[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (36)

Search Parameters:
Keywords = broomcorn millet

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 10735 KiB  
Article
Comparative Transcriptomic Analysis Reveals Domestication and Improvement Patterns of Broomcorn Millet (Panicum miliaceum L.)
by Xinyu Zhao, Minxuan Liu, Chunxiang Li, Jingyi Zhang, Tianshu Li, Fengjie Sun, Ping Lu and Yue Xu
Int. J. Mol. Sci. 2024, 25(20), 11012; https://doi.org/10.3390/ijms252011012 - 13 Oct 2024
Viewed by 324
Abstract
Broomcorn millet (Panicum miliaceum L.) is one of the earliest crops, domesticated nearly 8000 years ago in northern China. It gradually spread across the entire Eurasian continent, as well as to America and Africa, with recent improvement in various reproductive and vegetative [...] Read more.
Broomcorn millet (Panicum miliaceum L.) is one of the earliest crops, domesticated nearly 8000 years ago in northern China. It gradually spread across the entire Eurasian continent, as well as to America and Africa, with recent improvement in various reproductive and vegetative traits. To identify the genes that were selected during the domestication and improvement processes, we performed a comparative transcriptome analysis based on wild types, landraces, and improved cultivars of broomcorn millet at both seeding and filling stages. The variations in gene expression patterns between wild types and landraces and between landraces and improved cultivars were further evaluated to explore the molecular mechanisms underlying the domestication and improvement of broomcorn millet. A total of 2155 and 3033 candidate genes involved in domestication and a total of 84 and 180 candidate genes related to improvement were identified at seedling and filling stages of broomcorn millet, respectively. The annotation results suggested that the genes related to metabolites, stress resistance, and plant hormones were widely selected during both domestication and improvement processes, while some genes were exclusively selected in either domestication or improvement stages, with higher selection pressure detected in the domestication process. Furthermore, some domestication- and improvement-related genes involved in stress resistance either lost their functions or reduced their expression levels due to the trade-offs between stress resistance and productivity. This study provided novel genetic materials for further molecular breeding of broomcorn millet varieties with improved agronomic traits. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Principal component analysis (PCA) of broomcorn millet samples based on FPKM values of genes. (<b>A</b>) PCA plot of 18 samples of broomcorn millet at seedling stage. (<b>B</b>) PCA plot of 16 samples of broomcorn millet at filling stage excluding two outliers (WNM8F-2 and ZLN5F-3). Samples of wild types, landraces, and improved cultivars are indicated in red, green, and blue symbols, respectively. WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively; WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively.</p>
Full article ">Figure 2
<p>Venn diagram of genes expressed in wild types, landraces, and improved cultivars of broomcorn millets at seedling stage (<b>A</b>) and filling stage (<b>B</b>). WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively; WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively.</p>
Full article ">Figure 3
<p>Detection of co-expression network in experimental groups of wild types, landraces, and improved cultivars of broomcorn millets at seedling and filling stages. (<b>A</b>) Hierarchical clustering tree of genes with assigned modules indicated. The main tree branches constitute 21 modules labeled by different colors, and the genes are represented by the leaves in the branches. The gray module is reserved for genes not assigned to any of the other 20 modules. WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively, and WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively. (<b>B</b>) Associations between 21 modules (21 rows) and 6 traits (6 columns), i.e., WS, LS, IS, WF, LF, and IF. Each cell at the row–column intersection contains the correlation coefficient between the corresponding module and trait as well as the corresponding <span class="html-italic">p</span> values given in parentheses and are color-coded based on the color legend.</p>
Full article ">Figure 4
<p>Differential gene expression analysis between wild types and landraces as well as between landraces and improved cultivars of broomcorn millet at seedling and filling stages, showing volcano plots of gene expression variation between WS and LS (<b>A</b>), WF and LF (<b>B</b>), LS and IS (<b>C</b>), and LF and IF (<b>D</b>). Each dot represents a gene, with genes up-regulated and down-regulated indicated in red and green, respectively; black dots represent the genes showing no significant difference. (<b>E</b>) Histogram showing the number of differentially expressed genes (DEGs) up-regulated and down-regulated in WS vs. LS, WF vs. LF, LS vs. IS, and LF vs. IF. Two vertical dashed lines represent |fold change| = 2; the horizontal dashed line represents significant <span class="html-italic">p</span> value of 0.05. WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively; WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively.</p>
Full article ">Figure 5
<p>Gene Ontology (GO) annotation of differentially expressed genes (DEGs) detected in WS vs. LS (<b>A</b>), WF vs. LF (<b>B</b>), LS vs. IS (<b>C</b>), and LF vs. IF (<b>D</b>) of broomcorn millet. <span class="html-italic">X</span>-axis represents the GO terms of three categories, i.e., biological process, cellular component, and molecular function. <span class="html-italic">Y</span>-axis represent the number of DEGs annotated to the GO term (right) and percentage of number of DEGs annotated in all DEGs (left). WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively; WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively.</p>
Full article ">Figure 6
<p>Chord plot of GO enrichment analysis based on differentially expressed genes (DEGs) in WS vs. LS (<b>A</b>), WF vs. LF (<b>B</b>), LS vs. IS (<b>C</b>), and LF vs. IF (<b>D</b>) of broomcorn millet. The enriched DEGs are linked via ribbons to their assigned GO terms. The enriched DEGs are arranged according to log (Fold Change), which is displayed as a blue-to-red coding scheme next to the selected genes. The up-regulated and down-regulated genes are indicated in red and blue, respectively, and the color intensity represents the relative value of fold change. WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively; WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively.</p>
Full article ">Figure 7
<p>COG classification of differentially expressed genes (DEGs) detected in WS vs. LS (<b>A</b>), WF vs. LF (<b>B</b>), LS vs. IS (<b>C</b>), and LF vs. IF (<b>D</b>) of broomcorn millet. WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively; WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively.</p>
Full article ">Figure 8
<p>Kyoto Encyclopedia of Genes and Genomes (KEGG) metabolic pathway enrichment analysis of differentially expressed genes (DEGs) in WS vs. LS (<b>A</b>), WF vs. LF (<b>B</b>), LS vs. IS (<b>C</b>), and LF vs. IF (<b>D</b>) of broomcorn millet. The top 20 most enriched pathways are listed on <span class="html-italic">Y</span>-axis; the <span class="html-italic">X</span>-axis presents the enrichment factor based on the ratio of the proportion of the DEGs annotated to a pathway to the proportion of all genes annotated to the pathway. The color intensity of the dots stands for q value (adjusted <span class="html-italic">p</span> value), and the size of the dots represents the number of DEGs enriched in the pathway. WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively; WF, LF, and IF present wild types, landraces, and improved cultivars at filling stage, respectively.</p>
Full article ">Figure 9
<p>Quantitative real-time PCR (qRT-PCR) validation of expression levels of eight genes based on WS vs. LS (<b>A</b>–<b>D</b>) and LS vs. IS (<b>E</b>–<b>H</b>). <span class="html-italic">Y</span>-axis shows the relative gene expression levels based on qRT-PCR (<b>left</b>) and RNA-Seq (<b>right</b>). The expression levels of genes based on qRT-PCR are normalized by <span class="html-italic">18S rRNA</span> and determined using the 2<sup>–ΔΔCt</sup> method. The significant difference in the gene expression levels between WS vs. LS and between LS vs. IS was determined using Student’s <span class="html-italic">t</span>-test based on <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 (***), and <span class="html-italic">p</span> &lt; 0.0001 (****), respectively. WS, LS, and IS indicate wild types, landraces, and improved cultivars at seedling stage, respectively.</p>
Full article ">
15 pages, 2250 KiB  
Article
The Genetic Diversity Assessment of Broomcorn Millet (Panicum miliaceum) and the Construction of a Mini-Core Collection
by Jiandong Ren, Xiaohan Yu, Xiaoxing Wang, Yue Wang, Xuxia Xin, Ruonan Wang, Yingxing Zhang, Minxuan Liu and Jishan Xiang
Agronomy 2024, 14(10), 2226; https://doi.org/10.3390/agronomy14102226 - 27 Sep 2024
Viewed by 523
Abstract
Broomcorn millet (Panicum miliaceum L.) is a crop with a good ability to adapt to the environment. Over 8800 accessions have been collected in the national gene bank of China. The huge quantity of germplasms made it difficult for analysis and evaluation. [...] Read more.
Broomcorn millet (Panicum miliaceum L.) is a crop with a good ability to adapt to the environment. Over 8800 accessions have been collected in the national gene bank of China. The huge quantity of germplasms made it difficult for analysis and evaluation. Although a broomcorn millet core collection (CC) comprising 780 accessions was established, the number is still too large for researchers to explore in depth. In this study, the genetic diversity of 634 broomcorn millet accessions from the core collection was analyzed based on SSR markers. A mini-core collection (MC) containing 256 accessions was extracted. The mini-core collection accounted for less than half of the original core collection and only about 2.8% of the total resources but still provided a good representation. In addition, the results of this study validated that Shanxi is the origin of broomcorn millet, and accessions from the South region may contain novel genes. In conclusion, this study provides a comprehensive characterization of the genetic diversities of broomcorn millet core collection in China. Moreover, an MC may aid in reasonably and efficiently selecting materials for broomcorn millet breeding as researchers could screen for aimed genetic characters within a smaller scope. Full article
(This article belongs to the Section Crop Breeding and Genetics)
Show Figures

