in hormonal regulation and response to stress. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Chromosome distribution and synteny analysis of <span class="html-italic">BTB</span> genes in <span class="html-italic">P. trichocarpa</span>. (<b>a</b>) The chromosomal mapping of <span class="html-italic">PtrBTB</span> genes across all 19 chromosomes of <span class="html-italic">P. trichocarpa</span>. (<b>b</b>) Comparative analysis of the distribution and syntenic relationships within the <span class="html-italic">PtrBTB</span> gene family, with syntenic gene pairs visually represented by blue connecting lines.</p>
Full article ">Figure 1 Cont.
<p>Chromosome distribution and synteny analysis of <span class="html-italic">BTB</span> genes in <span class="html-italic">P. trichocarpa</span>. (<b>a</b>) The chromosomal mapping of <span class="html-italic">PtrBTB</span> genes across all 19 chromosomes of <span class="html-italic">P. trichocarpa</span>. (<b>b</b>) Comparative analysis of the distribution and syntenic relationships within the <span class="html-italic">PtrBTB</span> gene family, with syntenic gene pairs visually represented by blue connecting lines.</p>
Full article ">Figure 2
<p>A phylogenetic tree of BTB proteins from three plant species, <span class="html-italic">P. trichocarpa</span>, <span class="html-italic">P. alba</span> × <span class="html-italic">P. glandulosa</span>, and Arabidopsis. The analysis shows that 340 BTB proteins are classified into eight subgroups: MATH, Armadillo, BTB-only, TAZ, BACK, Ankyrin, TPR, and NPH3. The phylogenetic tree was constructed using MEGA-X 10.2 software using the neighbor-joining method at 1000 bootstrap replicates.</p>
Full article ">Figure 3
<p>Analysis of the conserved motifs and gene structures within the <span class="html-italic">PtrBTB</span> gene family. The clustering is performed according to the results of phylogenetic analysis. (<b>a</b>) Conserved motifs within <span class="html-italic">PtrBTB</span> proteins, identified by the MEME tool (version 5.5.6), yielding 10 distinct motifs, labeled as Motif 1 through 10. 100 aa is indicated by the scale bar. (<b>b</b>) Gene structure analysis of <span class="html-italic">PtrBTB</span> genes revealed the organization of UTR, intron, and exon regions, with UTR in green, exons in yellow, and introns in grey. The scale bar corresponds to 2 kb.</p>
Full article ">Figure 4
<p>Analysis of <span class="html-italic">cis</span>-elements in the 2000 bp upstream promoter of <span class="html-italic">PtrBTB</span> genes. The clustering is performed according to the results of phylogenetic analysis. Within the predicted promoter region of the <span class="html-italic">PtrBTB</span> genes, a total of 14 distinct regulatory motifs have been identified. These motifs are distinguished by a spectrum of colors, each representing a unique class of transcription factor binding sites.</p>
Full article ">Figure 5
<p>Collinear analysis of <span class="html-italic">PtrBTB</span> genes from <span class="html-italic">P. trichocarpa</span> with four dicotyledons (<span class="html-italic">P. alba × P. glandulosa</span>, <span class="html-italic">A. thaliana</span>, <span class="html-italic">S. lycopersicum</span>, and <span class="html-italic">E. grandis</span>) and two monocotyledons (<span class="html-italic">O. sativa</span> and <span class="html-italic">Z. mays</span>).</p>
Full article ">Figure 6
<p>Expression profiles of <span class="html-italic">PtrBTB</span> genes in different tissues, shoots, roots, leaves, xylem and phloem. Ⅰ−VI represent different clusters. All data of RNA−seq analysis was deposited in GEO (accession number: GSE81077). The color scale represents the FPKM values normalized by log<sub>2</sub>FPKM. Red represents highly expressed genes and blue represents low expressed genes.</p>
Full article ">Figure 7
<p>Expression profiles of <span class="html-italic">BTB</span> genes under exogenous hormones in <span class="html-italic">P. alba × P. glandulosa</span>. Color scale represents log<sub>2</sub> expression values, red represents highly expressed genes and blue represents low expressed genes. (<b>a</b>) JA treatment, (<b>b</b>) ABA treatment, (<b>c</b>) GA treatment, (<b>d</b>) NAA treatment, (<b>e</b>) SA treatment. All samples were collected from leaf tissues at specified time points, with three biological replicates for each treatment. Error bars indicate ± SE of the means (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 8
