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

Next Issue
Volume 15, September
Previous Issue
Volume 15, July
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

Forests, Volume 15, Issue 8 (August 2024) – 203 articles

Cover Story (view full-size image): Infestations have persisted following a sudden and intense outbreak of the bark beetle, Orthotomicus erosus, along the Croatian coast, necessitating a continuous battle against this pest. A recommended protective action is the utilization of pheromone traps for population surveillance. Previous monitoring efforts have recorded an exceptionally high capture rate of natural enemies using pheromone traps; these traps inadvertently prevented natural enemies from fulfilling their essential role in controlling bark beetle populations. To address and significantly diminish instances of this unintended capture, our study designed a modification to the Theysohn-type pheromone trap by integrating a metal mesh within the trapping container. An experimental setup was established in Marjan Forest Park, situated on a peninsula bordered by the sea on three sides and partly by the city of Split. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
15 pages, 3979 KiB  
Article
Accumulation of Glomalin-Related Soil Protein Regulated by Plantation Types and Vertical Distribution of Soil Characteristics in Southern China
by Miaolan Wu, Shaochun Zhang, Xiaojuan Gu, Zhihang He, Yue Liu and Qifeng Mo
Forests 2024, 15(8), 1479; https://doi.org/10.3390/f15081479 - 22 Aug 2024
Viewed by 761
Abstract
The glomalin-related soil protein (GRSP) is an important component of soil organic carbon (SOC), which plays an important role in maintaining soil structural stability, soil carbon (C), and nitrogen (N) fixation. However, little is known about the GRSP content in soil and its [...] Read more.
The glomalin-related soil protein (GRSP) is an important component of soil organic carbon (SOC), which plays an important role in maintaining soil structural stability, soil carbon (C), and nitrogen (N) fixation. However, little is known about the GRSP content in soil and its contribution to soil nutrients in plantations of different tree species. In this study, we determined the soil physicochemical characteristics and GRSP contents in different soil layers of four kinds of plantations, including Acacia mangium (AM), Pinus caribaea (PC), Eucalyptus urophylla (EU), and Magnoliaceae glanca (MG), to address how the plantation types affected the GRSP in different layers of soil in southern China. The results showed that with an increase in soil depth, the GRSP content decreased linearly, and the contribution rate of GRSP to SOC and total nitrogen (TN) in deep soil was 1.08–1.18 times that in surface soil. The tree species significantly affected the vertical distribution of GRSP in soil. Among the four plantations, the conifer species PC had the highest level of GRSP, while the N-fixing species AM had the lowest level. However, SOC, soil capillary porosity (CP), TN, soil water content (SWC), and total phosphorus (TP) were important factors regulating soil GRSP content. Additionally, the regulation effects of soil properties on GRSP were various in surface and deep soil among different plantations. In order to improve soil quality and C sequestration potential, conifer species can be planted appropriately, or conifer species and N-fixing species can be mixed to increase soil nutrient content and enhance soil structure and function in afforestation of southern China. Full article
(This article belongs to the Special Issue Biogeochemical Cycles in Forests)
Show Figures

Figure 1

Figure 1
<p>Geographical location information of the study site. The four points are the locations of the four plantations we studied. AM, <span class="html-italic">Acacia mangium</span>; PC, <span class="html-italic">Pinus caribaea</span>; EU, <span class="html-italic">Eucalyptus urophylla</span>; MG, <span class="html-italic">Magnoliaceae glanca</span>. The map was generated by the ArcGIS 10.6 software.</p>
Full article ">Figure 2
<p>Vertical distribution of easily extractable glomalin-related soil protein (EE-GRSP), total GRSP (T-GRSP), and EE-GRSP/T-GRSP ratio in soil of different plantations. AM, <span class="html-italic">Acacia mangium</span>; PC, <span class="html-italic">Pinus caribaea</span>; EU, <span class="html-italic">Eucalyptus urophylla</span>; MG, <span class="html-italic">Magnoliaceae glanca</span>. Different capital letters indicate that there are significant differences between different tree species at the same soil depth, and different lowercase letters indicate that there are significant differences between the same tree species at different soil depths (<span class="html-italic">p</span> &lt; 0.05). The graphic (<b>a</b>) is EE-GRSP concentration, (<b>b</b>) is T-GRSP concentration, (<b>c</b>) is the ratio of EE-GRSP to T-GRSP.</p>
Full article ">Figure 3
<p>Contributions of easily extracted glomalin-related soil protein (EE-GRSP) and total GRSP (T-GRSP) to SOC and TN in soils of different tree species plantations at different soil depths. AM, <span class="html-italic">Acacia mangium</span>; PC, <span class="html-italic">Pinus caribaea</span>; EU, <span class="html-italic">Eucalyptus urophylla</span>; MG, <span class="html-italic">Magnoliaceae glanca</span>. Different capital letters indicate that there are significant differences between different tree species at the same soil depth, and different lowercase letters indicate that there are significant differences between the same tree species at different soil depths (<span class="html-italic">p</span> &lt; 0.05). The graphic (<b>a</b>) is the contribution of EE-GRSP to SOC, (<b>b</b>) is the contribution of EE-GRSP to SOC, (<b>c</b>) is the contribution of EE-GRSP to TN, (<b>d</b>) is the contribution of T-GRSP to TN.</p>
Full article ">Figure 4
<p>Correlation heat map of soil GRSP and soil physical and chemical factors of four tree species. ‘**‘ indicates that the difference is extremely significant (<span class="html-italic">p</span> &lt; 0.01); ‘*‘ means significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
16 pages, 3392 KiB  
Article
Timber Tracking of Jacaranda copaia from the Amazon Forest Using DNA Fingerprinting
by Lorena Frigini Moro Capo, Bernd Degen, Celine Blanc-Jolivet, Niklas Tysklind, Stephen Cavers, Malte Mader, Barbara Rocha Venancio Meyer-Sand, Kathelyn Paredes-Villanueva, Eurídice Nora Honorio Conorado, Carmen Rosa García-Dávila, Valérie Troispoux, Adline Delcamp and Alexandre Magno Sebbenn
Forests 2024, 15(8), 1478; https://doi.org/10.3390/f15081478 - 22 Aug 2024
Viewed by 1386
Abstract
We investigated the utility of nuclear and cytoplasmic single nucleotide polymorphism (SNP) markers for timber tracking of the intensively logged and commercialized Amazonian tree Jacaranda copaia. Eight hundred and thirty-two trees were sampled (cambium or leaves) from 38 sampling sites in Bolivia, [...] Read more.
We investigated the utility of nuclear and cytoplasmic single nucleotide polymorphism (SNP) markers for timber tracking of the intensively logged and commercialized Amazonian tree Jacaranda copaia. Eight hundred and thirty-two trees were sampled (cambium or leaves) from 38 sampling sites in Bolivia, Brazil, French Guiana, and Peru. A total of 128 SNP markers (113 nuclear, 11 chloroplastic, and 4 mitochondrial) were used for genotyping the samples. Bayesian cluster analyses were carried out to group individuals into homogeneous genetic groups for tests to self-assign groups of individuals or individuals to their population of origin. Cluster analysis based on all the SNP markers detected seven main genetic groups. Genetic differentiation was high among populations (0.484) and among genetic groups (0.415), and populations showed a strong isolation-by-distance pattern. Self-assignment testing of the groups of individuals for all loci was able to determine the population origin of all the samples (accuracy = 100%). Self-assignment tests of individuals were able to assign the origin of 94.5%–100% of individuals (accuracy: 91.7%–100%). Our results show that the use of the 128 SNP markers is suitable to correctly determine the origin of J. copaia timber, and they should be considered a useful tool for customs and local and international police. Full article
(This article belongs to the Special Issue Development of Nuclear SNP Markers for Tracing Timber)
Show Figures

Figure 1

Figure 1
<p>Population structure of <span class="html-italic">Jacaranda copaia</span> in South America, showing the spatial distribution of reporting groups (i.e., genetic clusters) based on CpMtSNP (K = 4) (<b>A</b>), nSNP (K = 9) (<b>B</b>), and nCpMtSNP (K = 8) (<b>C</b>) markers estimated by STRUCTURE (K = 4 to 9).</p>
Full article ">Figure 1 Cont.
<p>Population structure of <span class="html-italic">Jacaranda copaia</span> in South America, showing the spatial distribution of reporting groups (i.e., genetic clusters) based on CpMtSNP (K = 4) (<b>A</b>), nSNP (K = 9) (<b>B</b>), and nCpMtSNP (K = 8) (<b>C</b>) markers estimated by STRUCTURE (K = 4 to 9).</p>
Full article ">Figure 2
<p>Pattern of isolation by distance in <span class="html-italic">Jacaranda copaia</span> population samples (<b>A</b>) and genetic groups determined by cluster analysis using STRUCTURE analysis (<b>B</b>–<b>D</b>). <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">F</mi> </mrow> <mrow> <mi mathvariant="bold-italic">s</mi> <mi mathvariant="bold-italic">t</mi> </mrow> </msub> </mrow> </semantics></math> is the pairwise genetic differentiation between population samples (<b>n</b> &gt; 16 individuals) for all 128 nCpMtSNP markers (<b>A</b>) and the pairwise genetic differentiations between clustered population samples as determined by STRUCTURE analysis for all 128 nCpMtSNP (<b>B</b>), 113 nSNP (<b>C</b>), and 15 CpMtSNP (<b>D</b>) markers. The Spearman correlation coefficient (<math display="inline"><semantics> <mrow> <mi>ρ</mi> </mrow> </semantics></math>) was significantly higher than zero (<span class="html-italic">p</span> &lt; 0.01) between the pairwise <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">F</mi> </mrow> <mrow> <mi mathvariant="bold-italic">s</mi> <mi mathvariant="bold-italic">t</mi> </mrow> </msub> </mrow> </semantics></math> and the geographical distance among populations (<b>A</b>) for all 128 nCpMtSNP markers (0.645), as well as between <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">F</mi> </mrow> <mrow> <mi mathvariant="bold-italic">s</mi> <mi mathvariant="bold-italic">t</mi> </mrow> </msub> </mrow> </semantics></math> and the spatial distance between clustered population samples as determined by STRUCTURE analysis (<b>B</b>–<b>D</b>) for the nCpMtSNPs (0.521) and nSNPs (0.552), whereas non-significant correlation was observed for the CpMtSNPs (0.142).</p>
Full article ">Figure 2 Cont.
<p>Pattern of isolation by distance in <span class="html-italic">Jacaranda copaia</span> population samples (<b>A</b>) and genetic groups determined by cluster analysis using STRUCTURE analysis (<b>B</b>–<b>D</b>). <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">F</mi> </mrow> <mrow> <mi mathvariant="bold-italic">s</mi> <mi mathvariant="bold-italic">t</mi> </mrow> </msub> </mrow> </semantics></math> is the pairwise genetic differentiation between population samples (<b>n</b> &gt; 16 individuals) for all 128 nCpMtSNP markers (<b>A</b>) and the pairwise genetic differentiations between clustered population samples as determined by STRUCTURE analysis for all 128 nCpMtSNP (<b>B</b>), 113 nSNP (<b>C</b>), and 15 CpMtSNP (<b>D</b>) markers. The Spearman correlation coefficient (<math display="inline"><semantics> <mrow> <mi>ρ</mi> </mrow> </semantics></math>) was significantly higher than zero (<span class="html-italic">p</span> &lt; 0.01) between the pairwise <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">F</mi> </mrow> <mrow> <mi mathvariant="bold-italic">s</mi> <mi mathvariant="bold-italic">t</mi> </mrow> </msub> </mrow> </semantics></math> and the geographical distance among populations (<b>A</b>) for all 128 nCpMtSNP markers (0.645), as well as between <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">F</mi> </mrow> <mrow> <mi mathvariant="bold-italic">s</mi> <mi mathvariant="bold-italic">t</mi> </mrow> </msub> </mrow> </semantics></math> and the spatial distance between clustered population samples as determined by STRUCTURE analysis (<b>B</b>–<b>D</b>) for the nCpMtSNPs (0.521) and nSNPs (0.552), whereas non-significant correlation was observed for the CpMtSNPs (0.142).</p>
Full article ">
20 pages, 5183 KiB  
Article
Spatial Pattern of Drought-Induced Mortality Risk and Influencing Factors for Robinia pseudoacacia L. Plantations on the Chinese Loess Plateau
by Zhong-Dian Zhang, Tong-Hui Liu, Ming-Bin Huang, Xiao-Ying Yan, Ming-Hua Liu, Jun-Hui Yan, Fei-Yan Chen, Wei Yan and Ji-Qiang Niu
Forests 2024, 15(8), 1477; https://doi.org/10.3390/f15081477 - 22 Aug 2024
Viewed by 765
Abstract
During the large-scale vegetation restoration on the Loess Plateau, the introduction of exotic species with high water consumption, such as Robinia pseudoacacia L., led to widespread soil desiccation, and resulted in severe drought stress and increasing risk of forest degradation and mortality. Accurate [...] Read more.
During the large-scale vegetation restoration on the Loess Plateau, the introduction of exotic species with high water consumption, such as Robinia pseudoacacia L., led to widespread soil desiccation, and resulted in severe drought stress and increasing risk of forest degradation and mortality. Accurate assessment of drought-induced mortality risk in plantation forests is essential for evaluating and enhancing the sustainability of ecological restoration, yet quantitative research at the regional scale on the Loess Plateau is lacking. With a focus on Robinia pseudoacacia L. plantations, we utilized a coupled model of the Biome BioGeochemical Cycles model and plant supply–demand hydraulic model (BBGC-SPERRY model) to simulate the dynamics of the annual average percentage loss of whole-plant hydraulic conductance (APLK) at 124 meteorological stations over an extended period (1961–2020) to examine changes in plant hydraulic safety in Robinia pseudoacacia L. plantations. Based on the probability distribution of APLK at each site, the drought-induced mortality risk probability (DMRP) in Robinia pseudoacacia L. was determined. The results indicate the BBGC-SPERRY model could effectively simulate the spatiotemporal variations in transpiration and evapotranspiration in Robinia pseudoacacia L. stands on the Loess Plateau. The mean APLK and DMRP exhibited increasing trends from southeast to northwest along a precipitation gradient, with their spatial patterns on the Loess Plateau mainly driven by mean annual precipitation and also significantly influenced by other climatic and soil factors. The low-risk (DMRP < 2%), moderate-risk (2% ≤ DMRP ≤ 5%), and high-risk (DMRP > 5%) zones for drought-induced mortality in Robinia pseudoacacia L. accounted for 60.0%, 30.7%, and 9.3% of the study area, respectively. These quantitative findings can provide an important basis for rational forestation and sustainable vegetation management on the Loess Plateau. Full article
(This article belongs to the Section Forest Hydrology)
Show Figures