Figure 1

Figure 1
<p>UPGMA dendrogram of the 17 populations in the broomcorn millet core collection based on 34 SSR markers.</p>
Full article ">Figure 2
<p>The factorial correspondence analysis (FCA) and its pattern of spatial distribution of genetic variability among 634 accessions of broomcorn millet core collection.</p>
Full article ">Figure 3
<p>UPGMA dendrogram of the 634 accessions in the broomcorn millet core collection based on 34 SSR markers.</p>
Full article ">Figure 4
<p>Genetic parameter comparison of 34 SSR markers in core collection and min-core collection.</p>
Full article ">Figure 5
<p>Genetic diversity indices compare in core collection and mini-core collection.</p>
Full article ">
15 pages, 3106 KiB  
Article
Ancient Genome of Broomcorn Millet from Northwest China in Seventh Century CE: Shedding New Light to Its Origin and Dispersal Patterns
by Xiaolan Sun, Yifan Wang, Yongxiu Lu, Yongxiang Xu, Bingbing Liu, Yishi Yang, Guoke Chen, Hongru Wang, Zihao Huang, Yuanyang Cai, Zhengquan Gu, Xiaoxia Wang, Guanghui Dong and Yucheng Wang
Agronomy 2024, 14(9), 2004; https://doi.org/10.3390/agronomy14092004 - 2 Sep 2024
Viewed by 605
Abstract
Broomcorn millet (Panicum miliaceum) is among the earliest domesticated staple crops in the world’s agricultural history and facilitated the development of several early agrarian cultures, particularly those originating in northern China. However, the propagation route of broomcorn millet in China from [...] Read more.
Broomcorn millet (Panicum miliaceum) is among the earliest domesticated staple crops in the world’s agricultural history and facilitated the development of several early agrarian cultures, particularly those originating in northern China. However, the propagation route of broomcorn millet in China from the Middle Ages to the present remains unclear. The aim of this study is to explore the genetic affinity between ancient and modern millet samples, trace the genetic origins and diffusion pathways of broomcorn millet, and provide insights into its domestication and spread. To achieve this, we sequence ancient DNA from broomcorn millet remains excavated from the Chashan Village cemetery (AD 691) in Gansu Province, China. Phylogenetic and population genetic analyses, integrating ancient and modern millet genomes, reveal a close genetic relationship between ancient millet and contemporary millet from Ningxia Province (445 km away from Chashan Village), suggesting a potential origin for the Chashan millet. This finding aligns with the tomb’s epitaph, which documents the reburial of the tomb’s owner, who was originally buried in Ningxia, and provides important archaeological evidence for understanding the interaction between geopolitical dynamics and the natural environment in northwest China during the late seventh century. Furthermore, outgroup-f3 and D statistics evidence suggests substantial genetic interactions between ancient millet and modern varieties from the Loess Plateau, Huang-Huai-Hai Plain, and Northeast Plain, indicating the dispersal route of broomcorn millet, along with human migration routes, from the northwest to northern China and ultimately to the northeast region, starting from the Middle Ages onward. This study enhances our understanding of millet’s genetic history, offers a novel perspective on burial archaeology, and provides valuable insights into the origins, domestication, and diffusion of broomcorn millet. Full article
(This article belongs to the Section Farming Sustainability)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The 9 divisions of agricultural regions in China, including the Northeast China Plain (NE), Southwest Yunnan–Guizhou Plateau (SW), Northern arid and semiarid region (NA), Southern China (SC), Sichuan Basin and surrounding regions (SB), Middle–Lower Yangtze Plain (M), Qinghai–Tibet Plateau (QT), Loess Plateau (LP), and Huang-Huai-Hai Plain (HH). (<b>b</b>) The location of the Chashan Village burial site. (<b>c</b>) West side of the tomb chamber. (<b>d</b>) Ultra-microscopic photos of broomcorn millet plant remains. (<b>e</b>) Plant remains collected from all the torn silk bags.</p>
Full article ">Figure 2
<p>(<b>a</b>) Ancient and modern sample locations. (<b>b</b>) The NJ phylogenetic tree of all the samples based on the p-distance. (<b>c</b>,<b>d</b>) PCA clustering of all the samples.</p>
Full article ">Figure 3
<p>Population structure of ancient broomcorn millet and modern samples based on the admixture analysis.</p>
Full article ">Figure 4
<p>(<b>a</b>) Outgroup-f3 statistics for different millet populations. (<b>b</b>) D-statistic for different millet populations. All results that reach statistical significance are indicated by asterisks (corresponding |Z| &gt; 3).</p>
Full article ">Figure 5
<p>Historical patterns of human dispersal in China. Route 1 represents the southward migration of northern populations after the An Lushan Rebellion (~1300 BP). Route 2 traces the migration from the LP region to the HH region (~600 BP). Route 3 reflects the movement from the HH region to the NE region (~200 BP to modern times).</p>
Full article ">
17 pages, 8680 KiB  
Article
Effects of Rehydration on Bacterial Diversity in the Rhizosphere of Broomcorn Millet (Panicum miliaceum L.) after Drought Stress at the Flowering Stage
by Yuhan Liu, Jiao Mao, Yuanmeng Xu, Jiangling Ren, Mengyao Wang, Shu Wang, Sichen Liu, Ruiyun Wang, Lun Wang, Liwei Wang, Zhijun Qiao and Xiaoning Cao
Microorganisms 2024, 12(8), 1534; https://doi.org/10.3390/microorganisms12081534 - 26 Jul 2024
Viewed by 658
Abstract
This study aimed to elucidate responses of the bacterial structure and diversity of the rhizosphere in flowering broomcorn millet after rehydration following drought stress. In this study, the broomcorn millet varieties ‘Hequ red millet’ (A1) and ‘Yanshu No.10′ (A2), known for their different [...] Read more.
This study aimed to elucidate responses of the bacterial structure and diversity of the rhizosphere in flowering broomcorn millet after rehydration following drought stress. In this study, the broomcorn millet varieties ‘Hequ red millet’ (A1) and ‘Yanshu No.10′ (A2), known for their different drought tolerance levels, were selected as experimental materials. The plants were subjected to rehydration after drought stress at the flowering stage, while normal watering (A1CK and A2CK) served as the control. Soil samples were collected at 10 days (A11, A21, A1CK1, and A2CK1) and 20 days (A12, A22, A1CK2, and A2CK2) after rehydration. High-throughput sequencing technology was employed to investigate the variations in bacterial community structure, diversity, and metabolic functions in the rhizosphere of the broomcorn millet at different time points following rehydration. The findings indicated that the operational taxonomic units (OTUs) of bacteria in the rhizosphere of broomcorn millet were notably influenced by the duration of treatment, with a significant decrease in OTUs observed after 20 days of rehydration. However, bacterial Alpha diversity was not significantly impacted by rehydration following drought stress. The bacterial community in the rhizosphere of broomcorn millet was mainly composed of Actinobacteria and Proteobacteria. After rewatering for 10 to 20 days after drought stress, the abundance of Sphingomonas and Aeromicrobium in the rhizosphere soil of the two varieties of broomcorn millet decreased gradually. Compared with Yanshu No.10, the abundance of Pseudarthrobacter in the rhizosphere of Hequ red millet gradually increased. A Beta diversity analysis revealed variations in the dissimilarities of the bacterial community which corresponded to different rehydration durations. The relative abundance of bacterial metabolic functions in the rhizosphere of broomcorn millet was lower after 20 days of rehydration, compared to measurements after 10 days of rehydration. This observation might be attributed to the exchange of materials between broomcorn millet and microorganisms during the initial rehydration stage to repair the effects of drought, as well as to the enrichment of numerous microorganisms to sustain the stability of the community structure. This study helps to comprehend the alterations to the bacterial structure and diversity in the rhizosphere of broomcorn millet following drought stress and rehydration. It sheds light on the growth status of broomcorn millet and its rhizosphere microorganisms under real environmental influences, thereby enhancing research on the drought tolerance mechanisms of broomcorn millet. Full article
Show Figures