<p>Expression profiles of <span class="html-italic">PagBTB</span> genes under different biotic/abiotic stress in poplar 84K. Color scale represents log<sub>2</sub> expression values, red represents highly expressed genes and blue represents low expressed genes. (<b>a</b>) NaCl−root, (<b>b</b>) NaCl−leaf, (<b>c</b>) <span class="html-italic">F. solani</span> stress−leaf, (<b>d</b>) Drought treatment−xylem. All samples were collected at the indicated time intervals from three biological replicates of each treatment. Error bars indicate ± SE of the means (<span class="html-italic">n</span> = 3).</p>
Full article ">
20 pages, 12372 KiB  
Article
Influence of Anatomical Spatial Architecture of Pinus devoniana on Pressure Gradients Inferred from Coupling Three-Dimensional CT Imaging and Numerical Flow Simulations
by Juan Gabriel Rivera-Ramos, José Cruz de León, Dante Arteaga, Raúl Espinoza-Herrera, Erica Arreola García, Manuel Arroyo-Albiter and Luis Olmos
Forests 2024, 15(8), 1403; https://doi.org/10.3390/f15081403 - 10 Aug 2024
Viewed by 687
Abstract
Conifer forests in Michoacán are facing climate change. Pinus devoniana Lindley, with natural distribution in the state, has shown certain adaptability, and knowing the influence of anatomy in the flow system is essential to delimit how it contributes to safety margins and water [...] Read more.
Conifer forests in Michoacán are facing climate change. Pinus devoniana Lindley, with natural distribution in the state, has shown certain adaptability, and knowing the influence of anatomy in the flow system is essential to delimit how it contributes to safety margins and water efficiency. For this, the pressure gradients in the cell lumens and their ramifications were analyzed by numerical simulations of flow throughout the real microstructure. Xylem were evaluated in radial, tangential and longitudinal directions. With the skeletonization of lumens and their constrictions, a branching system of interconnection between tracheids, ray cells, intercellular chambers, extensions, and blind pits were identified. In the simulation, the branched system bypasses the longitudinal fluid passage through the pores in membranes of pairs of pits to redirect it through the direct path branching, contributing to safety margins and water efficiency. Thus, resilience at low pressures because of the lower pressure drop in the extensions. The interface between the branching system and the cell lumens are sites of higher pressure gradient, more conducive to water-vapor formation or air leakage in the face of the lowest pressure system. The flow lines move along easy paths, regardless of the simulated flow direction. Deposits in the cell extensions were shown to be attached to the S3 layer of the cell wall, leaving the center of the duct free to flow. It is concluded that the spatial architecture of the xylem anatomy of Pinus dvoniana is a factor in the resilience at low pressures due to high water stress of the species. Full article
Show Figures

Figure 1

Figure 1
<p>Sample preparation: (<b>a</b>) <span class="html-italic">Pinus devoniana</span> tree, (<b>b</b>) sample extraction, and (<b>c</b>) specimens used for the study.</p>
Full article ">Figure 2
<p>CT image processing: (<b>a</b>) initial image, (<b>b</b>) 3D-filtered initial image, (<b>c</b>) binary image, (<b>d</b>) tangential slice, (<b>e</b>) radial slice view of uncompleted fibers due to the angle between the fiber inclination and the crop section, (<b>f</b>) cross section, (<b>g</b>) the yellow rectangle is the ROI extracted from a 3D image acquired with 4 µm voxel, (<b>h</b>) color distribution.</p>
Full article ">Figure 3
<p>Reconstruction of continuity of cell cavities, spaces and cell extensions in microstructure of <span class="html-italic">Pinus devoniana</span> wood CT images: (<b>a</b>) continuity between lumen, pit, extensions spaces and extensions chambers, (<b>b</b>) types of extensions, (<b>c</b>) continuity in crossing fields, (<b>d</b>) chambers between extensions spaces, (<b>e</b>) checking of branches in transverse microstructure, (<b>f</b>) checking of branches in tracheid overlap zone microstructure; Parenchyma cavity (PC), extension (E), tracheid cavity (TC), blind pits (BP), chambers in BS (Ch), branched system (BS).</p>