Figure 1

Figure 1
<p>Map depicting the location of the study area on the Loess Plateau and the distributions of mean annual precipitation (MAP, 1961–2020), 124 modeling sites, and 11 model evaluation sites in the region. ET<sub>a</sub>, annual actual evapotranspiration during the growing season.</p>
Full article ">Figure 2
<p>Flowchart depicting the methodology and data sources of the study. Firstly, the performance of the BBGC-SPERRY model was evaluated with data of annual actual evapotranspiration during the growing season (ET<sub>a</sub>) and daily stem sap flux density collected from previous studies. Then, we simulated the long-term dynamics of plant hydraulic traits with 60-year meteorological data and soil hydraulic parameters at 124 sites by the BBGC-SPERRY model. Finally, based on temporal changes of plant hydraulic traits at each site and plant hydraulic threshold, we evaluated the spatial pattern of drought-induced mortality risk for <span class="html-italic">Robinia pseudoacacia</span> L. plantations using a probabilistic approach.</p>
Full article ">Figure 3
<p>Flow chart of the BBGC-SPERRY model. The soil–plant–atmosphere continuum water transport is simulated with a plant supply–demand hydraulic model (Sperry model) for the coupled simulation of plant hydraulic traits with water, carbon, and nitrogen cycles in the Biome BioGeochemical Cycles (Biome-BGC) model. APLK, annual average percentage loss of whole-plant hydraulic conductance. HR, heterotrophic respiration.</p>
Full article ">Figure 4
<p>Schematic diagram of the determination of drought-induced mortality risk probability (shaded area) based on the fitted probability distribution of annual average percentage loss of whole-plant hydraulic conductance (APLK).</p>
Full article ">Figure 5
<p>Comparison of simulated versus observed annual actual evapotranspiration during the growing season (ET<sub>a</sub>) in <span class="html-italic">Robinia pseudoacacia</span> L. stands.</p>
Full article ">Figure 6
<p>Spatial distribution of minimum (<b>a</b>), maximum (<b>b</b>), and mean (<b>c</b>) annual average percentage loss of whole-plant hydraulic conductance (APLK) for <span class="html-italic">Robinia pseudoacacia</span> L. in the study area.</p>
Full article ">Figure 7
<p>Correlations between environmental factors, minimum, maximum, and mean annual average percentage loss of whole-plant hydraulic conductance (APLK), and drought-induced mortality risk probability (DMRP) for <span class="html-italic">Robinia pseudoacacia</span> L. MAP, mean annual precipitation; and MAT, mean annual temperature. Color gradient and size of the symbols indicate the correlation coefficient among different indicators. * means <span class="html-italic">p</span> &lt; 0.05; ** means <span class="html-italic">p</span> &lt; 0.01; and *** means <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Spatial distribution of drought-induced mortality risk probability (DMRP) of <span class="html-italic">Robinia pseudoacacia</span> L. in the study area.</p>
Full article ">
19 pages, 17938 KiB  
Article
Land Surface Temperature May Have a Greater Impact than Air Temperature on the Autumn Phenology in the Tibetan Plateau
by Hanya Tang, Xizao Sun, Xuelin Zhou, Cheng Li, Lei Ma, Jinlian Liu, Zhi Ding, Shiwei Liu, Pujia Yu, Luyao Jia and Feng Zhang
Forests 2024, 15(8), 1476; https://doi.org/10.3390/f15081476 - 22 Aug 2024
Viewed by 732
Abstract
The Tibetan Plateau (TP), with its unique geographical and climatic conditions, holds a significant role in global climate change. Therefore, it is particularly urgent to fully understand its vegetation phenology. Herbaceous plants are widely distributed in the TP. However, previous studies have predominantly [...] Read more.
The Tibetan Plateau (TP), with its unique geographical and climatic conditions, holds a significant role in global climate change. Therefore, it is particularly urgent to fully understand its vegetation phenology. Herbaceous plants are widely distributed in the TP. However, previous studies have predominantly examined the impact of air temperature on the end date of the vegetation growing season (EOS), with less emphasis on the influence of land surface temperature (LST). In this study, the dynamic changes in the EOS from 2001 to 2020 were analyzed by utilizing the Normalized Difference Vegetation Index (NDVI) data published by NASA. Furthermore, the impact of climate change on the EOS was examined, and the dominant factor (air temperature, LST, or precipitation) influencing the EOS was identified. The main findings were as follows: the average annual EOS predominantly occurred between day of year (DOY) 240 and 280, with an advance from the edge of the plateau to the center. The EOS across the entire region displayed a marginal tendency towards delay, with an average rate of 0.017 days/year. Among all vegetation, shrubs showed the most pronounced delay at a rate of 0.04 days/year. In terms of precipitation, the impact of climate change increased precipitation in both summer and autumn, which could delay EOS. In terms of temperature, an increase in summer Tmin, autumn air temperatures and summer LST delayed the EOS, while an increase in autumn LST advanced the EOS. Compared to air temperature and precipitation, LST had a stronger controlling effect on the EOS (the largest pixel area dominated by LST). These results could offer new insights for enhancing the parameters of vegetation phenology models across the TP. Full article
(This article belongs to the Special Issue Application of Remote Sensing in Vegetation Dynamic and Ecology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The location and elevation of the TP within China; (<b>b</b>) the distribution of land cover types.</p>
Full article ">Figure 2
<p>Valid areas in the TP.</p>
Full article ">Figure 3
<p>Concept map for phenological parameter extraction.</p>
Full article ">Figure 4
<p>(<b>a</b>) The spatial distribution of annual mean EOS; (<b>b</b>) the cumulative percentage of the EOS pixels for the four vegetation types.</p>
Full article ">Figure 5
<p>Spatial distribution of EOS trend during 2001–2020. The picture in the lower left corner shows the significant level of the trend.</p>
Full article ">Figure 6
<p>The spatial distribution of (<b>a</b>) the H values of the EOS; (<b>b</b>) the future trend of the EOS.</p>
Full article ">Figure 7
<p>Relationship between EOS and altitude in (<b>a</b>) whole region (<b>b</b>) forest (<b>c</b>) meadow (<b>d</b>) shrub (<b>e</b>) steppe of the TP.</p>
Full article ">Figure 8
<p>The spatial distribution pattern of partial correlation coefficient between EOS and summer climate factors across the TP from 2001 to 2020. The climate variable contains (<b>a</b>) precipitation, (<b>b</b>) T<sub>max</sub>, (<b>c</b>) T<sub>min</sub>, (<b>d</b>) daytime LST, (<b>e</b>) nighttime LST. The pictures in the higher right corner show the significant partial correlations, with the color blue representing negative partial correlations, and the color red representing positive partial correlations.</p>
Full article ">Figure 9
<p>The spatial distribution pattern of partial correlation coefficient between EOS and autumn climate factors across the TP from 2001 to 2020. The pictures in the higher right corner show the significant partial correlations (the same as <a href="#forests-15-01476-f008" class="html-fig">Figure 8</a>).</p>
Full article ">Figure 10
<p>Spatial distribution pattern of the dominant climatic factor on EOS (<b>a</b>) in summer and (<b>b</b>) in autumn based on the maximum partial correlation coefficient.</p>
Full article ">Figure 11
<p>Comparison of the EOS results of the study with the EOS values of other phenological products.</p>
Full article ">
26 pages, 18602 KiB  
Article
Integration of Phenotypes, Phytohormones, and Transcriptomes to Elucidate the Mechanism Governing Early Physiological Abscission in Coconut Fruits (Cocos nucifera L.)
by Lilan Lu, Zhiguo Dong, Xinxing Yin, Siting Chen and Ambreen Mehvish
Forests 2024, 15(8), 1475; https://doi.org/10.3390/f15081475 - 22 Aug 2024
Viewed by 896
Abstract
The abscission of fruits has a significant impact on yield, which in turn has a corresponding effect on economic benefits. In order to better understand the molecular mechanism of early coconut fruit abscission, the morphological and structural characteristics, cell wall hydrolysis and oxidase [...] Read more.
The abscission of fruits has a significant impact on yield, which in turn has a corresponding effect on economic benefits. In order to better understand the molecular mechanism of early coconut fruit abscission, the morphological and structural characteristics, cell wall hydrolysis and oxidase activities, phytohormones, and transcriptomes were analyzed in the abscission zone (AZ) from early-abscised coconut fruits (AFs) and non-abscised coconut fruits (CFs). These results indicated that the weight and water content of AFs are significantly lower than those of CFs, and the color of AFs is a grayish dark red, with an abnormal AZ structure. Cellulase (CEL), polygalacturonase (PG), pectinesterase (PE), and peroxidase (POD) activities were significantly lower than those of CFs. The levels of auxin (IAA), gibberellin (GA), cytokinins (CKs), and brassinosteroid (BR) in AFs were significantly lower than those in CFs. However, the content of abscisic acid (ABA), ethylene (ETH), jasmonic acid (JA), and salicylic acid (SA) in AFs was significantly higher than in CFs. The transcriptome analysis results showed that 3601 DEGs were functionally annotated, with 1813 DEGs upregulated and 1788 DEGs downregulated. Among these DEGs, many genes were enriched in pathways such as plant hormone signal transduction, carbon metabolism, peroxisome, pentose and gluconate interconversion, MAPK signaling pathway—plant, and starch and sucrose metabolism. Regarding cell wall remodeling-related genes (PG, CEL, PE, POD, xyloglucan endoglucosidase/hydrogenase (XTH), expansin (EXP), endoglucanase, chitinase, and beta-galactosidase) and phytohormone-related genes (IAA, GA, CKs, BR, ABA, JA, SA, and ETH) were significantly differentially expressed in the AZ of AFs. Additionally, BHLH, ERF/AP2, WRKY, bZIP, and NAC transcription factors (TFs) were significantly differently expressed, reflecting their crucial role in regulating the abscission process. This study’s results revealed the molecular mechanism of early fruit abscission in coconuts. This provided a new reference point for further research on coconut organ development and abscission. Full article
(This article belongs to the Section Genetics and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Development of early-abscised coconut fruits (AFs) and non-abscised coconut fruits (CFs). (<b>a</b>) The anatomical structure of fruit morphology and abscission zone (AZ). (<b>b</b>) Fresh weight and water content of coconut fruits. The data represent the mean ± standard deviation (SD) of ten samples, and the significance of fresh weight and water content of coconut fruits between CFs and AFs was determined using Student’s <span class="html-italic">t</span>-tests. * significant at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Enzyme activity and phytohormones in AZ of AFs and CFs. CEL, cellulase; PG, polygalacturonase; PE, pectinesterase; POD, peroxidase; IAA, auxin; GA, gibberellin; CK, cytokinin; BR, brassinosteroid; ABA, abscisic acid; ETH, ethylene; JA, jasmonic acid; SA, salicylic acid. The data represent the mean ± standard deviation (SD) of three biological replicates, and the significance of enzyme activity and plant hormone contents in AZ between CFs and AFs was determined using Student’s <span class="html-italic">t</span>-tests. * significant at <span class="html-italic">p</span> &lt; 0.05. ** significant at <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Differentially expressed genes (DEGs) enriched in the top 20 enriched GO terms in terms of biological processes, molecular functions, and cellular components in AFs and CFs.</p>
Full article ">Figure 4
<p>KEGG analysis from CF vs. AF group. (<b>a</b>) KEGG classification chart of DEGs. (<b>b</b>) KEGG enrichment bar chart of all DEGs. (<b>c</b>) KEGG enrichment bar chart of upregulated DEGs. (<b>d</b>) KEGG enrichment bar chart of downregulated DEGs.</p>
Full article ">Figure 5
<p>Expression of identified DEGs involved in the main KEGG enrichment pathways in CF vs. AF group.</p>
Full article ">Figure 6
<p>Expression of identified DEGs involved in cell wall modification in CF vs. AF group.</p>
Full article ">Figure 7
<p>Heatmap of relative changes in expression patterns of 8 phytohormone-related genes in AZ of CF vs. AF group. The color scales on each heatmap display their expression values.</p>
Full article ">Figure 8
<p>Transcription factor analysis in CF vs. AF group. (<b>a</b>) Expression of identified DEGs involved in transcription factors in CF vs. AF group. (<b>b</b>) Distribution of overexpression of the regulatory transcription factor family in CF vs. AF group.</p>
Full article ">Figure 9
<p>Verification of the expression of 11 coconut fruit abscission-related genes through qRT-PCR analysis. The bar chart represents the value of FPKM. The line graph represents qRT-PCR values. The error bar represents the standard deviation of three biological replicates (<b>a</b>–<b>k</b>). Correlation of expression changes observed through RNA-seq (<span class="html-italic">y</span>-axis) and qRT-PCR (<span class="html-italic">x</span>-axis) (<b>l</b>).</p>
Full article ">Figure 10
<p>A hypothetical model for coconut fruit abscission.</p>
Full article ">
21 pages, 11132 KiB  
Article
Construction of a Genetic Transformation System for Populus wulianensis
by Yan Wang, Chenxia Song, Yi Han, Ruilong Wang, Lingshan Guan, Yanjuan Mu, Tao Sun, Xiaoman Xie, Yunchao Zhao, Jichen Xu and Yizeng Lu
Forests 2024, 15(8), 1474; https://doi.org/10.3390/f15081474 - 22 Aug 2024
Viewed by 870
Abstract
Transgenic technology is a potent tool for verifying gene functions, and poplar serves as a model system for genetically transforming perennial woody plants. However, the current poplar genetic transformation system is limited to a few genotypes. In this study, we developed an efficient [...] Read more.
Transgenic technology is a potent tool for verifying gene functions, and poplar serves as a model system for genetically transforming perennial woody plants. However, the current poplar genetic transformation system is limited to a few genotypes. In this study, we developed an efficient transformation system based on the Agrobacterium-mediated transformation of Populus wulianensis, a rare and endangered tree species endemic to Shandong Province. Aseptic seedlings of P. wulianensis were used as experimental materials, and the optimal medium for inducing adventitious buds was explored as 1/2(NH4NO3) MS + 0.05 mg/L naphthalene acetic acid (NAA) + 0.5 mg/L 6-benzylaminopurine (6-BA), resulting in up to 35 adventitious buds. The selection resistance critical pressure of 300 mg/L for timentin can effectively inhibit the growth of Agrobacterium while promoting the induction of adventitious buds in leaves. The critical screening pressure for kanamycin for producing resistant adventitious buds and inducing resistant rooting seedlings was 100 mg/L. We optimized several independent factors, which significantly enhanced the efficiency of genetic transformation. The leaves were infected with Agrobacterium suspension diluted twice by adding 100 μmol/L acetylsyringone (β-AS) (OD600 = 0.6) for 15 min, followed by co-culture in the dark for 3 d. Using this improved transformation system, we obtained transgenic P. wulianensis clones overexpressing the enhanced green fluorescent protein (EGFP) gene through direct organogenesis. Among the 112 resistant buds obtained, 17 developed resistant rooting in seedlings. Eight positive plants were identified through DNA, RNA, and protein level analyses, with a positivity rate of 47.06%. This study provides a foundation for developing and utilizing P. wulianensis germplasm resources and lays the groundwork for resource improvement. Full article
(This article belongs to the Section Genetics and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Diagram of the blade cutting area. A, B, and C represent the tip, middle, and base parts of the leaf, respectively.</p>
Full article ">Figure 2
<p>Expression vector PEZR—(K)—LC map.</p>
Full article ">Figure 3
<p>Adventitious bud induction. (<b>A</b>,<b>B</b>) were the best induction medium for leaf differentiation of adventitious buds and the optimal induction medium for 35 d elongation of adventitious buds, respectively.</p>
Full article ">Figure 4
<p>Effects of timentin concentrations on <span class="html-italic">Agrobacterium</span> Growth. (<b>A</b>–<b>F</b>) were 0, 50, 100, 150, 200, and 250 mg/L, respectively.</p>
Full article ">Figure 5
<p>Effects of timentin concentrations on the induction of adventitious buds. (<b>i</b>,<b>ii</b>) indicate adventitious bud results in the tip part of the leaf and the middle/base part of the leaf differentiation after the addition of timentin, respectively. (A–S) timentin with concentrations of 0, 50, 100, 150, 200, 250, 300, 350, 400, 450, 500, 550, 600, 650, 700, 750, 800, and 850 mg/L was added to the medium.</p>
Full article ">Figure 6
<p>Effects of timentin concentrations on the rooting of adventitious buds. (A–I) timentin with concentrations of 0, 100, 200, 300, 400, 500, 600, 700, and 800 mg/L was added to the medium. Two replicates were used for each gradient in the figure.</p>
Full article ">Figure 7
<p>Effects of Kan concentrations on adventitious budding induction in the leaves. (A–I) Kan with concentrations of 0, 25, 50, 75, 100, 125, 150, 175, and 200 mg/L was added to the medium. Two replicates were used for each gradient in the figure.</p>
Full article ">Figure 8
<p>Effects of Kan concentrations on adventitious budding induction from rooting. (A–I) Kan with concentrations of 0, 25, 50, 75, 100, 125, 150, 175, and 200 mg/L was added to the medium. Each gradient was repeated twice.</p>
Full article ">Figure 9
<p>Leaf discs after infection with a high concentration of <span class="html-italic">Agrobacterium</span> (OD<sub>600</sub> = 0.6) and culturing for 10 d.</p>
Full article ">Figure 10
<p>Leaf discs were cultured for 1 d, and adventitious buds were differentiated.</p>
Full article ">Figure 11
<p>Rooted seedlings for Kan resistance screening.</p>
Full article ">Figure 12
<p>Transgenic <span class="html-italic">P. wulianensis</span>. (<b>A</b>,<b>B</b>) represent the electropherograms of seedlings transplanted, resistant rooting seedlings, and DNA molecular identification, respectively. M, Marker. PC, Positive plants. WT, Wild type. H<sub>2</sub>O, negative control. Numbers 1–17, transgenic lines L1–L17.</p>
Full article ">Figure 13
<p>Semi-quantitative PCR detection of <span class="html-italic">EGFP</span> in transgenic <span class="html-italic">P. wulianensis</span>. L1–5, L13, L15, and L17 were identified as positive clonal clones at the DNA level. WT, wild type.</p>
Full article ">Figure 14
<p>Monitoring of EGFP expression. The brightfield, EGFP fluorescence, DAPI fluorescence, and the combined field (merged EGFP fluorescence, DAPI fluorescence, and brightfield) were present for the fusion protein (EGFP) (<b>up</b>) and control (WT) (<b>down</b>), respectively. Arrows in the same direction indicate the expression of EGFP in the nucleus.</p>
Full article ">
20 pages, 7365 KiB  
Article
Increased Soil Moisture in the Wet Season Alleviates the Negative Effects of Nitrogen Deposition on Soil Microbial Communities in Subtropical Evergreen Broad-Leaved Forest
by Wen Chen, Zheng Hou, Donghui Zhang, Leixi Chen, Keqin Wang and Yali Song
Forests 2024, 15(8), 1473; https://doi.org/10.3390/f15081473 - 21 Aug 2024
Viewed by 794
Abstract
The rapid increase in reactive nitrogen (N) released into the environment by human activities has notably altered the structure and composition of forest soil microbial communities (SMCs), profoundly impacting the N cycle in terrestrial ecosystems. However, the response of soil microorganisms to nitrogen [...] Read more.
The rapid increase in reactive nitrogen (N) released into the environment by human activities has notably altered the structure and composition of forest soil microbial communities (SMCs), profoundly impacting the N cycle in terrestrial ecosystems. However, the response of soil microorganisms to nitrogen addition in different seasons is not clear. This study delved into how SMCs in a subtropical region of central Yunnan, China, specifically in an evergreen broad-leaved forest (EBLF), respond to N deposition during both the dry and wet seasons. Through high-throughput sequencing, we assessed the composition and structure of SMCs under varying N addition treatments across seasons, examining their interplay with soil chemical properties, enzyme activities, and community responses. The findings revealed significant outcomes following four years of N addition in the subtropical EBLF: (1) Significant changes were observed due to the interaction between N addition and seasonal changes. Soil pH significantly decreased, indicating increased soil acidification, particularly in the dry season. Increased moisture in the wet season mitigated soil acidification. (2) In the dry season, N addition led to a decrease in microbial richness and diversity. In the wet season, N addition increased microbial richness and diversity, alleviating the downward trend observed in the dry season. (3) N addition significantly impacted the composition of soil bacterial and fungal communities. Dominant fungal genera in the wet season were particularly sensitive to N addition. (4) Seasonal changes and N addition altered soil microbial community structures. Soil chemical properties and enzyme activities significantly influenced the microbial community structure. However, due to differences in soil moisture, the key environmental factors that regulate microbial communities have changed significantly during the dry and wet seasons. This study serves as a foundation for understanding how N deposition impacts SMCs in EBLF ecosystems in subtropical regions, offering valuable insights for the scientific management of forest ecological resources amidst global change trends. Full article
(This article belongs to the Special Issue Forest Soil Microbiology and Biogeochemistry)
Show Figures