Figure 1

Figure 1
<p>Heatmap of horizontal clustering of rhizobacteria. A12, A22: rewatering for 20 days after drought stress; A1CK1, A2CK1: no stress, rehydration for 10 days; A1CK2, A2CK2: rewatering for 20 days, and without stress.</p>
Full article ">Figure 2
<p>The first 50 cluster heatmaps of bacterial abundance in the rhizosphere of broomcorn millet. (<b>a</b>): class-level clustering heatmap of rhizosphere bacteria; (<b>b</b>): cluster heat map of rhizosphere bacteria at the order level; (<b>c</b>): family-level clustering heatmap of rhizosphere bacteria; (<b>d</b>): cluster heatmap of rhizosphere bacteria at the genus level. A11, A21: rewatering for 10 days after drought stress; A12, A22: rewatering for 20 days after drought stress; A1CK1, A2CK1: no stress, rehydration for 10 days; A1CK2, A2CK2: rewatering for 20 days, and without stress.</p>
Full article ">Figure 2 Cont.
<p>The first 50 cluster heatmaps of bacterial abundance in the rhizosphere of broomcorn millet. (<b>a</b>): class-level clustering heatmap of rhizosphere bacteria; (<b>b</b>): cluster heat map of rhizosphere bacteria at the order level; (<b>c</b>): family-level clustering heatmap of rhizosphere bacteria; (<b>d</b>): cluster heatmap of rhizosphere bacteria at the genus level. A11, A21: rewatering for 10 days after drought stress; A12, A22: rewatering for 20 days after drought stress; A1CK1, A2CK1: no stress, rehydration for 10 days; A1CK2, A2CK2: rewatering for 20 days, and without stress.</p>
Full article ">Figure 3
<p>The effect of rewatering after drought stress on the relative abundance of bacteria in the rhizosphere of broomcorn millet at the genus level. The formula for calculating the multiple change is as follows: (relative abundance under drought stress rewatering conditions/relative abundance under control conditions)-1. The error bars represent three independently repeated standard errors. Different lowercase letters indicated that the expression level was significantly different at the <span class="html-italic">p</span> &lt; 0.05 level. (<b>a</b>): <span class="html-italic">Pseudarthrobacter</span>; (<b>b</b>): <span class="html-italic">Streptomyces</span>; (<b>c</b>): <span class="html-italic">Lysobacter</span>; (<b>d</b>): <span class="html-italic">Roseiflexus</span>; (<b>e</b>): <span class="html-italic">Microvirga</span>.</p>
Full article ">Figure 4
<p>Classification analysis of the genus heatmap of the top 50 levels of abundance. (<b>a</b>,<b>b</b>): Heat map and cluster analysis of the comprehensive population composed of samples taken 10 days and 20 days after rehydration, respectively.</p>
Full article ">Figure 5
<p>Analysis of soil microbial Beta diversity. (<b>a</b>,<b>b</b>): Two-dimensional ordination of principal component analysis of rhizosphere microorganisms after 10 days and 20 days of rewatering; (<b>c</b>,<b>d</b>): PLS-discriminant analysis of graded microorganisms after 10 days and 20 days of rehydration. Identical groups of samples are marked with ellipses.</p>
Full article ">Figure 6
<p>Difference analysis of rhizosphere bacterial metabolic pathways.</p>
Full article ">
19 pages, 4922 KiB  
Article
Revealing the 2300-Year-Old Fermented Beverage in a Bronze Bottle from Shaanxi, China
by Li Liu, Yanglizheng Zhang, Wei Ge, Zhiwei Lin, Nasa Sinnott-Armstrong and Lu Yang
Fermentation 2024, 10(7), 365; https://doi.org/10.3390/fermentation10070365 - 18 Jul 2024
Viewed by 2276
Abstract
China has a 9000-year-long history of cereal-based alcohol production, with the use of molds (filamentous fungi) likely being one of the earliest fermentation techniques. This method later developed into the uniquely East Asian qu (koji) starter compound, containing grains, molds, yeasts, and bacteria. [...] Read more.
China has a 9000-year-long history of cereal-based alcohol production, with the use of molds (filamentous fungi) likely being one of the earliest fermentation techniques. This method later developed into the uniquely East Asian qu (koji) starter compound, containing grains, molds, yeasts, and bacteria. Recent studies have revealed that this method was already widely applied during the Neolithic period. However, much less is known about its development during the early dynastic times, and our knowledge of this innovation has mainly relied on textual materials. Here, we present direct evidence, based on microbotanical, microbial, and chemical analyses, for the fermentation method of a 2300-year-old liquid preserved in a sealed bronze bottle unearthed in a Qin tomb at Yancun, Shaanxi. The results of this research suggest that this liquid is likely a fermented beverage made from wheat/barley, rice, Job’s tears, broomcorn millet, and pulses. The fermentation starter may have been a cereal-based qu, consisting of a wide range of microorganisms, including molds (Aspergillus and Monascus), yeasts, and bacteria. Our findings suggest that the tradition of selecting suitable grains and microbial communities for brewing alcohol, possibly with a maiqu starter (primarily wheat/barley-based qu), may have been well established more than two thousand years ago. Full article
(This article belongs to the Special Issue Advances in Beverages, Food, Yeast and Brewing Research, 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Site locations and artifacts from tomb M41 discussed in this paper. (<b>A</b>) Locations of sites revealing alcohol remains. 1: Yancun; 2: Beibai’e; 3: Anyang (Liujiazhuang and Dasikong); 4: Qianzhangda; 5: Changzikou; and 6: Tianhu. (<b>B</b>) Tomb M41, showing grave goods in the left chamber; (<b>C</b>) the interior of the lid from the <span class="html-italic">hu</span> bottle, showing corroded surface and remaining textile used to help seal the vessel; (<b>D</b>) tightly lidded bronze <span class="html-italic">hu</span> bottle; (<b>E</b>) M41 plan; (<b>F</b>) liquid sample analyzed in this study; and (<b>G</b>) the <span class="html-italic">hu</span> bottle (middle) with other vessels excavated in the tomb (photographs of the site and artifacts were taken by Yanglizheng Zhang).</p>
Full article ">Figure 2
<p>Starch types from the Qin liquid. (<b>A</b>) Type I, Panicoideae, likely millet; (<b>B</b>) type I, Panicoideae, likely Job’s tears (arrow pointing to the zig-zag arm; (<b>C</b>) type II, pulses; (<b>D</b>) type III, Triticeae (A-type), showing deep channels; (<b>E</b>) type III, Triticeae (B-type), showing deep channels, a central depression, and pitting; (<b>F</b>) type IV, rice compound; (<b>G</b>) type IV, rice compound, partially missing; (<b>H</b>) mildly gelatinized starch masses, possibly rice; (<b>I</b>) UNID starch, showing hollowed center; (<b>J</b>) UNID starch, broken edges; (<b>K</b>) gelatinized starch mass, possibly rice; (<b>L</b>) gelatinized starch, resembling wheat/barley; and (<b>M</b>) UNID gelatinized starch. (<b>A</b>–<b>K</b>) Each image shown in bright field and polarized views).</p>
Full article ">Figure 3
<p>Fungi and bacteria in the Qin liquid. (<b>A</b>–<b>D</b>) Oval yeast cells in different stages of the budding process, cf. <span class="html-italic">S. cerevisiae</span>; (<b>E</b>) round yeast cell; (<b>F</b>) a cluster of oval yeasts; (<b>G</b>) elongated–oval yeast, cf. <span class="html-italic">Pichia</span> sp.; (<b>H</b>) rod-shaped yeast in budding; (<b>I</b>,<b>J</b>) rod-shaped bacteria in fission, cf. <span class="html-italic">Bacillum</span> sp.; (<b>K</b>) fungal spore in germination; (<b>L</b>) fungal hypha; (<b>M</b>) hypha associated with conidia; (<b>N</b>,<b>P</b>) conidial cluster connected with hyphae, resembling <span class="html-italic">A. niger</span>; (<b>O</b>) cleistothecium resembling <span class="html-italic">Monascus</span> sp., with a yeast cell on top; and (<b>Q</b>) mycelium attached to conidial mass, resembling <span class="html-italic">A. niger</span>.</p>
Full article ">Figure 4
<p>Modern yeasts and molds comparable to ancient fungal elements. (<b>A</b>) Oval and round yeast cells (wild <span class="html-italic">S. cerevisiae</span>) in Shimao millet beer; (<b>B</b>) elongated–oval yeast cells (wild <span class="html-italic">P. kudriavzevii</span>) in Shimao millet beer; (<b>C</b>) cultured <span class="html-italic">S. cerevisiae</span>; (<b>D</b>) cultured <span class="html-italic">P. kudriavzevii</span>; (<b>E</b>) rod-shaped yeast cell in <span class="html-italic">daqu</span>; (<b>F</b>) bacteria, cultured <span class="html-italic">Bacillum</span> sp.; (<b>G</b>) <span class="html-italic">Aspergillus oryza</span>, mycelium; (<b>H</b>) vesicle with conidia, <span class="html-italic">A. oryzae</span>; (<b>I</b>) mycelia and vesicles, <span class="html-italic">A. oryzae</span>; (<b>J</b>) vesicle with conidial, <span class="html-italic">A. niger</span>; (<b>K</b>) hypha with conidia, <span class="html-italic">A. niger</span>; (<b>L</b>) mycelia and conidia clusters, <span class="html-italic">A. niger</span>; (<b>M</b>) wheat starch and mycelium in millet beer, fermented for 12 days; (<b>N</b>) cliestothecium, <span class="html-italic">Monascus</span> sp.; (<b>O</b>) structure of <span class="html-italic">Aspergillus</span>; and (<b>P</b>) structure of <span class="html-italic">Monascus</span> cliestothecium.</p>
Full article ">Figure 5
<p>FTICR MS analysis showing the positive-ion mode of the Qin liquid sample, indicating the presence of miliacin.</p>
Full article ">Figure 6
<p>Comparison of major microfossil elements recovered from the Qin liquid.</p>
Full article ">Figure 7
<p>Schematic process of brewing fermented beverages. (<b>A</b>) Making <span class="html-italic">qu</span> starter compound based on the findings of this study. (<b>B</b>) An Eastern Han portrait brick illustrating: 1. working on the brewing material containing <span class="html-italic">qu</span> and steamed cereals; 2. filtering fermented liquid through funnels into globular jars for storage; and 3. transporting alcoholic beverages by humans (artifact in Sichuan Museum). (<b>C</b>,<b>D</b>) Fermentation vat and storage jars from the Danxi Brewing Company in Zhejiang (photos by Li Liu).</p>
Full article ">
13 pages, 3710 KiB  
Article
Effects of Long-Term Fertilizer Application on Crop Yield Stability and Water Use Efficiency in Diversified Planting Systems
by Nana Li, Tao Li, Jianfu Xue, Gaimei Liang and Xuefang Huang
Agronomy 2024, 14(5), 1007; https://doi.org/10.3390/agronomy14051007 - 10 May 2024
Cited by 1 | Viewed by 981
Abstract
Exploring crop yield stability and the relationship between the water–fertilizer effect and annual precipitation type in a broomcorn millet–potato–spring corn rotation system under long-term fertilization on chestnut cinnamon soil in loess tableland can provide a scientific basis for rational fertilization in the northwest [...] Read more.
Exploring crop yield stability and the relationship between the water–fertilizer effect and annual precipitation type in a broomcorn millet–potato–spring corn rotation system under long-term fertilization on chestnut cinnamon soil in loess tableland can provide a scientific basis for rational fertilization in the northwest Shanxi region in years with different precipitation. This study was based on a 33-year long-term fertilizer experiment, using four fertilizer treatments: no fertilizer as control (CT), single fertilizer nitrogen (N), single organic fertilizer (M), and nitrogen fertilizer with organic fertilizer (NM). The results showed that broomcorn millet and maize had the highest yield in wet years, while potatoes had the highest yield in normal years and the yield under NM treatment was the highest. The sustainable yield index (SYI) values for potato and maize were higher than the SYI for the broomcorn millet during years with different precipitation and the SYI for the NM treatment was the highest. The water use efficiency of NM treatment was the highest. The yield of broomcorn millet and maize was affected by nitrogen fertilizer, organic fertilizer, and precipitation during the growth period, while the potato yield was mainly affected by nitrogen fertilizer and organic fertilizer. Therefore, the rotation of potato–maize and the rational allocation of organic and inorganic fertilizer (NM) is the best planting system in this region. Full article
Show Figures