Full article ">Figure 4
<p>CT images of the microstructure of <span class="html-italic">Pinus devoniana</span> wood: (<b>a</b>) cell cavities in the microstructure, (<b>b</b>) <span class="html-italic">Pinus devoniana</span> cell wall layers identified with CT, (<b>c</b>) connectivity of extensions and wall roughness, (<b>d</b>) tracheid corner extension; middle lamella (ML), extension (E), wood layer (S2) and (S3), parenchyma cavity (PC), tracheid cavity (TC), pit (P).</p>
Full article ">Figure 5
<p>Pressure gradient plots simulated of <span class="html-italic">Pinus devoniana</span> wood microstructure issues from CT images: (<b>a</b>) flow pressure drop on entering the tracheid, (<b>b</b>) pressure gradient in longitudinal flow direction, (<b>c</b>) different pressure gradient on the extensions, (<b>d</b>) similar pressure gradient between epithelial cells, tracheids, pits, and parenchyma; extension (E), tracheid (T), pits (P).</p>
Full article ">Figure 6
<p>Pressure plots in BS of <span class="html-italic">Pinus devoniana</span> wood microstructure from CT images: (<b>a</b>) longitudinal flow pressure drop in tracheids and extensions, (<b>b</b>) pressure drop in BS in a high pressure system; extension (E), tracheid (T), pits (P).</p>
Full article ">Figure 7
<p>Pressure plots in radial flow of <span class="html-italic">Pinus devoniana</span> wood microstructure from CT images: (<b>a</b>) flow pressure drop on entering the tracheid, (<b>b</b>) pressure gradient in radial flow direction; tracheid (T), pits (P).</p>
Full article ">Figure 8
<p>Pressure plots in tangential flow of <span class="html-italic">Pinus devoniana</span> wood microstructure from CT images: (<b>a</b>) pits in crossing fields, (<b>b</b>) pressure gradient in extensions in tangential flow direction; extension (E).</p>
Full article ">Figure 9
<p>Connection of blind pits with extensions on <span class="html-italic">Pinus devoniana</span> wood from CT images: (<b>a</b>) continuity of the tracheid lumen through the extensions, (<b>b</b>) connection of the extensions with the ray; extension (E), tracheid (T).</p>
Full article ">Figure 10
<p>Pressure plots and flow lines inside of the microstructure of <span class="html-italic">Pinus devoniana</span> wood from CT images: (<b>a</b>) pressure within chambers linking the extension connected in turn to blind pit chambers between two tracheids, (<b>b</b>) longitudinal flow lines avoiding the pit pairs and continuing through the extensions; extension (E), tracheid (T), branched system (BS).</p>
Full article ">Figure 11
<p>Microstructure of <span class="html-italic">Pinus devoniana</span> wood in 4 µm voxel from CT images; (<b>a</b>) early wood and late wood, (<b>b</b>) obstacles to flow, (<b>c</b>) relationship between pressure drop and diameter of cell lumens. Late Xylem (XL), Early Xylem (XE).</p>
Full article ">Figure 12
<p>Radial flow in microstructure of <span class="html-italic">Pinus devoniana</span> wood from CT images.</p>
Full article ">Figure 13
<p>Tangential flow in microstructure of <span class="html-italic">Pinus devoniana</span> wood from CT images: (<b>a</b>) flow lines in ray-extension continuation, (<b>b</b>) periodic flow in tracheid rays and extensions; resiniferous channel (Rch), extension (E).</p>
Full article ">
19 pages, 4061 KiB  
Article
Dynamics of Physiological Properties and Endophytic Fungal Communities in the Xylem of Aquilaria sinensis (Lour.) with Different Induction Times
by Qingqing Zhang, Rongrong Li, Yang Lin, Weiwei Zhao, Qiang Lin, Lei Ouyang, Shengjiang Pang and Huahao Zeng
J. Fungi 2024, 10(8), 562; https://doi.org/10.3390/jof10080562 - 9 Aug 2024
Viewed by 703
Abstract
Xylem-associated fungus can secrete many secondary metabolites to help Aquilaria trees resist various stresses and play a crucial role in facilitating agarwood formation. However, the dynamics of endophytic fungi in Aquilaria sinensis xylem after artificial induction have not been fully elaborated. Endophytic fungi [...] Read more.