Figure 1

Figure 1
<p>Overview map of the research area.</p>
Full article ">Figure 2
<p>Soil chemical properties under different nitrogen addition treatments in dry and wet seasons. Different capital letters indicate significant differences between the same nitrogen treatments in different seasons, and different lowercase letters indicate significant differences between different nitrogen treatments in the same season (<span class="html-italic">p</span> &lt; 0.05). The meanings of the abbreviations in the figure are shown in <a href="#forests-15-01473-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>Soil enzyme activities under different nitrogen addition treatments in dry and wet seasons. Different capital letters above the bars indicate significant differences between the same treatment in different seasons, and different lowercase letters indicate significant differences between different treatments in the same season (<span class="html-italic">p</span> &lt; 0.05). The meanings of the abbreviations in the figure are shown in <a href="#forests-15-01473-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 4
<p>Microbial α-diversity under different nitrogen addition treatments in dry and wet seasons. Different capital letters above the bars indicate significant differences between the same treatment in different seasons, and different lowercase letters indicate significant differences between different treatments in the same season (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Principal coordinate analysis of bacteria in dry and wet seasons (<b>a</b>), <span class="html-italic">p</span> &lt; 0.001 ***; principal coordinate analysis of fungi in dry and wet seasons (<b>b</b>), <span class="html-italic">p</span> &lt; 0.001 ***; principal coordinate analysis of bacteria in different nitrogen treatments (<b>c</b>), <span class="html-italic">p</span> = 0.004 **; Principal coordinate analysis of fungi under different nitrogen treatments (<b>d</b>), <span class="html-italic">p</span> = 0.023 *. The horizontal and vertical axes of PCoA represent the two selected principal coordinate axes, and the percentage represents the interpretability value of the principal coordinate axis on the difference in sample composition; the scales of the horizontal and vertical axes are relative distances and have no practical meaning; points of different colors for samples representing different groups, the closer the two sample points are, the more similar the species composition of the two samples is.</p>
Full article ">Figure 6
<p>Relative abundance of bacterial phyla in the dry season (<b>a</b>); relative abundance of fungal phyla in the dry season (<b>b</b>); relative abundance of bacterial phyla in the wet season (<b>c</b>); relative abundance of fungal phyla in the wet season (<b>d</b>).</p>
Full article ">Figure 7
<p>Relative abundance of the bacterial genus in the dry season (<b>a</b>); relative abundance of the fungal genus in the dry season (<b>b</b>); relative abundance of the bacterial genus in the wet season (<b>c</b>); relative abundance of the fungal genus in the wet season (<b>d</b>). Only bacterial genera with a relative abundance &gt;1% are shown in the figure, and the rest are classified as other.</p>
Full article ">Figure 8
<p>Variance partitioning analysis (VPA) shows the explanation of soil nutrients, enzyme activity, and microbial diversity on the variance of bacterial (<b>a</b>) and fungal (<b>b</b>) communities.</p>
Full article ">Figure 9
<p>The Mantel test was used to explore the relationship between the microbial community structure and soil environmental factors in the dry and wet seasons (<b>a</b>,<b>c</b>), and RDA analysis was used to determine the main environmental factors controlling the microbial community structure in the dry and wet seasons (<b>b</b>,<b>d</b>). The Pearson correlation coefficient between different chemical properties, enzyme activities, and microbial community traits is shown in the right triangle diagram, and the correlation results of the Mantel test are shown in the right line diagram. The meanings of the abbreviations in the figure are shown in <a href="#forests-15-01473-t002" class="html-table">Table 2</a>.</p>
Full article ">
19 pages, 10880 KiB  
Article
Satellite Evidence for Increasing in Terrestrial Evapotranspiration over the Contiguous United States from 2001 to 2022
by Lu Liu, Yunjun Yao, Yufu Li, Zijing Xie, Jing Ning, Ruiyang Yu, Jiahui Fan, Yixi Kan, Luna Zhang and Jia Xu
Forests 2024, 15(8), 1472; https://doi.org/10.3390/f15081472 - 21 Aug 2024
Viewed by 728
Abstract
Evapotranspiration (ET) is a key process in the eco-hydrological cycle of a basin and a reliable indicator of climate change. However, the spatiotemporal alterations of ET in the contiguous United States (CONUS) over the recent two decades remain largely uncertain. In this study, [...] Read more.
Evapotranspiration (ET) is a key process in the eco-hydrological cycle of a basin and a reliable indicator of climate change. However, the spatiotemporal alterations of ET in the contiguous United States (CONUS) over the recent two decades remain largely uncertain. In this study, we used the recently proposed Priestley–Taylor (PT)-SinRH model to estimate the ET of CONUS during 2001–2022 based on satellite and reanalysis data. The results showed that the PT-SinRH model yielded superior performance at eddy covariance (EC) sites, and the root-mean-square error (RMSE) ranged from 6.0 to 33.5 W/m2, the Kling–Gupta efficiency (KGE) ranged from 0.22 to 0.66. The annual mean value of ET in CONUS from 2001 to 2022, estimated by the PT-SinRH model, was 42.54 W/m2, and the spatial pattern of seasonal and annual ET variations increased from west to east. From 2001 to 2022, seasonal and annual ET of CONUS showed linear trends, with an average increase of 0.76 W/m2/da (p < 0.05). The ET in the east of CONUS exhibited a rate of increase at 1.45 W/m2/da, and the ET in the west of CONUS exhibited a rate of increase at 0.42 W/m2/da (p < 0.05). Importantly, our analysis of ET trends highlights that the change of precipitation (P) and normalized difference vegetation index (NDVI) exerts a significant impact on the change of ET over CONUS. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Study area.</p>
Full article ">Figure 2
<p>Scatterplots of predicted ET and daily observed ET based on satellite and reanalysis data input at various land cover types.</p>
Full article ">Figure 3
<p>Annual spatial patterns of ET over the CONUS during 2001–2022.</p>
Full article ">Figure 4
<p>Seasonal spatial patterns of ET over the CONUS from 2001 to 2022. MAM (spring): March, April, and May; JJA (summer): June, July, and August; SON (Fall): September, October, and November; DJF (Winter): December, January, and February.</p>
Full article ">Figure 5
<p>Interannual variability of ET from 2001 to 2022 over the (<b>a</b>) CONUS, (<b>b</b>) west of CONUS, and (<b>c</b>) east of CONUS.</p>
Full article ">Figure 6
<p>Interannual seasonal variability of ET from 2001 to 2022 over the CONUS [(<b>a</b>) MAM, (<b>b</b>) JJA, (<b>c</b>) SON, and (<b>d</b>) DJF], west of CONUS [(<b>a1</b>) MAM, (<b>b1</b>) JJA, (<b>c1</b>) SON, and (<b>d1</b>) DJF], and east of CONUS [(<b>a2</b>) MAM, (<b>b2</b>) JJA, (<b>c2</b>) SON, and (<b>d2</b>) DJF].</p>
Full article ">Figure 7
<p>Annual spatiotemporal variations of ET over the CONUS from 2001 to 2022 (<span class="html-italic">p</span> &lt; 0.05). Black dots mean the regions have been checked by significant tests at a 95% confidence level.</p>
Full article ">Figure 8
<p>Seasonal spatiotemporal variations of ET over the CONUS from 2001 to 2022. Black dots mean the regions have been checked by significant tests at a 95% confidence level.</p>
Full article ">Figure 9
<p>Annual spatial patterns of NDVI, precipitation (P), and temperature (Ta) over the CONUS from 2001 to 2022.</p>
Full article ">Figure 10
<p>Annual temporal variations of precipitation (P), temperature (Ta), and NDVI over the CONUS [(<b>a</b>–<b>c</b>)], west of CONUS [(<b>a1</b>–<b>c1</b>)], and east of CONUS [(<b>a2</b>–<b>c2</b>)] from 2001 to 2022.</p>
Full article ">Figure 11
<p>Annual spatiotemporal variations of NDVI, precipitation (P), and temperature (T<sub>a</sub>) over the CONUS from 2001 to 2022. Black dots mean the regions have been checked by significant tests at a 95% confidence level.</p>
Full article ">Figure 12
<p>Evapotranspiration (ET), precipitation (P), and temperature (Ta) anomalies over the CONUS on 2012.</p>
Full article ">
19 pages, 2351 KiB  
Article
Unveiling the Centrality of Knowledge in Stakeholder Involvement Strategies Regarding Public Forest Management
by Carla Ferreira, Maria Eduarda Fernandes and Elisabete Figueiredo
Forests 2024, 15(8), 1471; https://doi.org/10.3390/f15081471 - 21 Aug 2024
Viewed by 1104
Abstract
Stakeholder involvement can foster more socially and environmentally sustainable management of natural resources, including forests. However, few studies have approached the effect of knowledge on stakeholders’ involvement in forest management. This study intends to contribute to filling this gap by exploring the relationship [...] Read more.
Stakeholder involvement can foster more socially and environmentally sustainable management of natural resources, including forests. However, few studies have approached the effect of knowledge on stakeholders’ involvement in forest management. This study intends to contribute to filling this gap by exploring the relationship between access to knowledge, involvement, stakeholders’ profiles, and levels of influence and interest regarding public forest management strategies. To this end, this article examines the data collected through a questionnaire directed to all the stakeholders potentially interested in the management of the Matas do Litoral. Matas do Litoral are part of the 3% of publicly managed forests in Portugal. The results reveal a discrepancy between the high levels of interest regarding Matas do Litoral management, and low levels of influence on those processes. Most of the stakeholders surveyed know forest management strategies, and their involvement in those strategies is limited. The proximity and role of governmental organizations are key factors underlying knowledge levels among the various stakeholders. Furthermore, knowledge acts as a critical factor in encouraging the stakeholders’ influence and involvement in management strategies and policies. This study gives insights regarding the need for knowledge management as a tool for empowering local stakeholders and promoting their involvement in bottom-up forest management strategies. Full article
(This article belongs to the Section Forest Economics, Policy, and Social Science)
Show Figures

Figure 1

Figure 1
<p>Location of the Matas do Litoral analyzed. The light green areas on the map correspond to the Matas Nacionais (National Woods) and the dark green areas to the Perímetros Florestais (Forest Perimeters). (Source: own elaboration, based on REFLOA-<a href="https://geocatalogo.icnf.pt/geovisualizador/refloa/" target="_blank">https://geocatalogo.icnf.pt/geovisualizador/refloa/</a>, accessed on 3 January 2024).</p>
Full article ">Figure 2
<p>Level of interest and influence regarding the management strategies of Matas do Litoral.</p>
Full article ">Figure 3
<p>Level of interest reported by stakeholder profile.</p>
Full article ">Figure 4
<p>Level of influence reported by stakeholders’ profiles.</p>
Full article ">Figure 5
<p>Knowledge about and the involvement of stakeholders in the Matas do Litoral management policies.</p>
Full article ">
13 pages, 4291 KiB  
Article
Effects of Soil Physical Properties on Soil Infiltration in Forest Ecosystems of Southeast China
by Di Wang, Jinhong Chen, Zhiying Tang and Yinghu Zhang
Forests 2024, 15(8), 1470; https://doi.org/10.3390/f15081470 - 21 Aug 2024
Viewed by 859
Abstract
Soil infiltration properties (SIPs) are important components of forest hydrological responses; however, few studies have investigated the mechanisms through which soil physical properties affect SIPs. In this study, two SIPs, the initial infiltration rate (IIR) and saturated hydraulic conductivity (Ks), were [...] Read more.
Soil infiltration properties (SIPs) are important components of forest hydrological responses; however, few studies have investigated the mechanisms through which soil physical properties affect SIPs. In this study, two SIPs, the initial infiltration rate (IIR) and saturated hydraulic conductivity (Ks), were quantified at five soil depths (0–10, 10–20, 20–30, 30–40, and 40–50 cm) in three forest stands (pine (Pinus taeda), oak (Quercus acutissima), and bamboo (Phyllostachys edulis) forests). We constructed a structural equation model (SEM) to analyze the main physical properties affecting the SIPs and their influence pathways, and the results show that the IIR and Ks values for the whole soil profile decreased as follows: pine forest > oak forest > bamboo forest. Soil total porosity (STP), soil field capacity (SFC), capillary water holding capacity (CMC), saturated water capacity (SWC), and initial soil water content (ISWC) were positively correlated with the SIPs, while soil bulk density (SBD) was negatively correlated with the SIPs. The SEM indicated that the main positive driver of soil infiltration was STP, while the sand content and SBD reduced soil infiltration. Soil texture indirectly affected SBD by mediating STP, and SBD indirectly affected the SIPs through SWC. These results provide data that support the simulation of subsurface hydrological responses in forests and have significant implications for forest management. Full article
(This article belongs to the Section Forest Hydrology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study area.</p>
Full article ">Figure 2
<p>Schematic diagram of the measurement principle of the device.</p>
Full article ">Figure 3
<p>Changes in initial infiltration rate and saturated hydraulic conductivity at 10 °C as a function of soil depth in three stands (pine (<span class="html-italic">Pinus taeda</span>), oak (<span class="html-italic">Quercus acutissima</span>), and bamboo (<span class="html-italic">Phyllostachys edulis</span>)): (<b>a</b>) Initial infiltration rate; (<b>b</b>) <span class="html-italic">Ks</span> (Saturated hydraulic conductivity at 10 °C). Capital letters A and B indicate significant differences in infiltration at different soil depths in the same stand; lowercase letters a and b denote significant differences in infiltration at the same soil depth across stands.</p>
Full article ">Figure 4
<p>Correlations between soil physical properties and initial infiltration rate, and saturated hydraulic conductivity at 10 °C in three stands (<b>a</b>), and redundancy analysis of soil physical properties and water infiltration properties (<b>b</b>), IIR: Initial infiltration rate; <span class="html-italic">Ks</span>: Saturated hydraulic conductivity at 10 °C.</p>
Full article ">Figure 5
<p>Causal relationships between dominant factors of soil physical properties and soil moisture infiltration properties in three stands were based on structural equation modeling (SEM) and standardized direct and total effects. (*** indicates significance at <span class="html-italic">p</span> &lt; 0.001. SBD, soil bulk density; STP, total capillary porosity; SWC, saturated water capacity. (<b>a</b>) IIR, initial infiltration rate; (<b>b</b>) <span class="html-italic">Ks</span>, saturated hydraulic conductivity at 10 °C in the three stands. The value next to each arrow represents the normalized path coefficient. Unidirectional arrows indicate the direct effect of a unidirectional causal relationship. The width of the arrow indicates the strength of the causal relationship.</p>
Full article ">
13 pages, 5155 KiB  
Article
Developing Forest Road Recognition Technology Using Deep Learning-Based Image Processing
by Hyeon-Seung Lee, Gyun-Hyung Kim, Hong Sik Ju, Ho-Seong Mun, Jae-Heun Oh and Beom-Soo Shin
Forests 2024, 15(8), 1469; https://doi.org/10.3390/f15081469 - 21 Aug 2024
Viewed by 800
Abstract
This study develops forest road recognition technology using deep learning-based image processing to support the advancement of autonomous driving technology for forestry machinery. Images were collected while driving a tracked forwarder along approximately 1.2 km of forest roads. A total of 633 images [...] Read more.
This study develops forest road recognition technology using deep learning-based image processing to support the advancement of autonomous driving technology for forestry machinery. Images were collected while driving a tracked forwarder along approximately 1.2 km of forest roads. A total of 633 images were acquired, with 533 used for the training and validation sets, and the remaining 100 for the test set. The YOLOv8 segmentation technique was employed as the deep learning model, leveraging transfer learning to reduce training time and improve model performance. The evaluation demonstrates strong model performance with a precision of 0.966, a recall of 0.917, an F1 score of 0.941, and a mean average precision (mAP) of 0.963. Additionally, an image-based algorithm is developed to extract the center from the forest road areas detected by YOLOv8 segmentation. This algorithm detects the coordinates of the road edges through RGB filtering, grayscale conversion, binarization, and histogram analysis, subsequently calculating the center of the road from these coordinates. This study demonstrates the feasibility of autonomous forestry machines and emphasizes the critical need to develop forest road recognition technology that functions in diverse environments. The results can serve as important foundational data for the future development of image processing-based autonomous forestry machines. Full article
Show Figures