Figure 1

Figure 1
<p>Water consumption during crop growth periods under different fertilization treatments in years with different precipitation. Note: different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Farmland precipitation utilization rates under different fertilization treatments in different precipitation years. Note: different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Crop water use efficiency under different fertilization treatments in years with different precipitation. Note: different lowercase letters indicate significant differences between treatments (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
14 pages, 7704 KiB  
Article
Alteration in Plant-Based Subsistence and Its Influencing Factors from Late Neolithic to Historical Periods in Hexi Corridor, Northwestern China: Archaeobotanical Evidence
by Wenyu Wei, Zhilin Shi, Yongxiu Lu, Linyao Du, Junmin Zhang, Guomu Zheng and Minmin Ma
Land 2024, 13(4), 419; https://doi.org/10.3390/land13040419 - 26 Mar 2024
Cited by 1 | Viewed by 1030
Abstract
The spatio-temporal evolution of human subsistence strategies and their driving force in prehistoric Eurasia has received increasing attention with the rapid development of archaeobotanical, zooarchaeological, and isotopic research in recent decades, while studies focusing on the historical periods are relatively absent. In the [...] Read more.
The spatio-temporal evolution of human subsistence strategies and their driving force in prehistoric Eurasia has received increasing attention with the rapid development of archaeobotanical, zooarchaeological, and isotopic research in recent decades, while studies focusing on the historical periods are relatively absent. In the Hexi Corridor in northwestern China, which has served as a hub for trans-Eurasian exchange since the late prehistoric period, archaeobotanical data have been reported from numerous Neolithic and Bronze Age sites, as well as sites from the Wei and Jin Dynasties (220–420 BCE) to the Yuan Dynasty (1271–1368 BCE). However, no archaeobotanical study has been conducted at sites of the Han Dynasty (202 BCE–220 CE), a crucial period connecting prehistoric and historical eras. In this study, we identified 32475 plant remains, including 31,463 broomcorn millets, 233 foxtail millets, and 780 weeds, from the Shuangdun North Beacon Tower (SDNBT) site of the Han Dynasty at the western end of the Hexi Corridor, suggesting that millets played a prominent part in human subsistence strategies in the area during this period. In addition, sheep, chicken, dog, horse, and rodent remains were also collected at the site. By applying a multi-disciplinary approach, we detected a remarkable change in plant-based subsistence in the ancient Hexi Corridor. Specifically, the importance of millet crops, compared with other crops (especially barley and wheat), in plant-based subsistence declined from the Late Neolithic to the Bronze Age; it apparently improved during the Han and Sui-Tang Dynasties (581–907 CE), when agricultural empires controlled the area, and then declined again during the Wei, Jin, Northern, and Southern Dynasties (220–581 CE) and the Song-Yuan Dynasty (960–1368 CE), when nomadic regimes controlled the area. Climate change, trans-Eurasian exchanges, and geopolitical shifts influenced the diachronic change in plant-based subsistence from the Late Neolithic to the historical periods in the Hexi Corridor. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of excavated and investigated sites in the Hexi Corridor from the Late Neolithic to the historical periods.</p>
Full article ">Figure 2
<p>Details of SDNBT site: (<b>a</b>) SDNBT site before excavation. (<b>b</b>) SDNBT site after excavation. (<b>c</b>,<b>d</b>) Dwelling remains at SDNBT site. (<b>e</b>,<b>f</b>) Archaeological remains from dwelling remains.</p>
Full article ">Figure 3
<p>Identified plant seeds from the SDNBT site (scale bar: 1 mm): (<b>a</b>) <span class="html-italic">Setaria italica</span>; (<b>b</b>) <span class="html-italic">Panicum miliaceum</span>; (<b>c</b>) <span class="html-italic">Cannabis sativa</span> L.; (<b>d</b>) <span class="html-italic">Setaria viridis</span>; (<b>e</b>) <span class="html-italic">Setaria pumila</span>; (<b>f</b>) <span class="html-italic">Echinochloa crusgalli</span>; (<b>g</b>) <span class="html-italic">Atriplex patens</span>; (<b>h</b>) <span class="html-italic">Erodium stephanianum</span> Willd.</p>
Full article ">Figure 4
<p>The alteration in plant-based subsistence in the Hexi Corridor from the Late Neolithic to historical periods was compared with climate records and δ<sup>13</sup>C values of human, pig, dog, and crop remains from sites in the study area: (<b>a</b>) Northern Hemisphere (30° to 90° N) temperature records compared with 1961–1990 instrumental mean temperatures [<a href="#B59-land-13-00419" class="html-bibr">59</a>]. (<b>b</b>) Reconstructed precipitation from DLH (Delingha) [<a href="#B60-land-13-00419" class="html-bibr">60</a>]. (<b>c</b>) δ<sup>13</sup>C values of the human, pig, and dog remains from sites in the Hexi Corridor. <span class="html-italic">n</span>: the number of isotopic data. (<b>d</b>) Weight percentages of the crop remains from sites in the Hexi Corridor [<a href="#B15-land-13-00419" class="html-bibr">15</a>,<a href="#B34-land-13-00419" class="html-bibr">34</a>]. (<b>e</b>) The archaeological framework of cultural evolution in the Hexi Corridor. Gray inverted triangles from left to right: Han government regained Hexi, Tang government regained Hexi, Xixia government occupied Hexi, and Mongolia government occupied Hexi.</p>
Full article ">
21 pages, 5399 KiB  
Article
Metabolomics and Physiological Methods Revealed the Effects of Drought Stress on the Quality of Broomcorn Millet during the Flowering Stage
by Jiangling Ren, Yuhan Liu, Jiao Mao, Yuanmeng Xu, Mengyao Wang, Yulu Hu, Shu Wang, Sichen Liu, Zhijun Qiao and Xiaoning Cao
Agronomy 2024, 14(2), 236; https://doi.org/10.3390/agronomy14020236 - 23 Jan 2024
Cited by 1 | Viewed by 1413
Abstract
The flowering stage is a critical period for water sensitivity and quality formation of broomcorn millets. However, the effects and mechanisms of drought stress on the quality formation of broomcorn millets are not clear. We used the drought-resistant variety Hequ red millet (H) [...] Read more.
The flowering stage is a critical period for water sensitivity and quality formation of broomcorn millets. However, the effects and mechanisms of drought stress on the quality formation of broomcorn millets are not clear. We used the drought-resistant variety Hequ red millet (H) and the drought-sensitive variety Yanshu No. 10 (Y) were used as materials for drought stress treatment during the flowering stage, metabolomics and physiological methods were used to study the differences in protein, starch, amino acids, medium and medium-long chain fatty acids, and their response characteristics to drought in broomcorn millet. The results showed that different genotypes of broomcorn millets exhibited different response mechanisms in the face of drought stress. In Hequ red millet, drought stress significantly increased the contents of amylopectin (2.57%), pyridoxine (31.89%), and anthocyanin, and significantly decreased the contents of water-soluble protein (5.82%), glutelin (10.07%), thiamine (14.95%) and nicotinamide (23.01%). In Yanshu No. 10, drought significantly decreased amylose by 6.05%, and significantly increased riboflavin and nicotinamide contents by 21.11% and 32.59%. Correlation analysis showed that total starch and amylose were highly significantly positively correlated with methyl palmitate; negatively correlated with amylopectin, vitamins, proteins, free amino acids, and medium-long chain fatty acids; and amylopectin was significantly positively correlated with water-soluble protein, riboflavin, and pyridoxine. Water-soluble protein and glutelin were significantly positively correlated with most free amino acids, medium-long chain fatty acids, and nicotinamide. Thiamine showed significant positive correlation with nicotinamide and significant negative correlation with pyridoxine. Riboflavin was significantly positively correlated with nicotinamide, pyridoxine, and water-soluble protein, and pyridoxine was significantly positively correlated with water-soluble protein. Hequ red millet transforms into amylopectin by consuming water-soluble protein and glutelin, and improves drought resistance by accumulating pyridoxine, and changes its physicochemical properties by decreasing the content of amylose and protein and elevating the content of amylopectin. Yanshu No. 10 resisted drought by catabolizing lipids to produce fatty acids and by consuming amylose for conversion into other metabolites. The present study helps to understand the response of the nutritional quality of millets to drought stress at the flowering stage and provides a theoretical basis for the selection and breeding of superior varieties of millets and drought resistance research. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of drought stress on starch content of broomcorn millet: (<b>a</b>) Total starch content; (<b>b</b>) Amylose content; (<b>c</b>) Amylopectin content. H: Hequ red millet; Y: Yanshu No. 10; CK: Control treatment; D: drought treatment, different lowercase letters indicated that the difference was significant at the <span class="html-italic">p</span> &lt; 0.05 level; * The difference was significant at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 2
<p>Effects of drought stress on protein composition of broomcorn millet: (<b>a</b>) Water-soluble protein content. (<b>b</b>) Globulin content. (<b>c</b>) Alcohol-soluble protein content. (<b>d</b>) Glutelin content. H: Hequ red millet; Y: Yanshu No. 10; CK: Control treatment; D: drought treatment, different lowercase letters indicated that the difference was significant at the <span class="html-italic">p</span> &lt; 0.05 level; <b>*</b> The difference was significant at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 3