Xylem-associated fungus can secrete many secondary metabolites to help Aquilaria trees resist various stresses and play a crucial role in facilitating agarwood formation. However, the dynamics of endophytic fungi in Aquilaria sinensis xylem after artificial induction have not been fully elaborated. Endophytic fungi communities and xylem physio-biochemical properties were examined before and after induction with an inorganic salt solution, including four different times (pre-induction (0M), the third (3M), sixth (6M) and ninth (9M) month after induction treatment). The relationships between fungal diversity and physio-biochemical indices were evaluated. The results showed that superoxide dismutase (SOD) and peroxidase (POD) activities, malondialdehyde (MDA) and soluble sugar content first increased and then decreased with induction time, while starch was heavily consumed after induction treatment. Endophytic fungal diversity was significantly lower after induction treatment than before, but the species richness was promoted. Fungal β-diversity was also clustered into four groups according to different times. Core species shifted from rare to dominant taxa with induction time, and growing species interactions in the network indicate a gradual complication of fungal community structure. Endophytic fungi diversity and potential functions were closely related to physicochemical indices that had less effect on the relative abundance of the dominant species. These findings help assess the regulatory mechanisms of microorganisms that expedite agarwood formation after artificial induction. Full article
Show Figures

Figure 1

Figure 1
<p>Number of sequences (<b>A</b>), unique OTUs (<b>B</b>) and diversity indices (<b>C</b>) of endophytic fungi at different induction times. 0M, 3M, 6M and 9M represent pre-induction, the third month, the sixth month and the ninth month after artificial induction, respectively. Different letters indicate significant differences between induction time groups (<span class="html-italic">p</span> &lt; 0.05, Tukey’s test).</p>
Full article ">Figure 2
<p>The number (<b>A</b>,<b>C</b>) and relative abundance (<b>B</b>,<b>D</b>) of dominant fungi at order (<b>A</b>,<b>B</b>) and genus (<b>C</b>,<b>D</b>) levels under different induction times.</p>
Full article ">Figure 3
<p>NMDS ((<b>A</b>), based on Bray–Curtis) and ANOSIM ((<b>B</b>), based on weighted UniFrac distances) analyses of endophytic fungi of <span class="html-italic">A. sinensis</span> at different induction times.</p>
Full article ">Figure 4
<p>Changes in the relative abundance of dominant orders (<b>A</b>,<b>B</b>) and genera (<b>C</b>,<b>D</b>) in different induction times. The relative abundance of each indicator in the above figure was converted to log<sub>10</sub>(X + 1) standards. Different letters indicate significant differences between induction time groups (<span class="html-italic">p</span> &lt; 0.05, Tukey’s test).</p>
Full article ">Figure 5
<p>Co-occurrence network of endophytic fungi at the order level during different induction times. The size and color of each node depend on its abundance and phylum category. The red and green links indicate positive and negative correlations between nodes, respectively. 0M, 3M, 6M and 9M represent pre-induction, the third month, the sixth month and the ninth month after artificial induction, respectively.</p>
Full article ">Figure 6