Figure 1

Figure 1
<p>Location of study site and tested forest road route.</p>
Full article ">Figure 2
<p>Vision cameras for image data collection.</p>
Full article ">Figure 3
<p>Visualization of mask image.</p>
Full article ">Figure 4
<p>YOLOv8 architecture [<a href="#B29-forests-15-01469" class="html-bibr">29</a>].</p>
Full article ">Figure 5
<p>Flow chart of image-based forested center-extraction algorithm.</p>
Full article ">Figure 6
<p>Binary image histogram.</p>
Full article ">Figure 7
<p>(<b>a</b>) YOLOv8: bounding box, segmentation, classification, and distribution focal loss for training dataset. (<b>b</b>) YOLOv8: bounding box, segmentation, classification, and distribution focal loss (dfl_loss) for validation dataset.</p>
Full article ">Figure 8
<p>Forest road detection and segmentation using YOLOv8 in test image.</p>
Full article ">Figure 9
<p>Image-based forest center extraction results.</p>
Full article ">Figure 9 Cont.
<p>Image-based forest center extraction results.</p>
Full article ">Figure 10
<p>Recognition results by algorithm on actual forest roads.</p>
Full article ">
13 pages, 2566 KiB  
Article
Changes in Soil Total and Microbial Biomass Nitrogen in Deforested and Eroded Areas in the Western Black Sea Region of Turkey
by İlyas Bolat and Huseyin Sensoy
Forests 2024, 15(8), 1468; https://doi.org/10.3390/f15081468 - 21 Aug 2024
Viewed by 605
Abstract
The microbial biomass in soil is an active and living constituent of organic matter. It is both a storage pool and a source of plant nutrients that can be used as required. In addition, each microbial indicator evaluates soil quality and health from [...] Read more.
The microbial biomass in soil is an active and living constituent of organic matter. It is both a storage pool and a source of plant nutrients that can be used as required. In addition, each microbial indicator evaluates soil quality and health from different perspectives, which are not necessarily very different. This study was conducted to compare some physical, chemical, and biochemical characteristics of the soils of forest (SF) and deforested (SDE) areas located on the slopes of the Kirazlıköprü area, which was previously deforested due to dam construction in Bartın province in northwestern Turkey. Soil samples were taken from the topsoil surface (0–5 cm) to determine the microbial soil characteristics of the SF and SDE sites. The soil microbial biomass N (Nmic) was determined by chloroform fumigation extraction, and the Cmic/Nmic ratio and Nmic/Ntotal percentage were calculated using the original values. Total N, Nmic and Cmic/Nmic values are higher in the forest area. The lowest and highest total N (Ntotal) contents in the SF and SDE soils varied between 1.50 and 3.47 g kg−1 and 0.91 and 1.46 g kg−1, respectively. Similarly, the Nmic contents of the SF and SDE soils varied between 75.56 and 143.42 μg g−1 and 10.40 and 75.96 μg g−1, respectively. A statistical analysis revealed that the mean Ntotal and mean Nmic values differed (p < 0.05) in the SF and SDE soils. The mean Cmic/Nmic values in the SF and SDE soils were 8.79 (±1.65) and 5.64 (±1.09), respectively, and a statistical difference was found between the fields (p < 0.05). Our findings indicate that the soil microbial community structure varies according to the site. As a result, it can be concluded that deforestation and erosion due to dam construction in the area led to the removal of plant nutrients from the soil; deterioration in the amount and activity of microbial biomass; and, consequently, soil losses and degradation of soil quality. Full article
(This article belongs to the Section Forest Soil)
Show Figures