<p>Effects of drought stress on contents of free amino acids and medium-long chain fatty acids in broomcorn millet: (<b>a</b>) Free amino acid content and (<b>b</b>) Medium-long chain fatty acid content. H: Hequ red millet; Y: Yanshu No. 10; CK: Control treatment; D: Drought treatment.</p>
Full article ">Figure 4
<p>Effects of drought stress on vitamin content of broomcorn millet: (<b>a</b>) Thiamine content. (<b>b</b>) Riboflavin content, (<b>c</b>) Nicotinamide content, and (<b>d</b>) Pyridoxine content. H: Hequ red millet; Y: Yanshu No. 10; CK: Control treatment; D: Drought treatment, different lowercase letters indicated that the difference was significant at the <span class="html-italic">p</span> &lt; 0.05 level; * The difference was significant at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 5
<p>Correlation heat map of broomcorn millet quality. Red indicates positive correlation, blue indicates negative correlation, color from deep to light and dot from large to small indicates correlation from strong to weak.</p>
Full article ">Figure 6
<p>(<b>a</b>) Principal component analysis (PCA) score plot. Different colors represent different treatments. The abscissa and ordinate represent the scores of PC1 and PC2, respectively, H: Hequ red millet; Y: Yanshu No. 10; CK: Control treatment; D: Drought treatment. (<b>b</b>) Hierarchical clustering heat map of relative abundance of two broomcorn millet quality. Different colors indicate that the abundance of quality accumulation in different samples is different. The abscissa represents the sample name, and the ordinate represents the quality index. H: Hequ red millet; Y: Yanshu No. 10; CK: Control treatment; D: Drought treatment.</p>
Full article ">
16 pages, 3751 KiB  
Article
Effects of Drought Stress during the Flowering Period on the Rhizosphere Fungal Diversity of Broomcorn Millet (Panicum miliaceum L.)
by Yuhan Liu, Jiangling Ren, Yulu Hu, Shu Wang, Jiao Mao, Yuanmeng Xu, Mengyao Wang, Sichen Liu, Zhijun Qiao and Xiaoning Cao
Agronomy 2023, 13(12), 2896; https://doi.org/10.3390/agronomy13122896 - 25 Nov 2023
Cited by 2 | Viewed by 1153
Abstract
Drought stress restricts plant growth and development. The flowering stage is a period of abundant microbial diversity in the rhizosphere of broomcorn millet. However, the diversity and community structure of rhizosphere fungi during the flowering stage of broomcorn millet and the response mechanism [...] Read more.
Drought stress restricts plant growth and development. The flowering stage is a period of abundant microbial diversity in the rhizosphere of broomcorn millet. However, the diversity and community structure of rhizosphere fungi during the flowering stage of broomcorn millet and the response mechanism to drought stress are still unclear. In this study, two broomcorn millet varieties, ‘Hequ red millet’ (A1) and ‘Yanshu No.10′ (A2), with different drought resistances, were used as experimental materials. Using the pot water control method, drought treatment at the flowering stage was carried out, and normal watering (A1CK, A2CK) was used as the control. High-throughput sequencing technology was used to study the diversity and structural changes in rhizosphere fungi in broomcorn millet. The results show that the number of fungi OTUs in the A1, A2, A1CK and A2CK samples were 445, 481, 467 and 434, respectively, of which 331 OTUs were shared by all groups. The fungal community in the rhizosphere of broomcorn millet was mainly composed of Ascomycota and Basidiomycota. Drought treatment significantly reduced the abundance of Mortierella and significantly increased the abundance of Phoma. The abundance of Nectriaceae in the rhizosphere soil of ‘Hequ Red millet’ was significantly increased. The abundance of Pseudocercospora in the rhizosphere soil of ‘Yanshu No.10′ was higher, and the lower was Hypocreales and Nectriaceae. However, there was no significant difference in the alpha diversity of fungal communities in the four treatments, and the fungal community structure between A2 and A1CK was more similar, whereas the difference between A1 and A2CK was larger. Correlation analysis showed that drought stress had little effect on the interaction of rhizosphere fungi, and metabolic functions such as nucleotide metabolism and electron transport in rhizosphere fungi accounted for a relatively high proportion. The results show that the diversity and community structure of rhizosphere fungi were less affected by drought, which may have been due to the close interaction between species, which made the fungal community more stable under drought stress, and the difference in planting varieties may have affected the enriched rhizosphere fungal species. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of drought stress on agronomic traits of broomcorn millet. (<b>a</b>) Plant height; (<b>b</b>) Number of internode; (<b>c</b>) Culm diameter; (<b>d</b>) Panicle length; (<b>e</b>) Dry weight of panicle; (<b>f</b>) Dry grain weight per plant.</p>
Full article ">Figure 2
<p>OTUs of each sample. (<b>a</b>) Venn diagram of the number of sample OTUs; (<b>b</b>) The number of sample OTUs at different classification levels.</p>
Full article ">Figure 3
<p>Diversity analysis of rhizosphere fungi in drought-tolerant and weak drought-tolerant broomcorn millet. (<b>a</b>) Rarefaction curves; (<b>b</b>): Species accumulation curves; (<b>c</b>): Rank abundance curves.</p>
Full article ">Figure 4
<p>Distribution and abundance of taxa. (<b>a</b>–<b>e</b>) The percentage of taxa at the phylum, order, class, genus and species levels.</p>
Full article ">Figure 5
<p>Taxonomic analysis through a phylogenetic tree and heat map. (<b>a</b>) Classification hierarchy tree showing the hierarchical relationships of all taxa from the phylum to the genus level in the sample population. (<b>b</b>) Combined heat level map of the community composition with cluster analysis.</p>
Full article ">Figure 6
<p>Beta diversity analysis. (<b>a</b>) Two-dimensional ranking of the PCA analysis. (<b>b</b>) Weighted UniFrac distance matrix. Samples are clustered according to their similarity. The shorter the branch length between samples, the more similar the two samples are. (<b>c</b>) Multiple sets of box plots for the weighted UniFrac distance. (<b>d</b>) PLS-discriminant analysis. Each point represents a sample, the same color points belong to the same group, and the same group points are marked by ellipses. The samples belonging to the same group are closer to each other, and the distances between the points of the different groups are farther apart.</p>
Full article ">Figure 7
<p>Species-related network diagram of the top 50 species at each treatment level. In the figure, the default shows the species with <span class="html-italic">p</span> &lt; 0.01. The size of the nodes in the graph represents the abundance of the species, and different colors represent different species. Line color represents positive correlations and negative correlations: yellow represents a positive correlation, and gray represents a negative correlation. The thickness of the line indicates the correlation size. With the increase of line thickness, the correlation between species increased. More spectral lines indicate a closer relationship between species and other species. (<b>a</b>): A1; (<b>b</b>): A2; (<b>c</b>): A1CK; (<b>d</b>): A2CK.</p>
Full article ">Figure 8
<p>Difference analysis of rhizosphere microbial metabolic pathways.</p>
Full article ">
15 pages, 1608 KiB  
Article
A Brief History of Broomcorn Millet Cultivation in Lithuania
by Giedrė Motuzaitė Matuzevičiūtė and Rimvydas Laužikas
Agronomy 2023, 13(8), 2171; https://doi.org/10.3390/agronomy13082171 - 18 Aug 2023
Cited by 3 | Viewed by 2506
Abstract
The eastern Baltic region represents the world’s most northerly limit of successful broomcorn millet (Panicum miliaceum) (hereafter, millet) cultivation in the past, yet this crop has been almost forgotten today. The earliest millet in the eastern Baltic region has been identified [...] Read more.
The eastern Baltic region represents the world’s most northerly limit of successful broomcorn millet (Panicum miliaceum) (hereafter, millet) cultivation in the past, yet this crop has been almost forgotten today. The earliest millet in the eastern Baltic region has been identified from macrobotanical remains which were directly dated to ca 1000 BCE. Between 800 and 500 BCE, millet was one of the major staple foods in the territory of modern-day Lithuania. Millet continued to play an important role in past agriculture up until the 15th century, with its use significantly declining during the following centuries. This paper analyses both the archaeobotanical records and written sources on broomcorn millet cultivation in Lithuania from its first arrival all the way through to the 19th century. The manuscript reviews the evidence of millet cultivation in the past as documented by archaeobotanical remains and historical accounts. In light of fluctuating records of millet cultivation through time, we present the hypothetical reasons for the decline in millet use as human food. The paper hypothesizes that the significant decrease in broomcorn millet cultivation in Lithuania from the 15th century onwards was likely influenced by several factors, which include climate change (the Little Ice Age) and the agricultural reforms of the 16th century. However, more detailed research is required to link past fluctuations in millet cultivation with climatic and historical sources, thus better understanding the roots of collapsing crop biodiversity in the past. Full article
Show Figures