<p>Function predicted of fungal community based on FUNGuild.</p>
Full article ">Figure 7
<p>The correlation heatmap between dominant species (<b>A</b>: order level) (<b>B</b>: genus level) and physio-biochemical properties in the xylem. SOD, POD and MDA represent superoxide dismutase, peroxidase and malondialdehyde, respectively. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Mantel test among diversity, community composition, function predicted of endophytic fungi and physio-biochemical properties in the xylem. SOD, POD and MDA represent superoxide dismutase, peroxidase and malondialdehyde, respectively. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
18 pages, 8128 KiB  
Article
Selection and Characterization of Somaclonal Variants of Prata Banana (AAB) Resistant to Fusarium Wilt
by Mileide dos Santos Ferreira, Tamyres Amorim Rebouças, Anelita de Jesus Rocha, Wanderley Diaciso dos Santos Oliveira, Ana Carolina Lima Santos dos Santos, João Pedro Falcón Lago de Jesus, Andresa Priscila de Souza Ramos, Claudia Fortes Ferreira, Janay Almeida dos Santos-Serejo, Fernando Haddad and Edson Perito Amorim
Agronomy 2024, 14(8), 1740; https://doi.org/10.3390/agronomy14081740 - 8 Aug 2024
Viewed by 622
Abstract
Fusarium wilt, caused by the fungus Fusarium oxysporum f. sp. cubense (Foc), is one of the most devastating diseases affecting banana cultivation worldwide. Although Foc tropical race 4 (TR4) has not yet been identified in Brazilian production areas, the damage caused by races [...] Read more.
Fusarium wilt, caused by the fungus Fusarium oxysporum f. sp. cubense (Foc), is one of the most devastating diseases affecting banana cultivation worldwide. Although Foc tropical race 4 (TR4) has not yet been identified in Brazilian production areas, the damage caused by races 1 and subtropical 4 is the main cause of production losses, especially affecting cultivars of the Prata subgroup. Thus, the induction of somaclonal variation is a promising strategy in biotechnology to generate genetic variability and develop resistant varieties. This study aimed to induce somaclonal variation in the Prata Catarina cultivar (AAB genome) using successive subcultures in Murashige and Skoog (MS) medium enriched with the plant regulator Thiadizuron (TDZ) at two concentrations: 1 and 2 mg/L. After evaluating the symptoms, we selected 13 resistant somaclones that were not infected by the fungus. Histochemical and histological analyses of the somaclones indicated possible defense mechanisms that prevented colonization and/or infection by Foc, such as intense production of phenolic compounds and the presence of cellulose and callose in the roots. Some somaclones showed no pathogen structures in the xylem-conducting vessels, indicating possible pre-penetration resistance. Furthermore, molecular studies indicated that the genetic alterations in the somaclones may have induced resistance to Foc without compromising the agronomic characteristics of the commercial genotype. Full article
(This article belongs to the Section Crop Breeding and Genetics)
Show Figures

Figure 1

Figure 1
<p>Internal symptoms of Fusarium wilt in the somaclones of Prata Catarina (AAB) banana plants evaluated in the greenhouse. (<b>A</b>) Bar graph with the number of plants with each grade of symptoms according to the grading scale, which varied from 1 to 4, and cross-section of the rhizome with the respective degrees of symptoms. (<b>B</b>) Boxplot of the internal disease symptom indices (DI%). Trat 1: treatment 1, with a TDZ dose of 1 mg/L; Trat 2: treatment 2, with a TDZ dose of 2 mg/L. The treatments differ statistically, as indicated by the letters a, b, and c.</p>