Figure 1

Figure 1
<p>Kirazlıköprü Dam and sampling points in forest and deforested sites.</p>
Full article ">Figure 2
<p>Images of grooves (<b>A</b>) and gullies (<b>B</b>) in the deforested site.</p>
Full article ">Figure 3
<p>Mean total N (<b>A</b>) and mean microbial biomass N (<b>B</b>) values in S<sub>F</sub> and S<sub>DE</sub> soils. The letters in parentheses represent a difference (<span class="html-italic">p</span> &lt; 0.05) between fields.</p>
Full article ">Figure 4
<p>Mean N<sub>mic</sub>/N<sub>total</sub> (%) values in S<sub>F</sub> and S<sub>DE</sub> soils. The letters in parentheses represent no difference (<span class="html-italic">p</span> &gt; 0.05) between fields.</p>
Full article ">Figure 5
<p>Mean C<sub>mic</sub>/N<sub>mic</sub> values in S<sub>F</sub> and S<sub>DE</sub> soils. The letters in parentheses represent the difference (<span class="html-italic">p</span> &lt; 0.05) between fields.</p>
Full article ">
18 pages, 6073 KiB  
Article
Estimation of NPP in Huangshan District Based on Deep Learning and CASA Model
by Ziyu Wang, Youfeng Zhou, Xinyu Sun and Yannan Xu
Forests 2024, 15(8), 1467; https://doi.org/10.3390/f15081467 - 21 Aug 2024
Viewed by 822
Abstract
Net primary productivity (NPP) is a key indicator of the health of forest ecosystems that offers important information about the net carbon sequestration capacity of these systems. Precise assessment of NPP is crucial for measuring carbon fixation and assessing the general well-being of [...] Read more.
Net primary productivity (NPP) is a key indicator of the health of forest ecosystems that offers important information about the net carbon sequestration capacity of these systems. Precise assessment of NPP is crucial for measuring carbon fixation and assessing the general well-being of forest ecosystems. Due to the distinct ecological characteristics of various forest types, accurately understanding and delineating the distribution of these types is crucial for studying NPP. Therefore, an accurate forest-type classification is necessary prior to NPP calculation to ensure the accuracy and reliability of the research findings. This study introduced deep learning technology and constructed an HRNet-CASA framework that integrates the HRNet deep learning model and the CASA model to achieve accurate estimation of forest NPP in Huangshan District, Huangshan City, Anhui Province. Firstly, based on VHR remote sensing images, we utilized the HRNet to classify the study area into six forest types and obtained the forest type distribution map of the study area. Then, combined with climate data and forest type distribution data, the CASA model was used to estimate the NPP of forest types in the study area, and the comparison with the field data proved that the HRNet-CASA framework simulated the NPP of the study area well. The experimental findings show that the HRNet-CASA framework offers a novel approach to precise forest NPP estimation. Introducing deep learning technology not only enables precise classification of forest types but also allows for accurate estimation of NPP for different types of forests. This provides a more effective tool for forest ecological research and environmental protection. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The study area; (<b>a</b>) location of Huangshan District; (<b>b</b>) location of sample plots.</p>
Full article ">Figure 2
<p>HRNet-V2 structure diagram.</p>
Full article ">Figure 3
<p>Ortho images of different classification types in VHR images.</p>
Full article ">Figure 4
<p>Technical flow chart.</p>
Full article ">Figure 5
<p>Confusion matrix heat map of HRNet classification.</p>
Full article ">Figure 6
<p>Forest-type distribution map.</p>
Full article ">Figure 7
<p>NPP distribution map.</p>
Full article ">Figure 8
<p>Comparison of NPP simulation and field data.</p>
Full article ">
12 pages, 5931 KiB  
Article
Soil-Moisture-Dependent Temperature Sensitivity of Soil Respiration in a Poplar Plantation in Northern China
by Huan He, Tonggang Zha and Jiongrui Tan
Forests 2024, 15(8), 1466; https://doi.org/10.3390/f15081466 - 21 Aug 2024
Viewed by 796
Abstract
The temperature sensitivity (Q10) of soil respiration (Rs) plays a crucial role in evaluating the carbon budget of terrestrial ecosystems under global warming. However, the variability in Q10 along soil moisture gradients remains a subject of debate, and the associated [...] Read more.
The temperature sensitivity (Q10) of soil respiration (Rs) plays a crucial role in evaluating the carbon budget of terrestrial ecosystems under global warming. However, the variability in Q10 along soil moisture gradients remains a subject of debate, and the associated underlying causes are poorly understood. This study aims to investigate the characteristics of Q10 changes along soil moisture gradients throughout the whole growing season and to assess the factors influencing Q10 variability. Changes in soil respiration (measured by the dynamic chamber method) and soil properties were analyzed in a poplar plantation located in the suburban area of Beijing, China. The results were as follows: (1) Q10 increased with the increasing soil water content up to a certain threshold, and then decreased, (2) the threshold was 75% to 80% of the field capacity (i.e., the moisture content at capillary rupture) rather than the field water-holding capacity, and (3) the dominant influence shifted from soil solid-phase properties to microbes with increasing soil moisture. Our results are important for understanding the relationship between the temperature sensitivity of soil respiration and soil moisture in sandy soil, and for the refinement of the modeling of carbon cycling in terrestrial ecosystems. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area.</p>
Full article ">Figure 2
<p>A conceptual diagram showing the factors influencing Q<sub>10</sub> in sandy soil of a poplar plantation in northern China.</p>
Full article ">Figure 3
<p>Soil water retention curve for the topsoil layer.</p>
Full article ">Figure 4
<p>Diurnal variation (<b>a</b>) and temporal variation (mean values ± SD) (<b>b</b>) of soil respiration.</p>
Full article ">Figure 5
<p>Relationships between daily Q<sub>10</sub> values and soil moisture during the growing season. R<sup>2</sup> is the proportion of variance explained. <span class="html-italic">P</span> is the significance level. The black arrows indicate the observed trend in the relationship between daily Q<sub>10</sub> values and soil water content.</p>
Full article ">Figure 6
<p>Correlation analysis of Q<sub>10</sub> and soil properties for soil moisture lower than the WRC (<b>a</b>) and soil moisture higher than the WRC (<b>b</b>). The blue circles indicate positive correlations, and the red circles indicate negative correlations. The coefficients in the figure represent a strong (* <span class="html-italic">p</span> &lt; 0.05), significant (** <span class="html-italic">p</span> &lt; 0.01), or strongly significant (*** <span class="html-italic">p</span> &lt; 0.001) correlation.</p>
Full article ">Figure 7
<p>Individual and shared effects of soil moisture, soil solid-phase properties, and microbial factors on Q<sub>10</sub>, as derived when the soil moisture was lower than the WRC (<b>a</b>) and when the soil moisture was higher than the WRC (<b>b</b>). Dynamics of the relationships among Q<sub>10</sub> and soil properties during the growing season (<b>c</b>).</p>
Full article ">
15 pages, 4201 KiB  
Article
Species and Stand Management Options for Wood Production from Small Grower Plantations in Central Vietnam
by Christopher E. Harwood, Le Xuan Toan, Pham Xuan Dinh and E. K. Sadanandan Nambiar
Forests 2024, 15(8), 1465; https://doi.org/10.3390/f15081465 - 21 Aug 2024
Viewed by 755
Abstract
Acacia hybrid (Acacia mangium Willd. × A. auriculiformis A. Cunn. Ex Benth.) dominates plantation wood production in central Vietnam. Dependence on a single species may increase biological risks. The potential of eucalypt as an alternative was examined by comparing the growth and [...] Read more.
Acacia hybrid (Acacia mangium Willd. × A. auriculiformis A. Cunn. Ex Benth.) dominates plantation wood production in central Vietnam. Dependence on a single species may increase biological risks. The potential of eucalypt as an alternative was examined by comparing the growth and survival of acacia hybrid and eucalypt hybrid (Eucalyptus urophylla S.T. Blake × E. pellita F. Muell.) clones in Quang Tri province at three planting densities (1333, 1667 and 2222 trees ha−1). The experiment was planted on an eroded shallow soil common in the region. At age 5 years, survival of acacia (74%) was higher than that of eucalypt (67%), a consequence of high mortality from wind damage for one eucalypt clone. Eucalypt was taller by about 2 m, but stem diameters of acacia and eucalypt were very similar. For both taxa, diameter decreased significantly as planting density increased. Across planting densities, mean standing volume was 107 and 108 m3 ha−1 for acacia and eucalyptus, respectively. Linear regressions of stocking at 5 years on volume accounted for over half of the variance in acacia and eucalypt plot volumes, demonstrating the strong effect of stocking on yield. There were similarly strong effects of stocking on stem diameter. Acacia hybrid plantations of nearby small growers had stockings at age 5 years that averaged over 2500 stems ha−1. Growers planted at higher densities and allowed their trees to multi-stem. Their standing volumes at age 5 years ranged from 83 to 102 m3 ha−1. Understanding how to reduce tree mortality would assist growers to choose planting densities and stand management that optimise growth, log diameter classes and net returns. Full article
(This article belongs to the Section Genetics and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) soil profile, the colour bands show 10 cm intervals, (<b>b</b>) typical understorey below acacia (left of observer) and eucalypt (right), (<b>c</b>), crowns of eucalypt trees and (<b>d</b>) crowns of acacia trees. Photos taken at a stand age of 5.5 years.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) soil profile, the colour bands show 10 cm intervals, (<b>b</b>) typical understorey below acacia (left of observer) and eucalypt (right), (<b>c</b>), crowns of eucalypt trees and (<b>d</b>) crowns of acacia trees. Photos taken at a stand age of 5.5 years.</p>
Full article ">Figure 2
<p>Relationships between stocking and growth attributes: (<b>a</b>) regression of standing volume on stocking; (<b>b</b>) regression of stem diameter on stocking.</p>
Full article ">Figure 2 Cont.
<p>Relationships between stocking and growth attributes: (<b>a</b>) regression of standing volume on stocking; (<b>b</b>) regression of stem diameter on stocking.</p>
Full article ">Figure 3
<p>Mortality trends of clones from age 1 to 5 years.</p>
Full article ">Figure 4
<p>Typical small grower plantation, showing multi-stemming of many trees.</p>
Full article ">
11 pages, 1823 KiB  
Article
Determinants of Deadwood Biomass under the Background of Nitrogen and Water Addition in Warm Temperate Forests
by Liang Hong, Shouzheng Tang, Tao Li, Liyong Fu, Xinyu Song, Guangshuang Duan, Jueming Fu and Lei Ma
Forests 2024, 15(8), 1464; https://doi.org/10.3390/f15081464 - 20 Aug 2024
Viewed by 647
Abstract
Climate change is exacerbating the vulnerability of temperate forests to severe disturbances, potentially increasing tree mortality rates. Despite the significance of this issue, there has been a lack of comprehensive research on tree survival across extensive forest areas under the background of global [...] Read more.
Climate change is exacerbating the vulnerability of temperate forests to severe disturbances, potentially increasing tree mortality rates. Despite the significance of this issue, there has been a lack of comprehensive research on tree survival across extensive forest areas under the background of global climate change. To fill this gap, we conducted a detailed analysis of tree survival within a canopy nitrogen and water addition experimental platform in central China, utilizing data from two censuses and evaluating contributing factors. Our findings revealed 283 dead trees within the plots, predominantly of very small diameters (1–10 cm). The distribution of these dead trees varied among subplots, influenced by both biotic and abiotic factors. Notably, three dominant tree species were responsible for 64.8% of the deadwood biomass. The study determined that both the breast diameter and the quantity of dead trees, affected by surrounding trees and environmental conditions, played a critical role in deadwood biomass accumulation. This research offers an in-depth examination of deadwood biomass patterns in a temperate forest, highlighting the need to consider both experiment treatments and abiotic elements like topography in studies of forest ecosystem carbon. The insights gained from this study enhance our understanding of warm temperate forests’ role in the global carbon cycle and offer valuable guidance for forest conservation and management strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study site and experiment design.</p>
Full article ">Figure 2
<p>Facilities for nitrogen and water application in forest canopies.</p>
Full article ">Figure 3
<p>Distribution of breast diameter of dead trees in the study site.</p>
Full article ">Figure 4
<p>Path analysis of the effects of nitrogen and water addition on the deadwood biomass.</p>
Full article ">
23 pages, 11067 KiB  
Article
A Down-Scaling Inversion Strategy for Retrieving Canopy Water Content from Satellite Hyperspectral Imagery
by Meihong Fang, Xiangyan Hu, Jing M. Chen, Xueshiyi Zhao, Xuguang Tang, Haijian Liu, Mingzhu Xu and Weimin Ju
Forests 2024, 15(8), 1463; https://doi.org/10.3390/f15081463 - 20 Aug 2024
Viewed by 707
Abstract
Vegetation canopy water content (CWC) crucially affects stomatal conductance and photosynthesis and, consequently, is a key state variable in advanced ecosystem models. Remote sensing has been shown to be an effective tool for retrieving CWCs. However, the retrieval of the CWC from satellite [...] Read more.
Vegetation canopy water content (CWC) crucially affects stomatal conductance and photosynthesis and, consequently, is a key state variable in advanced ecosystem models. Remote sensing has been shown to be an effective tool for retrieving CWCs. However, the retrieval of the CWC from satellite remote sensing data is affected by the vegetation canopy structure and soil background. This study proposes a methodology that combines a modified spectral down-scaling model with a high-universality leaf water content inversion model to retrieve the CWC through constraining the impacts of canopy structure and soil background on CWC retrieval. First, canopy spectra acquired by satellite sensors were down-scaled to leaf reflectance spectra according to the probabilities of viewing the sunlit foliage (PT) and background (PG) and the estimated spectral multiple scattering factor (M). Then, leaf water content, or equivalent water thickness (EWT), was obtained from the down-scaled leaf reflectance spectra via a leaf-scale EWT inversion model calibrated with PROSPECT simulation data. Finally, the CWC was calculated as the product of the estimated leaf EWT and canopy leaf area index. Validation of this coupled model was performed using satellite-ground synchronous observation data across various vegetation types within the study area, affirming the model’s broad applicability. Results indicate that the modified spectral down-scaling model accurately retrieves leaf reflectance spectra, aligning closely with site-level measured spectra. Compared to the direct inversion approach, which performs poorly with Hyperion satellite images, the down-scale strategy notably excels. Specifically, the Similarity Water Index (SWI)-based canopy EWT coupled model achieved the most precise estimation, with a normalized Root Mean Square Error (nRMSE) of 15.28% and an adjusted R2 of 0.77, surpassing the performance of the best index Shortwave Angle Normalized Index (SANI)-based model (nRMSE = 15.61%, adjusted R2 = 0.52). Given its calibration using simulated data, this coupled model proved to be a potent method for extracting canopy EWT from satellite imagery, suggesting its applicability to retrieve other vegetative biochemical components from satellite data. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Study area and sampling locations: (<b>a</b>) the study area was located in Menghai county (21.98° N, 100.29° E), Xishuangbanna, Yunnan province in southwest China. (<b>b</b>) The RGB true color image of Hyperion remote sensing data and (<b>c</b>) sampling points for ground synchronous observation.</p>
Full article ">Figure 2
<p>Workflow of the coupled down-scaling inversion strategy retrieving leaf-level and canopy-level water content from satellite data considering canopy structure and background effects.</p>
Full article ">Figure 3
<p>Sensitivity of the simulated probability of viewing sunlit foliage (PT) (<b>a</b>) and sunlit background (PG) (<b>b</b>) to the solar zenith Angle (SZA), LAI and radius of tree crowns. Here, the tree density was set at 4000 trees per hectare and VZA = 0°.</p>
Full article ">Figure 4
<p>Correlations of NSAI with PT (<b>a</b>) and PG (<b>b</b>) for the Hyperion synthetic data under conditions of normal soil (N, orange dots), dry soil (D, green dots), and wet soil (W, blue dots). A total of 13,950,144 Hyperion samples and 10,764 Hyperion scenes are available for analysis. On 33 Hyperion pixels, we compare PT (<b>c</b>) and PG (<b>d</b>) estimates using NSAI-based models to the reference values inverted with the 4-Scale GO model. The red straight lines are the 1:1 line. Correlations of NSAI with PT (<b>a</b>) and comparisons of PT estimated using NSAI with the reference values (<b>c</b>) are adapted from Fang et al. [<a href="#B25-forests-15-01463" class="html-bibr">25</a>].</p>
Full article ">Figure 5
<p>Spatial patterns of LAI (<b>a</b>) derived using MSR<sub>705</sub> and spatial distribution of PT (<b>b</b>) and PG (<b>c</b>) estimated using NASI retrieved from the Hyperion image over the study area at a spatial resolution of 30 m.</p>
Full article ">Figure 6
<p>Correlations between leaf EWT and SWI for the measured data (<b>a</b>) and the simulated data using the PROSPECT model (<b>b</b>). Validation of leaf EWT retrieved using the SWI-based model derived from measured data (<b>c</b>) and the simulated data (<b>d</b>). All leaf reflectance spectra were resampled to Hyperion-equivalent spectra.</p>
Full article ">Figure 7
<p>Spatial distribution of average leaf EWT inverted from coupled down-scaling inversion strategy (<b>a</b>) and canopy water content per unit ground surface area derived based on the LAI image and retrieved average leaf EWT (<b>b</b>). The spatial resolution of the image is 30 × 30 m.</p>
Full article ">Figure 8
<p>Comparison of measured CWC against those retrieved from SWI-based (<b>a</b>) and SANI-based (<b>b</b>) coupled models using the down-scaling inversion strategy.</p>
Full article ">Figure A1
<p>After preprocessing of the Hyperion image, spectra were compared between (<b>a</b>) vegetation with different crown closures (high, moderate, and low coverage) and (<b>b</b>) soil background with red and gray hues. Data were adapted from Fang et al. [<a href="#B25-forests-15-01463" class="html-bibr">25</a>].</p>
Full article ">Figure A2
<p>Correlations of the viewing probabilities of sunlit crown and background (PT and PG) based on the 4-Scale GO model simulations from the Hyperion synthetic data, which includes 10,764 scenes.</p>
Full article ">
18 pages, 14491 KiB  
Article
Influence of Main Flight Parameters on the Performance of Stand-Level Growing Stock Volume Inventories Using Budget Unmanned Aerial Vehicles
by Marek Lisańczuk, Grzegorz Krok, Krzysztof Mitelsztedt and Justyna Bohonos
Forests 2024, 15(8), 1462; https://doi.org/10.3390/f15081462 - 20 Aug 2024
Viewed by 848
Abstract
Low-altitude aerial photogrammetry can be an alternative source of forest inventory data and a practical tool for rapid forest attribute updates. The availability of low-cost unmanned aerial systems (UASs) and continuous technological advances in terms of their flight duration and automation capabilities makes [...] Read more.
Low-altitude aerial photogrammetry can be an alternative source of forest inventory data and a practical tool for rapid forest attribute updates. The availability of low-cost unmanned aerial systems (UASs) and continuous technological advances in terms of their flight duration and automation capabilities makes these solutions interesting tools for supporting various forest management needs. However, any practical application requires a priori empirical validation and optimization steps, especially if it is to be used under different forest conditions. This study investigates the influence of the main flight parameters, i.e., ground sampling distance and photo overlap, on the performance of individual tree detection (ITD) stand-level forest inventories, based on photogrammetric data obtained from budget unmanned aerial systems. The investigated sites represented the most common forest conditions in the Polish lowlands. The results showed no direct influence of the investigated factors on growing stock volume predictions within the analyzed range, i.e., overlap from 80 × 80 to 90 × 90% and GSD from 2 to 6 cm. However, we found that the tree detection ratio had an influence on estimation errors, which ranged from 0.6 to 15.3%. The estimates were generally coherent across repeated flights and were not susceptible to the weather conditions encountered. The study demonstrates the suitability of the ITD method for small-area forest inventories using photogrammetric UAV data, as well as its potential optimization for larger-scale surveys. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of applied methodology.</p>
Full article ">Figure 2
<p>Investigation area: range of Polish lowlands (<b>a</b>), Skierniewice forest district (<b>b</b>), selected sites (<b>c</b>). Outline of the district in orange. A, B, C, D—selected stands.</p>
Full article ">Figure 3
<p>The most common forest site types in Polish lowlands. Abbreviations: C—coniferous, M—mixed, F—fresh, B—broadleaved.</p>
Full article ">Figure 4
<p>Fragments of orthomosaics presenting areas of interest within analyzed stands. (<b>A</b>–<b>D</b>)—selected stands.</p>
Full article ">Figure 5
<p>Visual results of sample segmentation over test stand A.</p>
Full article ">Figure 6
<p>3D model of validation stand A based on TLS data, with detected stems.</p>
Full article ">Figure 7
<p>Scatterplots of the winning model, (<b>a</b>)—fitted vs observed, (<b>b</b>)—fitted vs residuals (m<sup>3</sup>).</p>
Full article ">Figure 8
<p>Histogram of residual frequency.</p>
Full article ">Figure 9
<p>GSV estimates according to GSD. Missing values were interpolated.</p>
Full article ">Figure 10
<p>Relationship between tree detection ratio and GSV estimation error.</p>
Full article ">
26 pages, 10877 KiB  
Article
Phytosociological Analysis of the Boreal Oligotrophic Pine Forests in the Southern Ural Region (Russia)
by Vasiliy Martynenko, Pavel Shirokikh, Elvira Baisheva, Albert Muldashev, Nikolay Fedorov, Svetlana Zhigunova and Leniza Naumova
Forests 2024, 15(8), 1461; https://doi.org/10.3390/f15081461 - 20 Aug 2024
Viewed by 774
Abstract
Floristic composition and syntaxonomy of the boreal oligotrophic pine forests covering vast areas in the boreal, hemiboreal, and temperate zones of continental Eurasia still remain poorly studied in the Southern Ural region. Using the Braun–Blanquet approach and TURBOVEG and JUICE software, the phytocoenotic [...] Read more.
Floristic composition and syntaxonomy of the boreal oligotrophic pine forests covering vast areas in the boreal, hemiboreal, and temperate zones of continental Eurasia still remain poorly studied in the Southern Ural region. Using the Braun–Blanquet approach and TURBOVEG and JUICE software, the phytocoenotic diversity of boreal oligotrophic pine forests of the Southern Ural region was studied, and their position in the system of ecological and floristic classification of Eurasian vegetation was determined. Geobotanical data on boreal oligotrophic pine forests of Europe, including the European part of Russia; the Southern Urals; and Siberia were compared. A new alliance of oligotrophic boreal pine forests Brachypodio pinnati-Pinion sylvestris all. nov. hoc loco is described. The communities of this new alliance (i.e., five associations from the Southern Ural region) are characterized by a special floristic composition, occupying an intermediate position between the typical European oligotrophic pine forests of the alliance Dicrano-Pinion (Libbert 1933) Matuszkiewicz 1962 and oligotrophic (mainly psammophilous) South Siberian pine forests of the alliance Hieracio-Pinion Anenkhonov et Chytrý 1998. The communities of the alliance Brachypodio pinnati-Pinion sylvestris prefer to grow on poor soils with different moisture conditions. Due to intensive forestry activities, the distribution area of these forests has decreased, and these communities have been replaced by secondary birch forests. We have proposed a set of conservation measures to preserve these communities. A new association of oligotrophic pine forests Psephello sumensis-Pinetum sylvestris ass. nov. hoc loco is also described. These communities from the Kurgan region of Western Siberia were ordered into the alliance Dicrano-Pinion. It confirms the idea that the distribution area of this alliance may reach Siberia. Unlike the Southern Ural pine forests of the alliance Brachypodio pinnati-Pinion sylvestris, the recovery of these West Siberian pine forests after felling is quite high, and these communities do not require special measures for their protection. Full article
(This article belongs to the Special Issue Forest Biodiversity Conservation)
Show Figures

Figure 1

Figure 1
<p>Localities of relevés of boreal oligotrophic pine forests in the SUR and the Kurgan region.</p>
Full article ">Figure 2
<p>Distribution of syntaxa of boreal oligotrophic pine forests in the SUR and the Kurgan region.</p>
Full article ">Figure 3
<p>Tree diagram (Ward’s method, Bray-Curtis distances) of 61 syntaxa of boreal oligotrophic pine forests from alliances Cladonio stellaris-<span class="html-italic">Pinion</span> sylvestris (A: 1–16), Dicrano-<span class="html-italic">Pinion</span> (B: 17–42), Brachypodio-<span class="html-italic">Pinion</span> (C: 43–51, red color), Hieracio-<span class="html-italic">Pinion</span> (D: 52–61). Numbers in the tree diagram correspond to the numbers of the syntaxa in <a href="#app1-forests-15-01461" class="html-app">Table S1</a>.</p>
Full article ">Figure 4
<p>Boreal oligotrophic pine forests in the SUR (authors of photos Shirokikh P.S., Muldashev A.A. ((<b>a</b>) Ass. <span class="html-italic">Zigadeno sibirici</span>-<span class="html-italic">Pinetum sylvestris</span>; (<b>b</b>) Ass. <span class="html-italic">Pleurospermo uralensis</span>-<span class="html-italic">Pinetum sylvestris</span>, subass. <span class="html-italic">digitalietosum grandiflorae</span>; (<b>c</b>) Ass. <span class="html-italic">Pleurospermo uralensis</span>-<span class="html-italic">Pinetum sylvestris</span>, subass. <span class="html-italic">anemonastretosum biarmiensis</span>).</p>
Full article ">
13 pages, 9031 KiB  
Article
Impacts of Downed Dead Wood Poplar Trees on Forest Regeneration in the Semi-Arid Region of Northern China
by Pengwu Zhao, Lijuan Guan, Huaxia Yao, Yang Shu, Yongjie Yue, Furen Liu, Yaxiong Zheng, Longfei Hao, Changlin Xiang and Liwen Zhou
Forests 2024, 15(8), 1460; https://doi.org/10.3390/f15081460 - 19 Aug 2024
Viewed by 748
Abstract
In the past few decades, due to rising temperatures and changes in precipitation, the degree of drought in semi-arid areas has increased, leading to a large number of tree deaths and threatening the natural forests distributed in the semi-arid areas of North China. [...] Read more.
In the past few decades, due to rising temperatures and changes in precipitation, the degree of drought in semi-arid areas has increased, leading to a large number of tree deaths and threatening the natural forests distributed in the semi-arid areas of North China. This article takes the forest ecosystem of Saihanwula Nature Reserve in the southern section of Greater Khingan Mountains in China’s semi-arid region as a research area and studies the distribution of downed dead wood and its impact on forest renewal in the area. We used the sample plot survey method, investigated the number of downed dead wood, decay class, dumping direction, existence form, and the number of regenerated seedlings in the sample plot, and calculated the density of regenerated seedlings in different plots. The renewal density is 4050 ± 824, 2950 ± 265, plants/ha, and 2625 ± 237 plants/ha, respectively, in the sample plots for Later-death plot, Mid-death plot, and Early-death plot. The average storage of downed dead wood in Saihanwula Nature Reserve is 58.51 ± 16.56 m3/ha. The distribution densities of downed dead wood are 50 ± 21, 806 ± 198, 189 ± 76, and 22 ± 5 plants/ha for decay classes II, III, IV, and V respectively. The main form of downed dead wood in the research area is “trunk base fracture”, accounting for 68.78% of the total number of downed dead wood. A large number of downed dead wood had serious negative effects, such as crushing and injuring the regeneration seedlings and other plants under the forest at the moment of dumping and for a long time after dumping. The crushed and injured rate is 5.3~7.8%, with downed dead wood accumulated in the forest from the early stage of downed dead wood. It had negative effects on the regeneration of seeds, seedlings, and young trees, such as obstructing and hiding the light from the soil surface and inhibiting the regeneration and growth of seedlings. However, after the trees were dumped, large gaps appeared in the forest, increasing the sunlight area on the soil surface. In the later stage of tree death, moderately high decayed downed dead wood changed the soil structure in terms of soil softness, water holding capacity, and nutrient content, thus promoting the growth of seedlings and young trees. Reasonably utilizing the relationship between downed dead wood and forest renewal can effectively promote the healthy development of forests. Full article
(This article belongs to the Special Issue Ecosystem Degradation and Restoration: From Assessment to Practice)
Show Figures