Figure 1

Figure 1
<p>The locations of millet discoveries in Lithuania categorized according to the chronology of their cultivation. The site locations corresponding to numbers on the map are listed in <a href="#agronomy-13-02171-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>The directly dated millet caryopses and fruits from multiple sites across Lithuania. Top row: Dzūkų, Grikapėdžio, and Panemuninkai; middle row: Liejyklos g. 8 Vilnius, Stumbragalvė, and Vingrėnai; bottom row: Pylimo 7, Turlojiškė after [<a href="#B57-agronomy-13-02171" class="html-bibr">57</a>], p. 56, and Tarbiskės after [<a href="#B49-agronomy-13-02171" class="html-bibr">49</a>], p. 166. Scale bar: 1 mm.</p>
Full article ">
18 pages, 8333 KiB  
Article
Biochar and Nitrogen Fertilizer Change the Quality of Waxy and Non-Waxy Broomcorn Millet (Panicum miliaceum L.) Starch
by Miaomiao Zhang, Bauyrzhan Mukhamed, Qinghua Yang, Yan Luo, Lixin Tian, Yuhao Yuan, Yani Huang and Baili Feng
Foods 2023, 12(16), 3009; https://doi.org/10.3390/foods12163009 - 9 Aug 2023
Cited by 1 | Viewed by 1282
Abstract
The overuse of nitrogen fertilizers has led to environmental pollution, which has prompted the widespread adoption of biochar as a soil conditioner in agricultural production. To date, there has been a lack of research on the effects of biochar and its combination with [...] Read more.
The overuse of nitrogen fertilizers has led to environmental pollution, which has prompted the widespread adoption of biochar as a soil conditioner in agricultural production. To date, there has been a lack of research on the effects of biochar and its combination with nitrogen fertilizer on the quality of broomcorn millet (Panicum miliaceum L.) starch. Thus, this study examined the physicochemical characteristics of starch in two types of broomcorn millet (waxy and non-waxy) under four different conditions, including a control group (N0), nitrogen fertilizer treatment alone (N150), biochar treatment alone (N0+B), and a combination of biochar and nitrogen fertilizer treatments (N150+B). The results showed that, in comparison to the control, all the treatments, particularly N150+B, decreased the content of amylose and gelatinization temperature and enhanced the starch transparency gel consistency and swelling power. In addition, biochar can improve the water solubility of starch and the gelatinization enthalpy. Importantly, the combination of biochar and nitrogen fertilizer increased the proportion of A-granules, final viscosity, starch content, and the average degree of amylopectin in polymerization. Thus, this research indicates that the combinations of biochar and nitrogen fertilizer result in the most significant improvement in the quality of starch produced from broomcorn millet. Full article
(This article belongs to the Section Grain)
Show Figures

Figure 1

Figure 1
<p>The morphologies of starch granules under normal light microscopy (NLM), polarized light microscopy (PLM), and scanning electron microscopy (SEM). N0, control; N150, nitrogen fertilizer; N0+B, biochar; N150+B, nitrogen fertilizer and biochar.</p>
Full article ">Figure 2
<p>The distribution of starch granule sizes. (<b>A</b>) SM1; (<b>B</b>) SM2. N0, control; N150, nitrogen fertilizer; N0+B, biochar; N150+B, nitrogen fertilizer and biochar.</p>
Full article ">Figure 3
<p>The distributions of the AP chain length of amylopectin. (<b>A</b>) SM1; (<b>B</b>) SM2. N0, control; N150, nitrogen fertilizer; N0+B, biochar; N150+B, nitrogen fertilizer and biochar.</p>
Full article ">Figure 4
<p>Ordered structure (FTIR) of starch. (<b>A</b>) SM1; (<b>B</b>) SM2. FTIR, Fourier transfer infrared spectroscopy; N0, control; N150, nitrogen fertilizer; N0+B, biochar; N150+B, nitrogen fertilizer and biochar.</p>
Full article ">Figure 5
<p>(<b>A</b>) Water solubility of starch; (<b>B</b>) Swelling power of starch; (<b>C</b>) Light transmittance of starch; (<b>D</b>) Retrogradation curves of starch. Different letters indicate variability among the eight different treatments. N0, control; N150, nitrogen fertilizer; N0+B, biochar; N150+B, nitrogen fertilizer and biochar.</p>
Full article ">Figure 6
<p>Pearson’s correlation coefficients of the physicochemical properties on broomcorn millet starch. A-granules (&gt;15 μm), B-granules (5–15 μm), and C-granules (&lt;5 μm); S, amylose content; F1, 1045/1022; F2, 1022/955; L, Light transmittance; To, onset gelatinization temperature; Tp, peak gelatinization temperature; Tc, conclusion gelatinization temperature; ΔH, gelatinization enthalpy; PV, peak viscosity; HV, hot viscosity; BV, breakdown viscosity; FV, final viscosity; SV, setback viscosity PT, peak viscosity. Data are the means of three replicates.</p>
Full article ">
30 pages, 51010 KiB  
Article
Montane Ecoclines in Ancient Central Asia: A Preliminary Study of Agropastoral Economies in Juuku, Kyrgyzstan
by Claudia Chang, Sergei S. Ivanov, Robert N. Spengler, Basira Mir-Makhamad and Perry A. Tourtellotte
Land 2023, 12(7), 1406; https://doi.org/10.3390/land12071406 - 13 Jul 2023
Cited by 1 | Viewed by 1963
Abstract
In this paper, we use preliminary archaeological data spanning the Iron Age through Medieval periods (ca. 800 BCE to 1200 CE) in the Juuku Valley in Kyrgyzstan on the south side of Lake Issyk-Kul to model land use across vertical mountain zones. We [...] Read more.
In this paper, we use preliminary archaeological data spanning the Iron Age through Medieval periods (ca. 800 BCE to 1200 CE) in the Juuku Valley in Kyrgyzstan on the south side of Lake Issyk-Kul to model land use across vertical mountain zones. We have (1) established a radiometric chronology; (2) conducted test excavations of an Iron Age settlement at 2100 m asl and a Turkic period burial at 1934 m asl; (3) undertaken preliminary archaeobotanical research; and (4) performed pedestrian surveys. Archaeobotanical remains of wheat (Triticum aestivum), barley (Hordeum vulgare), broomcorn millet (Panicum milaceum), foxtail millet (Setaria italica), and legumes were recovered in very small quantities from both sites. We compare these preliminary archaeobotanical results with previously published data from Talgar Iron Age settlements on the north side of the Tian Shan Mountain range in Kazakhstan. A small assemblage of faunal remains found at the Turkic period kurgan and from a profile at the upland Iron Age settlement demonstrates the practice of herding sheep/goats, cattle, and horses in the Juuku Valley. The goal of this study was to test the hypothesis that pastoral transhumance and agropastoralism were interchangeable economic strategies used by peoples in the Iron Age through Medieval periods in mountain-river valleys between 600 m to 2100 m asl. These economic strategies combined the pasturing of sheep, goats, cattle, and horses with the cultivation of cereals in a system that was adapted to different vegetational zones along a vertical gradient. This paper is based on preliminary research using survey data and test excavations and initiates a long-term research study of four millennia of settlements that appear to have ranged from pastoral transhumance and combined mountain agriculture. Full article
(This article belongs to the Special Issue Modeling Land Use Change Using Historical and Archaeological Datasets)
Show Figures