Full article ">Figure 2
<p>Cross-sectional micrographs of the roots of somaclones of the cultivar Prata Catarina, considered resistant to infection by Foc isolate CNPMF 229. The red dots indicate the presence of phenolic compounds. Controls of (<b>A</b>) T1 and (<b>I</b>) T2: (<b>B</b>) S1, (<b>C</b>) S2, (<b>D</b>) S3, (<b>E</b>) S4, (<b>F</b>) S5, (<b>G</b>) S6, and (<b>H</b>) S7 are resistant somaclones in T1; (<b>J</b>) S8, (<b>K</b>) S9, (<b>L</b>) S10, (<b>M</b>) S11, (<b>N</b>) S12, and (<b>O</b>) S13, are resistant somaclones in T2.</p>
Full article ">Figure 3
<p>Fluorescence micrographs of cross-sections of the roots of somaclones of the cultivar Prata Catarina, considered resistant to infection by Foc isolate CNPMF 229. The yellow arrows indicate fluorescent regions with the presence of callose. Controls of (<b>A</b>) T1 and (<b>I</b>) T2: (<b>B</b>) S1, (<b>C</b>) S2, (<b>D</b>) S3, (<b>E</b>) S4, (<b>F</b>) S5, (<b>G</b>) S6, and (<b>H</b>) S7 represent the resistant somaclones in T1; (<b>J</b>) S8, (<b>K</b>) S9, (<b>L</b>) S10, (<b>M</b>) S11, (<b>N</b>) S12, and (<b>O</b>) S13 represent the resistant somaclones in T2.</p>
Full article ">Figure 4
<p>Fluorescence micrographs of cross-sections of the roots of somaclones of the cultivar Prata Catarina, considered resistant to infection by Foc isolate CNPMF 229. The yellow arrows indicate the presence of cellulose. Controls of (<b>A</b>) T1 and (<b>I</b>) T2: (<b>B</b>) S1, (<b>C</b>) S2, (<b>D</b>) S3, (<b>E</b>) S4, (<b>F</b>) S5, (<b>G</b>) S6, and (<b>H</b>) S7represent the resistant somaclones in T1; (<b>J</b>) S8, (<b>K</b>) S9, (<b>L</b>) S10, (<b>M</b>) S11, (<b>N</b>) S12, and (<b>O</b>) S13 represent the resistant somaclones in T2.</p>
Full article ">Figure 5
<p>Micrographs of root fragments from somaclones of the cultivar Prata Catarina, considered resistant to infection by Foc isolate CNPMF 229. The arrows indicate chlamydospores (Chl) and fungal hyphae (Hyp). Controls of (<b>A</b>) T1 and (<b>I</b>) T2: (<b>B</b>) S1, (<b>C</b>) S2, (<b>D</b>) S3, (<b>E</b>) S4, (<b>F</b>) S5, (<b>G</b>) S6, and (<b>H</b>) S7 represent the resistant somaclones in T1; (<b>J</b>) S8, (<b>K</b>) S9, (<b>L</b>) S10, (<b>M</b>) S11, (<b>N</b>) S12, and (<b>O</b>) S13 represent the resistant somaclones in T2.</p>
Full article ">Figure 6
<p>Molecular analysis to identify genetic changes in somaclones resistant to Foc isolate CNPMF 229. (<b>A</b>) Inter-retrotransposon amplified polymorphism (IRAP) markers Sukula + LTR6149 combination; (<b>B</b>) retrotransposon–microsatellite amplified polymorphism (REMAP) markers REMAP: LTR reverse 7286 + 8387; (<b>C</b>) inter-simple sequence repeat (ISSR) markers ISSR-7. 1kb Invitrogem<sup>®®</sup> (Waltham, MA, USA) marker; Prata Catarina cultivar controls (1 and 9), 2: S1, 3: S2, 4: S3, 5: S4, 6: S5, 7: S6, and 8: S7; these correspond to the resistant somaclones in T1. 10: S8, 11: S9, 12: S10, 13: S11, 14: S12, and 15: S13; these correspond to the resistant somaclones in T2.</p>
Full article ">
15 pages, 6966 KiB  
Article
Xylogenesis Responses to a Mediterranean Climate in Holm Oak (Quercus ilex L.)