Figure 1

Figure 1
<p>Location and site characteristics of the study area. <b>Upper left</b>: The location of the study area (<b>green triangle</b>) in northern China, with Inner Mongolia shown in light blue color. <b>Center left</b>: Elevation map of the Saihanwula Nature Reserve. <b>Right and bottom row:</b> Four pictures showing the landscape and downed dead wood.</p>
Full article ">Figure 2
<p>The generation process of downed dead wood, regeneration approach of seedlings, and decay class of downed dead wood.</p>
Full article ">Figure 3
<p>Downed dead wood reserves in Saihanwula Nature Reserve forest.</p>
Full article ">Figure 4
<p>The distribution of downed dead wood in plots.</p>
Full article ">Figure 5
<p>Decay class of downed dead wood in Saihanwula Nature Reserve forest.</p>
Full article ">Figure 6
<p>Precipitation and temperature changes from 1961 to 2022.</p>
Full article ">Figure 7
<p>The regeneration density of Saihanwula Nature Reserve forest. Note: “NS” is no significance (<span class="html-italic">p</span> &gt; 0.05) for the difference between two plots.</p>
Full article ">Figure 8
<p>The crushed and injured of downed dead wood in Saihanwula Nature Reserve forest.</p>
Full article ">
23 pages, 10102 KiB  
Article
Heat Mitigation Benefits of Street Tree Species during Transition Seasons in Hot and Humid Areas: A Case Study in Guangzhou
by Senlin Zheng, Caiwei He, Haodong Xu, Jean-Michel Guldmann and Xiao Liu
Forests 2024, 15(8), 1459; https://doi.org/10.3390/f15081459 - 19 Aug 2024
Viewed by 772
Abstract
The potential microclimatic effects of street trees are influenced by their ecological characteristics, planting patterns, and street orientations, especially in subtropical hot and humid areas. To investigate these effects, four typical street tree species in Guangzhou were selected for study during the transition [...] Read more.
The potential microclimatic effects of street trees are influenced by their ecological characteristics, planting patterns, and street orientations, especially in subtropical hot and humid areas. To investigate these effects, four typical street tree species in Guangzhou were selected for study during the transition seasons: Khaya senegalensis, Terminalia neotaliala, Ficus microcarpa, and Mangifera indica. Air temperature (AT), relative humidity (RH), solar radiation (SR), surface temperature (ST), wind speed (WS), and the leaf area index (LAI) were monitored. The cooling effects of these four species and the resulting improvements in human thermal comfort (HTC) were assessed. The influences of tree planting patterns and street orientations on cooling benefits were systematically analyzed. The results indicate that, during transition seasons, the four street trees, on average, can block 96.68% of SR, reduce AT by 1.45 °C and ST by 10.25 °C, increase RH by 5.26%, and lower the physiologically equivalent temperature (PET) by 8.34 °C. Terminalia neotaliala, reducing AT and PET by 1.76 °C and 12.4 °C, respectively, offers the greatest potential for microclimate improvement. Among the four tree species, the variations in ST (ΔST) and PET (ΔPET) were minimal, at only 0.76 °C and 0.25 °C, respectively. The average differences in AT and PET between inter-tree and under-tree environments were 0.06 °C and 0.98 °C, respectively. The AT reduction rate was 1.7 times higher in the double-row planting pattern compared to the single-row planting pattern. Street trees planted in the northwest–southeast (NW-SE) orientation exhibited a 16.96% lower WS reduction than those in other orientations. The northeast–southwest (NE-SW) orientation showed the least potential to enhance human thermal comfort. Compared to NE-SW, the northwest–southeast (NW-SE) orientation achieved twice the rate of AT reduction, while the north–south (N-S) orientation improved it by 1.3 times. This data analysis aids in assessing the impact of green infrastructure on urban climates and demonstrates the year-round microclimatic benefits of street trees. Full article
Show Figures

Figure 1

Figure 1
<p>Methodological flowchart.</p>
Full article ">Figure 2
<p>(<b>a</b>) Meteorological data for a typical year; (<b>b</b>) wind rose diagram in Guangzhou (China Meteorological Data Service Center, 2020).</p>
Full article ">Figure 3
<p>Tested street tree species.</p>
Full article ">Figure 4
<p>Location and orientation of the experimental test sites.</p>
Full article ">Figure 5
<p>Schematic diagram of measuring points.</p>
Full article ">Figure 6
<p>Instrument arrangement for experimental testing: (<b>a</b>) Louvered box radiation shield; (<b>b</b>) HOBO thermocouple.</p>
Full article ">Figure 7
<p>Daily variations in AT between trees, under trees, and in the open area at a reference height (1.5 m) for four tree species. (<b>a</b>) is <span class="html-italic">Khaya senegalensis</span>, (<b>b</b>) is <span class="html-italic">Terminalia neotaliala</span>, (<b>c</b>) is <span class="html-italic">Mangifera indica</span>, and (<b>d</b>) is <span class="html-italic">Ficus microcarpa</span>.</p>
Full article ">Figure 8
<p>Comparison of the air-cooling effect in the shaded area at the reference height (1.5 m) among the four tree species: (<b>a</b>) ΔAT = AT in the open area − AT in the shaded area; (<b>b</b>) reduction rate of AT = ΔAT/AT in the open area × 100%.</p>
Full article ">Figure 9
<p>Daily variations in RH between trees, under trees, and in the open area at a reference height (1.5 m) among the four tree species. (<b>a</b>) is <span class="html-italic">Khaya senegalensis</span>, (<b>b</b>) is <span class="html-italic">Terminalia neotaliala</span>, (<b>c</b>) is <span class="html-italic">Mangifera indica</span>, and (<b>d</b>) is <span class="html-italic">Ficus microcarpa</span>.</p>
Full article ">Figure 10
<p>Comparison of the rate of increase in RH in the shaded area at reference height (1.5 m) among four tree species: (<b>A</b>) ΔRH = RH in the shaded area − RH in the open area; (<b>B</b>) increase rate of RH = ΔRH/RH in the open area × 100%.</p>
Full article ">Figure 11
<p>Daily variations in SR between trees, under trees, and in the open area for the four tree species. (<b>a</b>) is <span class="html-italic">Khaya senegalensis</span>, (<b>b</b>) is <span class="html-italic">Terminalia neotaliala</span>, (<b>c</b>) is <span class="html-italic">Mangifera indica</span>, and (<b>d</b>) is <span class="html-italic">Ficus microcarpa</span>.</p>
Full article ">Figure 12
<p>Comparison of the degree of SR modification in the shaded area among the four tree species. (<b>A</b>) ΔSR = SR in the open area − SR in the shaded area; (<b>B</b>) reduction rate of SR = ΔSR/SR in the open area × 100%.</p>
Full article ">Figure 13
<p>Differences between under-tree and inter-tree SR for the four tree species. ΔSR = inter-tree SR − under-tree SR.</p>
Full article ">Figure 14
<p>Daily variations in ST between trees, under trees, and in the open area for the four tree species. (<b>a</b>) is <span class="html-italic">Khaya senegalensis</span>, (<b>b</b>) is <span class="html-italic">Terminalia neotaliala</span>, (<b>c</b>) is <span class="html-italic">Mangifera indica</span>, and (<b>d</b>) is <span class="html-italic">Ficus microcarpa</span>.</p>
Full article ">Figure 15
<p>Comparison of the degree of reduction in ST in the shaded area among the four tree species. (<b>A</b>) ΔST = ST in the open area − ST in the shaded area; (<b>B</b>) Reduction rate of ST = ΔST/ST in the open area × 100%.</p>
Full article ">Figure 16
<p>Differences between under-tree and inter-tree ST among the four tree species: (<b>A</b>) ΔST = inter-tree ST − under-tree ST; (<b>B</b>) differential rate of ST = ΔST/under-tree ST × 100%.</p>
Full article ">Figure 17
<p>Daily variation in WS in open and shaded areas at a reference height (1.5 m) for the four tree species. (<b>a</b>) is <span class="html-italic">Khaya senegalensis</span>, (<b>b</b>) is <span class="html-italic">Terminalia neotaliala</span>, (<b>c</b>) is <span class="html-italic">Mangifera indica</span>, and (<b>d</b>) is <span class="html-italic">Ficus microcarpa</span>.</p>
Full article ">Figure 18
<p>Comparison of the degree of reduction in WS in the shaded area among the four tree species: (<b>A</b>) ΔWS = WS in the open area − WS in the shaded area; (<b>B</b>) reduction rate of WS = ΔWS/WS in the shaded area × 100%.</p>
Full article ">Figure 19
<p>Daily variation in PET between trees, under trees, and in the open area at a reference height (1.5 m) for the four tree species. The 80% acceptable rate of PET in Guangzhou is 27.25 °C in summer. The neutral PET in Guangzhou is 24.41 °C in summer. (<b>a</b>) is <span class="html-italic">Khaya senegalensis</span>, (<b>b</b>) is <span class="html-italic">Terminalia neotaliala</span>, (<b>c</b>) is <span class="html-italic">Mangifera indica</span>, and (<b>d</b>) is <span class="html-italic">Ficus microcarpa</span>.</p>
Full article ">Figure 20
<p>Comparison of the PET modification between open and shaded areas among four tree species: (<b>A</b>) ΔPET = PET in the open area − PET in the shaded area; (<b>B</b>) reduction rate of PET = ΔPET/PET in the open area × 100%.</p>
Full article ">Figure 21
<p>Differences between under-tree and inter-tree PET among the four tree species: (<b>A</b>) ΔPET = inter-tree PET − under-tree PET; (<b>B</b>) differential rate of PET = ΔPET/under-tree PET × 100%.</p>
Full article ">Figure 22
<p>Street orientation and the average reduction rates of four parameters: ΔWS, ΔAT, ΔPET, ΔST, and ΔSR.</p>
Full article ">
15 pages, 2505 KiB  
Article
Is It Possible to Predict a Forest Insect Outbreak? Backtesting Using Remote Sensing Data
by Anton Kovalev, Olga Tarasova, Vladislav Soukhovolsky and Yulia Ivanova
Forests 2024, 15(8), 1458; https://doi.org/10.3390/f15081458 - 19 Aug 2024
Viewed by 1107
Abstract
In this study, methods are proposed for analyzing the susceptibility of forest stands to attacks by forest insects on the basis of Earth remote sensing data. As an indicator of the state of forest stands, we proposed to use a parameter of the [...] Read more.
In this study, methods are proposed for analyzing the susceptibility of forest stands to attacks by forest insects on the basis of Earth remote sensing data. As an indicator of the state of forest stands, we proposed to use a parameter of the sensitivity of a vegetation index (normalized difference vegetation index; NDVI) during a vegetative period to changes in the radiative temperature of the territory (land surface temperature; LST) determined from satellite data of the Terra/Aqua system. The indicator was calculated as a spectrum of a response function in an integral equation linking changes of NDVI to those of LST. Backtesting was carried out using data from two outbreaks of the Siberian silk moth Dendrolimus sibiricus Tschetv. and outbreaks of the white mottled sawyer Monochamus urussovi Fischer and of the four-eyed fir bark beetle Polygraphus proximus Blandford in taiga forests of Krasnoyarsk Territory in Russia. In addition, the state of fir stands in the year 2023 was examined when damage to the forest stands was not yet noticeable, but Siberian silk moth adults were found in pheromone traps. It was shown that the proposed indicator of susceptibility of forest stands changed significantly 2–3 years before the pest outbreak in outbreak foci of the studied areas. Thus, the proposed indicator can be used to predict outbreaks of insect pests. The proposed approach differs from commonly used remote sensing methods in that, rather than using absolute values of remote indicators (such as, for example, NDVI), it focuses on indicators of the susceptibility of these remote indicators to the characteristics of the natural environment. Since any given point on the planet is characterized by a seasonally varying temperature, it is always possible to determine the sensitivity of a remote sensing indicator to changes in the environment that are not directly related to the absolute value of the indicator. Future studies are expected to examine susceptibility indices as a function of forest stand location and species, and to examine the length of spatial correlation of susceptibility indices, which may provide information on the possible extent of future insect outbreaks. Full article
(This article belongs to the Special Issue Risk Assessment and Management of Forest Pest Outbreaks)
Show Figures

Figure 1

Figure 1
<p>Outbreak foci in Krasnoyarsk Territory (a white line, filled in red on a small map) in Russia. A: An outbreak focus of the Siberian silk moth <span class="html-italic">Dendrolimus sibiricus</span> Tschetv. in the years 2015–2018. B: An outbreak focus of the white mottled sawyer <span class="html-italic">Monochamus urussovi</span> Fisch. C: An outbreak focus of <span class="html-italic">D. sibiricus</span> in 2019–2020. D: An outbreak focus of <span class="html-italic">Polygraphus proximus</span> Blandford. E: The zone where adults of <span class="html-italic">D. sibiricus</span> were found in 2023 and a potential outbreak focus of this pest.</p>
Full article ">Figure 2
<p>A typical time series of seasonal dynamics of NDVI in taiga coniferous forests.</p>
Full article ">Figure 3
<p>A typical time series of seasonal dynamics of LST in taiga coniferous forests.</p>
Full article ">Figure 4
<p>Typical shapes of spectrum <span class="html-italic">H</span>(<span class="html-italic">f</span>) of the function of the response of NDVI to a change in LST for a fir stand in taiga forests of Siberia. 1: Control, 2: the year of the Siberian silk moth outbreak in the Yeniseisk Dist.</p>
Full article ">Figure 5
<p>Dynamics of average seasonal values of NDVI in control forest stands (1) and in foci of outbreaks of Siberian silk moths in the Yenisei District (2). Arrow: the year the outbreak began.</p>
Full article ">Figure 6
<p>Parameters LF and HF of plots damaged by the Siberian silk moth in outbreak focus A and on control plots in different years. 1: At 5 years before the start of visible damage to tree crowns by the pest; 2: 1–2 years before the damage; 3: in the year of the beginning of visible damage to the crowns; 4: within 4 years after the outbreak began; 5: control undamaged stands 1 year before the outbreak.</p>
Full article ">Figure 7
<p>Parameters LF and HF of response function spectra for white mottled sawyer outbreak focus B and for intact stands. 1: Control forest stands for all years before and after the outbreak began; 2: forest stands in the outbreak focus in 2009–2010 (i.e., 3–4 years before the start of the outbreak); 3: forest stands in the outbreak zone 1–2 years before the outbreak and in the year of change in the tree crown color when trees were damaged by the pest (in the year 2013); 4: forest stands in the outbreak focus in 2014.</p>
Full article ">Figure 8
<p>Parameters LF and HF for the 2008–2020 period of coniferous forest stands in outbreak focus C of the Siberian silk moth. 1: Future outbreak foci C in 2010–2014; 2: future outbreak foci C in 2015–2018; 3: outbreak focus C at the beginning of visible damage in 2019, 4: outbreak focus C in the years 2020–2021, 5: control (undamaged) forest stands in 2008–2020.</p>
Full article ">Figure 9
<p>Parameters LF and HF for the years 2006–2019 of fir stands in outbreak focus D of the four-eyed fir bark beetle. 1: Control (undamaged) forest stands in 2006–2019; 2: future outbreak foci assessed in 2006–2010; 3: the foci in 2011–2014, i.e., before the onset of visible damage; 4: areas of foci in the years 2015–2019, i.e., after the trees were cut down.</p>
Full article ">Figure 10
<p>Average components LF and HF for sample plots in zones E1 and E2. 1: Zone E1 in the years 2014–2019; 2: zone E1 in 2020–2023; 3: zone E2 in 2014–2019; E2: zone E2 in 2020–2023.</p>
Full article ">
20 pages, 2086 KiB  
Article
How Can Ecological Product Value Realization Sustainably Enhance the Well-Being of Farmers? A Case Study of Xingyuan Village in China
by Yanqiang Du, Jiying Wang and Juankun Li
Forests 2024, 15(8), 1457; https://doi.org/10.3390/f15081457 - 19 Aug 2024
Viewed by 932
Abstract
Although recent years have witnessed a considerable increase in studies on the economic value of ecological products, the extant literature has overlooked the multidimensional enhancement of ecological product value from the perspective of farmers’ well-being. This research aims to examine how the Realization [...] Read more.
Although recent years have witnessed a considerable increase in studies on the economic value of ecological products, the extant literature has overlooked the multidimensional enhancement of ecological product value from the perspective of farmers’ well-being. This research aims to examine how the Realization of Ecological Product Value (EPVR) serves as a crucial pathway to promoting the overall well-being of farmers in developing countries. Through a case study of a village in southeastern China, this research reveals that EPVR can enhance farmers’ well-being via various mechanisms as follows: (1) Economic solution to enrich farmers’ livelihood diversity, achieved by leveraging rural resource endowments and comparative advantages; (2) Fair social protection program enabling farmers to enjoy ecological benefits and further achieving urban-rural integration; (3) Environmental protection plan that balances production, living, and ecology; and (4) Grassroots governance tool promoting the governance ability to form collaborative governance model in a community of shared interests. This study offers theoretical support for enhancing human well-being through the realization of ecological product value in rural areas. Full article
(This article belongs to the Special Issue Forest Ecosystem Services and Strategy Development)
Show Figures