Figure 1

Figure 1
<p>Locator map showing the Juuku Valley in northeast Kyrgyzstan and the Talgar Fan in Southeastern Kazakhstan. Juuku Valley and the Talgar Fan are 135 km apart.</p>
Full article ">Figure 2
<p>The model of vertical zonation and farming and herding strategies in high mountain areas [<a href="#B19-land-12-01406" class="html-bibr">19</a>] (p. 10).</p>
Full article ">Figure 3
<p>Map showing site locations on the Talgar Fan.</p>
Full article ">Figure 4
<p>Map showing the site locations in Juuku.</p>
Full article ">Figure 5
<p>A Google Earth image of the Juuku Valley showing the location of EJS1 (Iron Age settlement in the Eastern Juuku Branch and the location of LJK1 (Turkic period kurgan) in Lower Juuku valley.</p>
Full article ">Figure 6
<p>Oblique View of LJK 1, view to the north.</p>
Full article ">Figure 7
<p>EJS1 Excavation and Profile Cleaning, July 2022.</p>
Full article ">Figure 8
<p>Basira Mir-Makhamad and Malike Primidova taking flotation samples from ESJ1, Profile 6.</p>
Full article ">Figure 9
<p>Taking flotation samples from ESJ1, Profile 4: Robert Spengler, Sergei Ivanov, and Claudia Chang in background.</p>
Full article ">Figure 10
<p>Photograph of Profile 4 at ESJ1 showing stratigraphic layers: 1—sod layer, dark humic soil (0–22 cm below surface), 2—natural sand, clay, and silt wash levels, 3—wash layers of fine sand, 3a—a series of fine sand laminae where a mudbrick feature is located, 4—definite layer of sheet midden from site occupation, the dotted lines show a thick ashy layer where flotation samples were taken.</p>
Full article ">Figure 11
<p>EJS1 Profile 1 from 2021 field season. Note where the flotation samples and C-14 samples were taken.</p>
Full article ">Figure 12
<p>A plan drawing of LJK1.</p>
Full article ">Figure 13
<p>LJK 1: East–west and north–south profiles.</p>
Full article ">Figure 14
<p>LJK1: View of grave shaft and burial chamber. Bones of ankles and feet in their original position.</p>
Full article ">Figure 15
<p>Photograph of grinding stone from Lower Juuku Kurgan 1 surface.</p>
Full article ">Figure 16
<p>Plan view of block excavations at EJS1 at 68 cm to 210 cm below datum. The numbers are elevation readings (in cm) below datum. The plan view shows Floor 2—a hard-packed yellow clay surface, Wall—a small fragment of clay wall just above Floor 2, Wall Foundation—a melted mudbrick foundation, and Ash Pit 1—a circular pit filled with small pebbles and ashy soil on the surface of Floor 2; fill levels are loose red sandy deposits.</p>
Full article ">Figure 17
<p>Cultivated plants the burial mound in Juuku, (<b>a</b>)—barley and (<b>b</b>)—wheat.</p>
Full article ">Figure 18
<p>(<b>a</b>)—hulled barley, (<b>b</b>)—highly compact barley, (<b>c</b>)—wheat, (<b>d</b>)—culm node, (<b>e</b>)—foxtail millet, (<b>f</b>)—broomcorn millet, (<b>g</b>)—pea, and (<b>h</b>)—cf. grass pea.</p>
Full article ">Figure 19
<p>A scatter plot based on barley grain measurements from Chap I [<a href="#B6-land-12-01406" class="html-bibr">6</a>], Tuzusai [<a href="#B65-land-12-01406" class="html-bibr">65</a>], and EJS 1 (2019 and 2022 seasons).</p>
Full article ">Figure 20
<p>A scatter plot based on wheat grain measurements from Chap I 2021a [<a href="#B6-land-12-01406" class="html-bibr">6</a>], Tuzusai [<a href="#B65-land-12-01406" class="html-bibr">65</a>], and EJS1 (2019 and 2022 seasons).</p>
Full article ">
15 pages, 2939 KiB  
Article
Microbiota Composition during Fermentation of Broomcorn Millet Huangjiu and Their Effects on Flavor Quality
by Ke Wang, Huijun Wu, Jiaxuan Wang and Qing Ren
Foods 2023, 12(14), 2680; https://doi.org/10.3390/foods12142680 - 11 Jul 2023
Cited by 3 | Viewed by 1530
Abstract
Broomcorn millet Huangjiu brewing is usually divided into primary fermentation and post-fermentation. Microbial succession is the major factor influencing the development of the typical Huangjiu flavor. Here, we report the changes in flavor substances and microbial community during the primary fermentation of broomcorn [...] Read more.
Broomcorn millet Huangjiu brewing is usually divided into primary fermentation and post-fermentation. Microbial succession is the major factor influencing the development of the typical Huangjiu flavor. Here, we report the changes in flavor substances and microbial community during the primary fermentation of broomcorn millet Huangjiu. Results indicated that a total of 161 volatile flavor compounds were measured during primary fermentation, and estragole was detected for the first time in broomcorn millet Huangjiu. A total of 82 bacteria genera were identified. Pediococcus, Pantoea, and Weissella were the dominant genera. Saccharomyces and Rhizopus were dominant among the 30 fungal genera. Correlation analysis showed that 102 microorganisms were involved in major flavor substance production during primary fermentation, Lactobacillus, Photobacterium, Hyphodontia, Aquicella, Erysipelothrix, Idiomarina, Paraphaeosphaeria, and Sulfuritalea were most associated with flavoring substances. Four bacteria, Lactobacillus (R1), Photobacterium (R2), Idiomarina (R3), and Pediococcus (R4), were isolated and identified from wheat Qu, which were added to wine Qu to prepare four kinds of fortified Qu (QR1, QR2, QR3, QR4). QR1 and QR2 fermentation can enhance the quality of Huangjiu. This work reveals the correlation between microorganisms and volatile flavor compounds and is beneficial for regulating the micro-ecosystem and flavor of the broomcorn millet Huangjiu. Full article
(This article belongs to the Special Issue The Microbial Community and Its Functions in Fermented Foods)
Show Figures