by Iqra Liyaqat, Angela Balzano, Francesco Niccoli, Jerzy Piotr Kabala, Maks Merela and Giovanna Battipaglia
Forests 2024, 15(8), 1386; https://doi.org/10.3390/f15081386 - 8 Aug 2024
Viewed by 755
Abstract
Quercus ilex L., an evergreen oak species typical of the western and central Mediterranean basin, is facing decline and dieback episodes due to the increase in the severity and frequency of heat waves and drought events. Studying xylogenesis (the wood formation process) is [...] Read more.
Quercus ilex L., an evergreen oak species typical of the western and central Mediterranean basin, is facing decline and dieback episodes due to the increase in the severity and frequency of heat waves and drought events. Studying xylogenesis (the wood formation process) is crucial for understanding how trees respond with their secondary growth to environmental conditions and stress events. This study aimed to characterize the wood formation dynamics of Quercus ilex and their relationship with the meteorological conditions in an area experiencing prolonged drought periods. Cambial activity and xylem cell production were monitored during the 2019 and 2020 growing seasons in a Q. ilex forest located at the Vesuvius National Park (southern Italy). The results highlighted the significant roles of temperature and solar radiation in stimulating xylogenesis. Indeed, the correlation tests revealed that temperature and solar radiation positively influenced growth and cell development, while precipitation had an inhibitory effect on secondary wall formation. The earlier cell maturation in 2020 compared to 2019 underscored the impact of global warming trends. Overall, the trees studied demonstrated good health, growth and adaptability to local environmental fluctuations. This research provides novel insights into the intra-annual growth dynamics of this key Mediterranean species and its adaptation strategies to climatic variability, which will be crucial for forest management in the context of climate change. Full article
(This article belongs to the Section Forest Ecology and Management)
Show Figures

Figure 1

Figure 1
<p>Study site located within Vesuvius National Park, Naples. Red triangle indicates the <span class="html-italic">Quercus ilex</span> stand.</p>
Full article ">Figure 2
<p>Weather conditions of the study site during the monitoring period of 2019 and 2020. In red, the maximum temperature; in black, the average temperature; in grey, the minimum temperature. The blue bars represent precipitation.</p>
Full article ">Figure 3
<p>Number of cambial cells and width of different developmental xylem zones in <span class="html-italic">Quercus ilex</span> trees in 2019 (<b>A</b>–<b>D</b>) and 2020 (<b>E</b>–<b>H</b>): cambial cells (CCs), enlarging post-cambial cells (PCs), cells developing secondary walls (SW) cells and mature (MT) cells with a lignified secondary wall. Mean values are shown for the days of the year (DOY) when the sampling was performed.</p>
Full article ">Figure 4
<p>Final tree ring width (TRW) containing fully mature cells formed for the years 2019 and 2020. Scale bar = 200 µm.</p>
Full article ">Figure 5
<p>Non-collapsed phloem (NCP) width in 2019 (<b>A</b>) and 2020 (<b>B</b>). Mean values are shown on the days of the year (DOY) when the sampling was performed.</p>
Full article ">Figure 6
<p>Correlations between xylogenesis and meteorological conditions. CC = cambial cells (number of cells), PC = post-cambial cells (width, μm), SW = secondary wall-forming cells (width, μm), MT = mature cells (width, μm). Meteorological variables: mean radiation (MJ/day), maximum temperature (°C), mean temperature (°C), minimum temperature (°C), and total precipitation (mm/day). Positive correlations are displayed in red, negative correlations in blue, non-significant ones in grey.</p>
Full article ">
Back to TopTop