Figure 1

Figure 1
<p>Data coding process.</p>
Full article ">Figure 2
<p>The Process of EPVR in Xingyuan Village.</p>
Full article ">Figure 3
<p>A conceptual framework of the mechanism for promoting the well-being of farmers through EPVR.</p>
Full article ">
15 pages, 6347 KiB  
Article
Distribution Characteristics and Driving Factors of the Bacterial Community Structure in the Soil Profile of a Discontinuous Permafrost Region
by Qilong Liu, Liquan Song, Siyuan Zou, Xiaodong Wu and Shuying Zang
Forests 2024, 15(8), 1456; https://doi.org/10.3390/f15081456 - 18 Aug 2024
Viewed by 1024
Abstract
Global warming leads to the melting of permafrost, affects changes in soil microbial community structures and related functions, and contributes to the soil carbon cycle in permafrost areas. Located at the southern edge of Eurasia’s permafrost region, the Greater Khingan Mountains are very [...] Read more.
Global warming leads to the melting of permafrost, affects changes in soil microbial community structures and related functions, and contributes to the soil carbon cycle in permafrost areas. Located at the southern edge of Eurasia’s permafrost region, the Greater Khingan Mountains are very sensitive to climate change. Therefore, by analyzing the bacterial community structure, diversity characteristics, and driving factors of soil profiles (active surface layer, active deep layer, transition layer, and permafrost layer) in this discontinuous permafrost region, this research provides support for the study of the carbon cycling process in permafrost regions. The results show that the microbial diversity (Shannon index (4.81)) was the highest at 0–20 cm, and the Shannon index of the surface soil of the active layer was significantly higher than that of the other soil layers. Acidobacteria and Proteobacteria were the dominant bacteria in the active layer soil of the permafrost area, and Chloroflexi, Actinobacteria, and Firmicutes were the dominant bacteria in the permafrost layer. Chloroflexi made the greatest contribution to the bacterial community in the permafrost soil, and Bacteroidota made the greatest contribution to the bacterial community in the active surface soil. The structure, richness, and diversity of the soil bacterial community significantly differed between the active layer and the permafrost layer. The number of bacterial species was the highest in the active layer surface soil and the active layer bottom soil. The difference in the structure of the bacterial community in the permafrost soil was mainly caused by changes in electrical conductivity and soil–water content, while that in the active layer soil was mainly affected by pH and soil nutrient indices. Soil temperature, NO3-N, and pH had significant effects on the structure of the bacterial community. The active layer and permafrost soils were susceptible to environmental information processing and genetic information processing. Infectious disease: the number of bacterial species was significantly higher in the surface and permafrost layers than in the other layers of the soil. In conclusion, changes in the microbial community structure in soil profiles in discontinuous permafrost areas sensitive to climate change are the key to soil carbon cycle research. Full article
(This article belongs to the Section Forest Soil)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Study area location map. (<b>b</b>) Soil profile in permafrost area (0–20 cm: upper active layer; 20–60 cm: lower active layer; 60–80 cm: transition layer; 80–120 cm: permafrost layer).</p>
Full article ">Figure 2
<p>Venn diagram of the structure of the soil bacterial community at different OTU levels in the permafrost region.</p>
Full article ">Figure 3
<p>Principal coordinate analysis (PCoA) of soil bacteria in different soil layers.</p>
Full article ">Figure 4
<p>Composition of soil microbial communities in different soil layers: (<b>a</b>) analysis at phylum level; (<b>b</b>) analysis at genus level.</p>
Full article ">Figure 5
<p>Difference analysis of dominant soil microbial species in different soil layers: (<b>a</b>) analysis at phylum level; (<b>b</b>) analysis at genus level. * <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 ">Figure 6
<p>Correlation analysis of soil physicochemical properties and relative abundance of the major bacterial phyla. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Analysis of the prediction function of microbial communities in different soil layers: (<b>a</b>) grade 1 functional group; (<b>b</b>) grade 2 functional group. a, b, c: If there is one identical marking letter, the difference is not significant, and if there is different marking letter, the difference is significant.</p>
Full article ">Figure 8
<p>(<b>a</b>) Analysis of the metabolic pathways of soil bacteria regulated by thawing permafrost. The red line represents the positive path, and the blue line represents the negative path. Insignificant effects are indicated by dotted arrows. * <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. (<b>b</b>) Normalization effects between variables in different permafrost regions (direct and indirect normalization effects).</p>
Full article ">
22 pages, 10272 KiB  
Article
Monitoring of Flash Drought on the Loess Plateau and Its Impact on Vegetation Ecosystems
by Yanmin Jiang, Haijing Shi, Zhongming Wen, Xihua Yang, Youfu Wu, Li Li, Yuxin Ma, John R. Dymond, Minghang Guo, Junfeng Shui and Hong Hu
Forests 2024, 15(8), 1455; https://doi.org/10.3390/f15081455 - 18 Aug 2024
Viewed by 917
Abstract
Flash drought (FD) has attracted much attention due to its severe stress on vegetation ecosystems. Yet to date, the impacts of FD on vegetation ecosystems in different regions have not been fully evaluated and explored, especially for ecologically fragile areas. In this study, [...] Read more.
Flash drought (FD) has attracted much attention due to its severe stress on vegetation ecosystems. Yet to date, the impacts of FD on vegetation ecosystems in different regions have not been fully evaluated and explored, especially for ecologically fragile areas. In this study, we identified the FD events in the Loess Plateau from 2000 to 2023 based on the attenuation rate in soil moisture percentile over time. The evolution process of FD, the driving roles of meteorological conditions and the responses of different vegetation types to FD were explored by vegetation indicators such as solar-induced chlorophyll fluorescence (SIF), SIFyield, SIF-RCI, etc. The results showed that FD events were predominantly concentrated in wetter areas with dense vegetation, with the highest frequency being 29. Meteorological factors contributed differently to the occurrence and development of FD. The responses of vegetation to FD were not only related to vegetation types (cropland was more sensitive to FD than forest and grassland) but were also significantly influenced by background climate. The SIFyield anomaly of vegetation was more sensitive than SIF anomaly and SIF-RCI. The results advance our understanding of the formation mechanisms of FD and facilitate the exploration of vegetative photosynthetic responses to FD. Full article
(This article belongs to the Special Issue Application of Remote Sensing in Vegetation Dynamic and Ecology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The location of the study area, China’s LP; (<b>b</b>) The climate areas and distribution of meteorological stations in the LP; (<b>c</b>) Annual trend of NDVI in the LP from 1999 to 2020; (<b>d</b>–<b>f</b>) Land uses of the LP in 1990, 2010, and 2020, respectively.</p>
Full article ">Figure 2
<p>Pentad average SM percentile in the LP of 2017. The data come from a grid unit on the Loess Plateau (39.399° N, 108.299° E).</p>
Full article ">Figure 3
<p>Data processing framework diagram.</p>
Full article ">Figure 4
<p>Frequency (<b>a</b>), average decline rate (<b>b</b>), average duration (<b>c</b>), and average severity (<b>d</b>) of FDs on the LP from 2000 to 2023.</p>
Full article ">Figure 5
<p>Statistical charts for different climate areas. Subfigures (<b>a</b>–<b>d</b>) represent the frequency, average RI, average duration and average severity of flash drought, respectively.</p>
Full article ">Figure 6
<p>Variation of SM percentile FD events of all grids in four different climate areas. The t represents the onset of FD. The t − 2 and t − 1 denote the 1 pentad and 2 pentads prior to t, while t + 1–t + 7 represent the lagged 1–7 pentads of t, respectively.</p>
Full article ">Figure 7
<p>Changes of SM percentile in 2017.</p>
Full article ">Figure 8
<p>Changes of RI in 2017.</p>
Full article ">Figure 9
<p>Variations in meteorological factors during the FD period. Subfigures (<b>a</b>–<b>f</b>) represent the Rainfall, relative humidity, maximum temperature, mean temperature, potential evapotranspiration and average wind speed, respectively.</p>
Full article ">Figure 10
<p>Spatial variations in meteorological factors anomalies during the FD period from June to July 2017.</p>
Full article ">Figure 11
<p>Temporal variations in SM and ecological indicators of SIF, SIF<sub>yield</sub>, SIF-RCI, NDVI, and APAR during FD events for the four climate areas. Thick lines indicate the median values, while the shaded regions depict the range of variability, spanning from the 25th to the 75th percentiles observed during flash drought events.</p>
Full article ">Figure 12
<p>SM and photosynthesis anomalies of different vegetation types in three climate areas during the FD in 2017.</p>
Full article ">
12 pages, 3083 KiB  
Article
Needle Biomass Turnover Rate in Scots Pine Stands of Different Ages
by Mieczysław Turski, Ignacy Korczyński, Adrian Łukowski and Andrzej Węgiel
Forests 2024, 15(8), 1454; https://doi.org/10.3390/f15081454 - 18 Aug 2024
Viewed by 626
Abstract
Understanding needle biomass turnover rates in Scots pine (Pinus sylvestris L.) stands is crucial for modelling forest ecosystem dynamics and nutrient cycling. This study examined needle litterfall and biomass turnover in Scots pine stands of varying ages in temperate forests (western Poland). [...] Read more.
Understanding needle biomass turnover rates in Scots pine (Pinus sylvestris L.) stands is crucial for modelling forest ecosystem dynamics and nutrient cycling. This study examined needle litterfall and biomass turnover in Scots pine stands of varying ages in temperate forests (western Poland). The research focused on determining how stand age affects needle biomass, litterfall and the associated turnover rates. Data were collected from 20 Scots pine stands aged 26 to 90 years, and needle litterfall was measured and analysed in relation to stand characteristics such as age, density and biomass. The average annual needle litter production of the sampled Scots pine stands was 2008 kg·ha−1·year−1, similar to the values previously reported for this tree species in other temperate forests in Europe. The average needle biomass turnover rate for sampled Scots pine stands was 23.4%. We could not support the hypothesis that this parameter depended on the age of the Scots pine stand. The needle biomass turnover rate showed a positive correlation with crown length and a negative correlation with stand density due to the very weak correlations; however, further research is needed to confirm these relationships. Despite this, the parameter can be used to estimate needle litterfall and can be applicable to conditions corresponding to those of temperate forests in Central and Western Europe. This study also highlights the need for further research on needle biomass turnover in temperate forests to improve the accuracy of carbon and nutrient cycling models. This work contributes to a deeper understanding of the role of needle litterfall in maintaining soil fertility and forest productivity, offering insights into sustainable forest management and conservation strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Location of sampled stands in western Poland. Sample plot numbers are explained in <a href="#forests-15-01454-t001" class="html-table">Table 1</a>. (source of spatial data: <a href="http://OpenStreetMap.org" target="_blank">OpenStreetMap.org</a>, accessed on 9 July 2024).</p>
Full article ">Figure 2
<p>An example of the randomly distributed eight sample plots (green circles) in a sampled stand, with the locations of the trays for collecting fallen needles (purple circles).</p>
Full article ">Figure 3
<p>Correlation matrix for different parameters of 20 sampled Scots pine stands. Significance level: * <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 ">Figure 4
<p>Correlation scatter plots with linear regression (solid line) and its confidence interval (dashed lines) between four variables: needle mass fraction (<b>a</b>), needle biomass (<b>b</b>), crown length (<b>c</b>), stand density (<b>d</b>) and needle biomass turnover rate.</p>
Full article ">
17 pages, 4059 KiB  
Article
Factors Influencing the Change of Phyllosphere Microbial Community of Three Populus spp. in the Same Habitat
by Xin Yin, Weixi Zhang, Dan Li, Ran Wang, Xinyao Cong, Zhongyi Pang, Yanhui Peng, Yang Ge, Wenxu Zhu and Changjun Ding
Forests 2024, 15(8), 1453; https://doi.org/10.3390/f15081453 - 18 Aug 2024
Viewed by 674
Abstract
Plant leaves harbor a rich diversity of bacteria and fungi that, through their interactions with host plants, assume an influential role in plant physiological and metabolic processes. The unique phyllosphere environment of different plant species may shape and select distinct phyllosphere microbial communities. [...] Read more.
Plant leaves harbor a rich diversity of bacteria and fungi that, through their interactions with host plants, assume an influential role in plant physiological and metabolic processes. The unique phyllosphere environment of different plant species may shape and select distinct phyllosphere microbial communities. While most academic research has focused on the phyllosphere microorganisms within the same plant variety, there is relatively limited research on the phyllosphere microbial communities between different varieties. Populus L. is a typical tree species in temperate monsoon climates, widely distributed in northern China, and it constitutes a crucial component of China’s forestry resources. For the purpose of this study, we investigated the community structure and diversity of phyllosphere fungi and bacteria in different poplar varieties under identical growth conditions to elucidate the main factors contributing to differences in phyllosphere microbial communities among these varieties. Our findings revealed variations in nitrogen, phosphorus, starch, and soluble sugar contents among the three poplar species studied. Additionally, there were considerable disparities in both abundance and α diversity index of phyllosphere fungal and bacterial communities among these species. At the phylum level, Ascomycota and Basidiomycota have been identified as the dominant fungal communities; while Proteobacteria and Actinomycetes were dominant bacterial communities. The correlation analysis pointed out that chemical traits in the leaves, in particular the total phosphorus and the quantity of soluble sugar, had a significant correlation with the structure and diversity of the microbial community residing in the phyllosphere. Overall, our results demonstrate that even under identical site conditions, each poplar species harbors its own unique phyllosphere microbial community composition as well as distinct leaf characteristics—highlighting host plant diversity as a crucial factor driving differences in phyllosphere microbial composition. Full article
Show Figures

Figure 1

Figure 1
<p>Venn diagram illustrating unique and shared ASVs of phyllosphere microbial community for three different samples. (<b>a</b>) Unique and shared ASVs of phyllosphere fungal communities in three different samples; (<b>b</b>) Unique and shared ASVs of phyllosphere bacteria communities in three different samples.</p>
Full article ">Figure 2
<p>Rarefaction Curve. The observed diversity of samples is influenced by the sequencing depth, which is reflected in the flattening of the curve. The lack of steepness in the curve suggests that the sequencing outcomes successfully reflect the existing diversity within the current samples, encompassing numerous novel ASVs. Conversely, a non-flattened curve suggests that alpha diversity has not reached saturation. (<b>a</b>) Rarefaction Curve of fungi. (<b>b</b>) Rarefaction Curve of bacteria.</p>
Full article ">Figure 3
<p>Phyllosphere microbial diversity (* Reveals a notable variance with a confidence level of 95%. The black line represents a direct and significant difference between the two) (<b>a</b>) Alpha diversity analysis of phyllosphere fungal community. (<b>b</b>) Alpha diversity analysis of phyllosphere bacterial community.</p>
Full article ">Figure 4
<p>Analysis of taxonomic composition at the phylum level. (<b>a</b>) Analysis of fungal taxonomic composition at the phylum level. (<b>b</b>) Analysis of bacterial taxonomic composition at the phylum level.</p>
Full article ">Figure 5
<p>Analysis of taxonomic composition at the genus level. (<b>a</b>) Analysis of fungal taxonomic composition at the genus level. (<b>b</b>) Analysis of bacterial taxonomic composition at the genus level.</p>
Full article ">Figure 6
<p>Hierarchical clustering analysis of phyllosphere microbial communities of different <span class="html-italic">Populus</span> spp. (at the genus level). (<b>a</b>) Hierarchical clustering analysis of phyllosphere fungal communities of different <span class="html-italic">Populus</span> spp. (<b>b</b>) Hierarchical clustering analysis of phyllosphere bacterial communities of different <span class="html-italic">Populus</span> spp.</p>
Full article ">Figure 7
<p>LEFse analysis of phyllosphere fungal and bacterial community at the genus level. (<b>a</b>) LEFse analysis of phyllosphere fungal community at the genus level. (<b>b</b>) LEFse analysis of phyllosphere bacterial community at the genus level.</p>
Full article ">Figure 8
<p>Correlation heatmap of leaf nutrient factors and α diversity of phyllosphere fungal and bacterial community. (<b>a</b>) Correlation heatmap of leaf nutrient factors and α diversity of phyllosphere fungal community. (<b>b</b>) Correlation heatmap of leaf nutrient factors and α diversity of phyllosphere bacterial community. * indicates that there is a significant correlation at the level of <span class="html-italic">p</span> &lt; 0.05; ** indicates a significant correlation at the level of <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 9
<p>Correlation heatmap of leaf nutrient factors and α diversity of phyllosphere fungal and bacterial community at genus level. (<b>a</b>) Correlation heatmap of leaf nutrient factors and α diversity of phyllosphere fungal community at genus level. (<b>b</b>) Correlation heatmap of leaf nutrient factors and α diversity of phyllosphere bacterial community at genus level. * signifies a statistically significant correlation at <span class="html-italic">p</span> &lt; 0.05, whereas ** denotes an even stronger statistical significance at <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
21 pages, 6440 KiB  
Article
Improving Urban Forest Expansion Detection with LandTrendr and Machine Learning
by Zhe Liu, Yaru Zhang and Xi Zheng
Forests 2024, 15(8), 1452; https://doi.org/10.3390/f15081452 - 17 Aug 2024
Viewed by 1062
Abstract
Annual urban forest expansion dynamics are crucial for assessing the benefits and potential issues associated with vegetation accumulation over time. LandTrendr (Landsat-Based Detection of Trends in Disturbance and Recovery) can efficiently detect the dynamics of interannual land cover change, but it has difficulty [...] Read more.
Annual urban forest expansion dynamics are crucial for assessing the benefits and potential issues associated with vegetation accumulation over time. LandTrendr (Landsat-Based Detection of Trends in Disturbance and Recovery) can efficiently detect the dynamics of interannual land cover change, but it has difficulty distinguishing urban forest expansion from urban surface rapid conversions, as changes are usually filtered by magnitude-of-change thresholds. To accurately detect annual urban forest expansion dynamics, we developed an improved method using random forest-supervised classification to filter urban forests. We further enhanced the performance of the improved method by incorporating trend features between segments. Additionally, we tested two threshold-based filtering baseline methods. These methods were tested with various spectral and parameter combinations in Beijing’s Central District and the 1st Greenbelt from 1994 to 2022. The improved method with trend features achieved the highest average accuracy of 89.35%, representing a 25% improvement over baseline methods. Post-change trend features aided in accurate identification, while quantitative features rather than extremum features were more important in filtering. The improved method with trend features tested in Beijing’s 2nd Greenbelt also showed an accuracy of 88.27%, confirming its stability. SWIR2 and a higher maximum segment number are efficient for filtering by providing the most detailed dynamics. Accurate annual expansion dynamic mapping offers insights into change rates and precise expansion years, providing a new perspective for urban forest research and management. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The four steps of this study. RF: random forest, MAG: magnitude, DUR: duration, DSNR: disturbance signal-to-noise ratio.</p>
Full article ">Figure 2
<p>Overview of the study area and data utilized. (<b>a</b>) location of the study area; (<b>b</b>) boundary of the study area with verified points and land cover types.</p>
Full article ">Figure 3
<p>An example of the land cover changes keyframes with the corresponding segments fitted by LandTrendr under the SWIR2. The red cross in keyframes represents the sample point processed by the LandTrendr, the green arrow indicates the process extracted by the greatest magnitude of change, and the dark red arrow shows the actual urban forest construction event.</p>
Full article ">Figure 4
<p>Segment identification results at all preset parameter combinations: (<b>a</b>) amount of all segments fitted by LandTrendr from the training and validation sets; (<b>b</b>) amount of the identified changed segments from the training and validation sets.</p>
Full article ">Figure 5
<p>Urban land use transition pattern in urban forest construction areas (using the segmentation results of the SWIR2 band as an example).</p>
Full article ">Figure 6
<p>Average performance and standard deviation of the single-band/index filtering with thresholds in identifying change events and their timings. The higher brightness indicates a higher max_segments.</p>
Full article ">Figure 7
<p>Performance evaluation of the multispectral secondary classification in identifying change events and timings: (<b>a</b>) the ratio of true positives and true negatives in identifying construction events; (<b>b</b>) the AUC in identifying construction events and the accuracy in identifying change time.</p>
Full article ">Figure 8
<p>Accuracy of identifying change time for all parameter combinations using the improved method based on single spectral supervised classification: (<b>a</b>) accuracy of the improved method based on basic features; (<b>b</b>) accuracy of the improved method based on relative trend features.</p>
Full article ">Figure 9
<p>The mean and standard deviation of feature importance under different combinations of preset parameters and bands/indices in the improved method with trend features.</p>
Full article ">Figure 10
<p>(<b>a</b>) Urban forest dynamic map of the study area; (<b>b</b>) annual and cumulative increase in urban forest area in the study area.</p>
Full article ">Figure 11
<p>Expansion of vegetation types in the Central District and the 1st Greenbelt: (<b>a</b>) the distributions of vegetation type before and after expansion; (<b>b</b>) expansion process during the expansion process.</p>
Full article ">Figure 12
<p>Examples of change time map comparison between the multispectral secondary classification method and the improved method with trend features. Areas in red boxes are areas with differential identification results. CT: change time, NC: no change: (<b>a</b>) keyframe Google Earth <sup>TM</sup> historical image of the areas over the time series; (<b>b</b>) change time maps based on the improved method with trend features; (<b>c</b>) change time maps based on the baseline method of multispectral secondary classification.</p>
Full article ">Figure A1
<p>Feature importance of the secondary classification.</p>
Full article ">
14 pages, 4349 KiB  
Article
Alien Species Introduction and Demographic Changes Contributed to the Population Genetic Structure of the Nut-Yielding Conifer Torreya grandis (Taxaceae)
by Yuming Tan, Qian Ou, Xin Huang, Yujin Wang and Yixuan Kou
Forests 2024, 15(8), 1451; https://doi.org/10.3390/f15081451 - 17 Aug 2024
Viewed by 940
Abstract
Understanding population genetic structure and its possible causal factors is critical for utilizing genetic resources and genetic breeding of economically important plants. Although Torreya grandis is an important conifer producing nuts in China, little is known about its population structure, let alone the [...] Read more.
Understanding population genetic structure and its possible causal factors is critical for utilizing genetic resources and genetic breeding of economically important plants. Although Torreya grandis is an important conifer producing nuts in China, little is known about its population structure, let alone the causal factors that shaped its genetic variation pattern and population structure. In this work, we intended to characterize the genetic variation pattern and population structure of the nut-yielding conifer T. grandis throughout its whole geographical distribution and further explore the potentially causal factors for the population structure using multiple approaches. A moderate level of genetic diversity and a novel population structure were revealed in T. grandis based on eleven robust EST-SSR loci and three chloroplast fragments. Alien genetic composition derived from the closely related species T. nucifera endemic to Japan was detected in the Kuaiji Mountain area, where the seed quality of T. grandis is considered the best in China. Demography history and niche modeling were inferred and performed, and the contribution of geographic isolation to its population structure was compared with that of environmental isolation. Significant demographic changes occurred, including a dramatic population contraction during the Quaternary, and population divergence was significantly correlated with geographic distance. These results suggested that early breeding activities and demographic changes significantly contributed to the population structure of T. grandis. In turn, the population structure was potentially associated with the excellent variants and adaptation of cultivars of T. grandis. The findings provide important information for utilizing genetic resources and genetic breeding of T. grandis in the future. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical distribution and network of chloroplast haplotypes (H1-H7) in <span class="html-italic">T. grandis</span> and its two closely related species, <span class="html-italic">T. nucifera</span> and <span class="html-italic">T. jackii</span>. The numbers 1-18 correspond to the population codes in <a href="#forests-15-01451-t001" class="html-table">Table 1</a>. The sizes of circles in the network are proportional to the haplotype frequencies, and the mutation steps among haplotypes of more than one are marked on each branch. The photograph on the right illustrates an ancient tree in population 16 (Zhuji, Zhejiang, China) that is more than 400 years old.</p>
Full article ">Figure 2
<p>Population structure of <span class="html-italic">T. grandis</span>, its genetic relationship with <span class="html-italic">T. nucifera</span> and <span class="html-italic">T. jackii</span>, and the correlation of its population divergence with geographic distances. (<b>a</b>) Population structure was inferred using Structure analysis with the number of clusters (<span class="html-italic">K</span>) varying from 2 to 12 with different colors and (<b>b</b>) principal coordinates analysis (PCoA). The numbers 1-18 correspond to the population codes in <a href="#forests-15-01451-t001" class="html-table">Table 1</a>. (<b>c</b>) Correlation among genetic distances (<span class="html-italic">F</span><sub>ST</sub>) and geographic distances was estimated by the Mantel test. The blue circles and gray background refer to the genetic distances (<span class="html-italic">F</span><sub>ST</sub>) and confidence interval of the linear correlation, respectively.</p>
Full article ">Figure 3
<p>Eight alternative demographic scenarios for <span class="html-italic">T. grandis</span> simulated by approximate Bayesian computation in DIYABC based on EST-SSR loci. <span class="html-italic">N</span><sub>1</sub> and <span class="html-italic">N</span><sub>2</sub> represent current population sizes, and <span class="html-italic">N</span><sub>A</sub> represents the ancestral population size. <span class="html-italic">N</span><sub>1a</sub>, <span class="html-italic">N</span><sub>1b</sub>, <span class="html-italic">N</span><sub>2a</sub>, and <span class="html-italic">N</span><sub>2b</sub> represent population sizes between the ancestral population and the current population. <span class="html-italic">t</span><sub>1</sub> and <span class="html-italic">t</span><sub>2</sub> represent times of population changes. The optimal scenario is shown with gray background.</p>
Full article ">Figure 4
<p>Climate niches of <span class="html-italic">T. grandis</span> predicted using ecological niche modeling in MAXENT. Predicted distributions are shown during present-day (PRESENT), Mid-Holocene (MH), Last Glacial Maximum (LGM), and Last Interglacial (LIG) climatic periods. Paleoclimate data for the MH, LGM, and LIG under MIROC model were employed.</p>
Full article ">Figure 5
<p>The effects of heterogeneous environment on population genetic divergence evaluated using principal component analysis (PCA) of 19 bioclimatic variables. The numbers 1–16 correspond to the population codes in <a href="#forests-15-01451-t001" class="html-table">Table 1</a>. Four genetically distinct groups revealed in Structure analysis are shown with different colors.</p>
Full article ">
17 pages, 4904 KiB  
Article
Reconstructing the Temperature and Precipitation Changes in the Northern Part of the Greater Khingan Mountains over the Past 210 Years Using Tree Ring Width Data
by Zhaopeng Wang, Dongyou Zhang, Tongwen Zhang, Xiangyou Li, Xinrui Wang, Taoran Luo, Shubing Zhong and Kexin Song
Forests 2024, 15(8), 1450; https://doi.org/10.3390/f15081450 - 16 Aug 2024
Viewed by 618
Abstract
In northeastern China, simultaneous reconstruction of temperature and precipitation changes in the same region using tree ring data has not yet been reported, limiting our understanding of the historical climate. Using tree ring samples from the Greater Khingan Mountains, it was established that [...] Read more.
In northeastern China, simultaneous reconstruction of temperature and precipitation changes in the same region using tree ring data has not yet been reported, limiting our understanding of the historical climate. Using tree ring samples from the Greater Khingan Mountains, it was established that there are five standardized tree ring width chronologies of Pinus sylvestris var. mongolica at five elevations. Correlation analyses revealed significant relationships between the tree ring chronologies and climate data for multiple months. Specifically, the correlation coefficient between the average minimum temperature from May to July and the composite chronologies of mid–high and mid-elevations was 0.726, whereas that between the total precipitation from August to July and the low-elevation chronology was 0.648 (p < 0.01). Based on these findings, we reconstructed two series: the average minimum temperature from May to July over the past 211 years and the total precipitation from August to July over the past 214 years. The reconstructed sequences revealed changes in the average minimum temperature from 1812 to 2022 and precipitation from 1809 to 2022 in the northern part of the Greater Khingan Mountains. The variances explained by the reconstruction equations were 0.528 and 0.421 (adjusted R-squared: 0.520 and 0.411), with F-test values of 65.896 and 42.850, respectively, exceeding the significance level of 0.01. The reliability of the reconstructed sequences was validated by historical records of meteorological disasters and the reconstruction results in the surrounding area. The reconstructed temperature and precipitation sequences exhibited distinct patterns of temperature fluctuations, dry–wet changes, and periodic oscillations. The region experienced two warm periods (1896–1909 and 2006–2020), two cold periods (1882–1888 and 1961–1987), a wet period (1928–1938), a drought period (1912–1914), and a period prone to severe drought events (1893–1919) during the past 210 years. The temperature series showed periodicities of 2–2.5 years, 3.9 years, 5.2 years, and 68 years, while the precipitation series exhibited periodicities of 2.1 years, 2.5 years, and 2.8 years, possibly related to El Niño–Southern Oscillation (ENSO) events, quasi-biennial oscillation, and Pacific Decadal Oscillation (PDO). Spatial correlation analysis indicated that the reconstructed temperature and precipitation sequences accurately represented the hydrothermal changes in the study area. Full article
(This article belongs to the Section Forest Ecology and Management)
Show Figures

Figure 1

Figure 1
<p>Climatic characteristics of the northern Greater Khingan Mountains from 1960 to 2020. (<b>A</b>) Changes in mean monthly temperature and total monthly precipitation. (<b>B</b>) Changes in annual mean temperature. (<b>C</b>) Changes in annual precipitation. <span class="html-italic">P</span>, precipitation; <span class="html-italic">T</span>, mean temperature; <span class="html-italic">T</span><sub>max</sub>, mean maximum temperature; <span class="html-italic">T</span><sub>min</sub>, mean minimum temperature.</p>
Full article ">Figure 2
<p>Standardized chronology, sample depth, expressed population signal, and mean inter-series correlation at different elevations. (<b>A</b>) High elevation. (<b>B</b>) Medium–high elevation. (<b>C</b>) Medium elevation. (<b>D</b>) Medium–low elevation. (<b>E</b>) Low elevation.</p>
Full article ">Figure 3
<p>Pearson correlation between climate data and chronology. (<b>A</b>) High elevation. (<b>B</b>) Medium–high elevation. (<b>C</b>) Medium elevation. (<b>D</b>) Medium–low elevation. (<b>E</b>) Low elevation. p, previous year; c, current year. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Comparison of reconstructed and measured values. ((<b>A</b>) minimum temperature; (<b>B</b>) precipitation) and reconstructed sequence ((<b>C</b>) minimum temperature; (<b>D</b>) precipitation).</p>
Full article ">Figure 5
<p>Reconstruction of the mean minimum temperature and precipitation series and the 210-year sliding average (<b>A</b>,<b>C</b>) and the cumulative distance horizon (<b>B</b>,<b>D</b>).</p>
Full article ">Figure 6
<p>Reconstruction of wavelet analysis of mean minimum temperature (<b>A</b>) and precipitation (<b>B</b>).</p>
Full article ">Figure 7
<p>Multi-window spectral analysis for reconstruction of mean minimum temperature (<b>A</b>) and precipitation (<b>B</b>). <span style="color:#ED0000">— </span>(red line), <span class="html-italic">p</span> &lt; 0.05; <span style="color:#4472C4">— </span>(blue line), <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Comparison of reconstructed precipitation sequences with neighboring sequences in this paper: (<b>A</b>) reconstruction of minimum temperature in Hailar; (<b>B</b>) reconstruction of minimum temperature in the Mangui area (this study).</p>
Full article ">Figure 9
<p>Reconstruction of spatial correlation between mean minimum temperature and land (<b>A</b>) and sea (<b>B</b>) temperatures.</p>
Full article ">Figure 10
<p>Spatial correlation analysis of measured and reconstructed mean minimum temperature and precipitation series with grid point data. (<b>A</b>) measured mean minimum temperature; (<b>B</b>) reconstructed mean minimum temperature; (<b>C</b>) measured precipitation; (<b>D</b>) reconstructed precipitation.</p>
Full article ">
Previous Issue
Back to TopTop