Figure 1

Figure 1
<p>Relative abundance levels of bacterial phyla (<b>a</b>) and genera (<b>b</b>) during different fermentation stages of Huangjiu brewed from broomcorn millet.</p>
Full article ">Figure 2
<p>Bacteria correlation network analysis of Huangjiu brewed from broomcorn millet.</p>
Full article ">Figure 3
<p>Relative abundance levels of fungi phyla (<b>a</b>) and genera (<b>b</b>) during different fermentation stages of Huangjiu brewed from broomcorn millet.</p>
Full article ">Figure 4
<p>Fungi correlation network analysis of Huangjiu brewed from broomcorn millet.</p>
Full article ">Figure 5
<p>Correlation between microorganisms and flavor compounds in broomcorn millet.</p>
Full article ">Figure 6
<p>Comparison of bioamine content in wheat qu (control) and fortified qu (QR1, QR2, QR3, and QR4) fermentation. (<b>A</b>) Putrescine, (<b>B</b>) Cadaverine, (<b>C</b>) Histamine, (<b>D</b>) Tyramine, (<b>E</b>) Spermine, (<b>F</b>) Spermidine, (<b>G</b>) Phenylethylamine, (<b>H</b>) Total content. Different lowercase letters indicate significant differences in different groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
19 pages, 4267 KiB  
Article
Unveiling the Dynamics of Millet Spread into Xinjiang: New Evidence of the Timing, Pathways, and Cultural Background
by Duo Tian, Jingbo Li, Yongqiang Wang, Zhihao Dang, Xiangpeng Zhang, Chunchang Li and Youcheng Xu
Agronomy 2023, 13(7), 1802; https://doi.org/10.3390/agronomy13071802 - 6 Jul 2023
Cited by 4 | Viewed by 2176
Abstract
Xinjiang, in Northwestern China, was a key point in the prehistoric trans-Eurasian network of exchange and played an important role in facilitating the dispersal of crops across Eurasia. Millet crops were first cultivated and used ca. 10,000 years ago in Northern China, from [...] Read more.
Xinjiang, in Northwestern China, was a key point in the prehistoric trans-Eurasian network of exchange and played an important role in facilitating the dispersal of crops across Eurasia. Millet crops were first cultivated and used ca. 10,000 years ago in Northern China, from where they spread via different routes, leaving intriguing traces in various sites across Xinjiang. This paper presents the latest data on millet in Xinjiang. By employing a multidisciplinary approach, including radiocarbon dating, archaeobotanical evidence, and carbon isotope datasets, this study explores potential routes by which millet entered Xinjiang and traces its expansion from the third millennium BC to the 10th century AD. The research highlights the significant role of millet in shaping the ancient economies and cultures of Xinjiang and Central Asia, while also underscoring the importance of further investigation to uncover the complex pathways of its dispersal across Eurasia. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The distribution of the sites relevant to millet in Phase 1 and Phase 2 in Xinjiang: 1: Ayituohan; 2: Tongtiandong; 3: Xiaohe; 4: Xintala; 5: Tianshanbeilu; 6: Keriya North; 7: Saensayi; 8: Kalasu (Ili); 9: Goukou; 10: Adunqiaolu; 11: Liushugou; 12: Xiabandi. (<b>b</b>) Map of the study area. The red arrows I, II, and III represent the potential routes of millet expansion. Area A is the Hexi Corridor, where millet remains have been found that date from the sixth millennium BC to the third millennium BC. Area B is the Dzhungar Mountain region, which includes the millet-related sites of Dali, Begash, and Tasbas, the remains from which date to the third millennium BC. Area C is the Kashmir region, with the Pethpuran Teng site, where third-millennium-BC broomcorn millet was recovered. The base map was sourced from Natural Earth (<a href="https://www.naturalearthdata.com/" target="_blank">https://www.naturalearthdata.com/</a>, accessed on 2 May 2023).</p>
Full article ">Figure 2
<p>Examples of carbonized and desiccated broomcorn millet (<span class="html-italic">Panicum miliaceum</span>) recovered from sites in Xinjiang: (<b>a</b>) carbonized seed from Kuola, ventral side; (<b>b</b>) desiccated floret from Yanbulake, ventral and dorsal sides; (<b>c</b>) desiccated floret from Wupu, ventral and dorsal side; (<b>d</b>) carbonized seed from Kuiyukexiehai’er, ventral side; (<b>e</b>) desiccated leaf from Kalaya; (<b>f</b>) desiccated leaf from Wupu; (<b>g</b>) desiccated husks from Kuiyukexiehai’er; (<b>h</b>) desiccated florets from Kalaya; (<b>i</b>) desiccated panicle from Wupu; (<b>j</b>) desiccated florets from Lafuqueke. Scale bars: (<b>a</b>–<b>g</b>,<b>j</b>) 1 mm; (<b>h</b>) 5 mm; (<b>i</b>) 10 cm.</p>
Full article ">Figure 3
<p>Examples of carbonized and desiccated foxtail millet (<span class="html-italic">Setaria italica</span>) recovered from sites in Xinjiang: (<b>a</b>) carbonized seed from Wumachang, ventral side; (<b>b</b>) desiccated floret from Yanbulake, dorsal and ventral sides; (<b>c</b>) desiccated inflorescence from Wupu; (<b>d</b>) desiccated floret from Wupu, ventral and dorsal sides; (<b>e</b>) carbonized seeds from Kuiyukexiehaier, ventral side. Scale bars: (<b>a</b>,<b>b</b>,<b>d</b>,<b>e</b>) 1 mm; (<b>c</b>) 3 cm.</p>
Full article ">Figure 4
<p>Pairwise scatterplot matrix with regression lines and correlation coefficients for earliest inferred date (year), latitude (degree), longitude (degree), and altitude (meters) of millet.</p>
Full article ">Figure 5
<p>Distribution and percentage of millet sites in each phase. (<b>a</b>) Grouped violin plot with scatters and fitted lines showing the distribution of latitude (degree), longitude (degree), and altitude (meters). The violin plots show the probability density of the data. The smooth fitted lines display the trends of data from Phases 1 to 5. (<b>b</b>) Stacked bar plot showing the percentages of millet sites in mountain and oasis regions for each phase.</p>
Full article ">Figure 6
<p>Distribution of human stable carbon isotope values by phase, colored by altitude (in meters).</p>
Full article ">
17 pages, 3378 KiB  
Article
Pollen and Molecular Biomarkers from Sedimentary Archives in the Central Po Plain (N Italy): Assessing Their Potential to Deepen Changes in Natural and Agricultural Systems
by Assunta Florenzano, Eleonora Clò and Jérémy Jacob
Sustainability 2023, 15(13), 10408; https://doi.org/10.3390/su151310408 - 1 Jul 2023
Viewed by 1179
Abstract
This paper proposes to improve the information provided by biological indicators from sedimentary archives by integrating biomolecular techniques and botanical skills. This study represents a first proposal for combining pollen and biomolecular markers to detect land use and improve knowledge of past environmental [...] Read more.
This paper proposes to improve the information provided by biological indicators from sedimentary archives by integrating biomolecular techniques and botanical skills. This study represents a first proposal for combining pollen and biomolecular markers to detect land use and improve knowledge of past environmental change drivers. The specific aim of the research is to verify the relationship between miliacin (a pentacyclic triterpene methyl ether, usually interpreted as a broomcorn millet biomarker) and Panicum pollen in three near-site stratigraphic sequences of the Terramara S. Rosa di Poviglio (Po Plain, N Italy). The three cores span the last ~15,000 years and potentially record the beginning of Panicum miliaceum cultivation attested in the area since at least the Bronze Age within the Terramare culture. Despite the fact that Panicum pollen grains were rare in the spectra and miliacin was barely detectable in most of the 31 samples selected for biomolecular analyses, their combined evidence testifies to the local presence of the plant. Panicum pollen and sedimentary miliacin suggest the adoption of millet crops during the Recent Bronze Age by the Terramare culture, when climatic instability led to the diversification of cereal crops and the shift to drought-tolerant varieties. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the Terramara Santa Rosa di Poviglio (PVG) site (VP = “Villaggio Piccolo”— “Small Village”; VG = “Villaggio Grande”— “Large Village”) in the central Po Plain, N Italy, and position of the three cores (N-S3, C-S1, and F-S2) drilled in a SW-NE direction at ~150 m, ~320 m, and ~525 m, respectively, from the Terramara (GoogleEarth™).</p>
Full article ">Figure 2
<p>Near-site cores (N-S3, C-S1, and F-S2) of the Terramara S. Rosa di Poviglio: (<b>a</b>) stratigraphy and uncalibrated age BP marked by *; C = clay, Si = silt, Sa = sand, CSa = coarse sand; (<b>b</b>) AMS–<sup>14</sup>C ages for N-S3 and C-S1; the calibration into calendar years (2σ range) was calculated with R Clam using IntCal20 Northern Hemisphere <sup>14</sup>C calibration curves [<a href="#B47-sustainability-15-10408" class="html-bibr">47</a>].</p>
Full article ">Figure 3
<p>Poaceae pollen grains: (<b>a</b>) wild grass group (23 μm); (<b>b</b>) <span class="html-italic">Panicum</span> s.l. (43 μm) and details of its verrucate exine; (<b>c</b>) <span class="html-italic">Hordeum</span> group (40 μm); (<b>d</b>) <span class="html-italic">Avena/Triticum</span> group (51 μm); (<b>e</b>) <span class="html-italic">Zea mays</span> (101 μm). The scale is 10 µm.</p>
Full article ">Figure 4
<p>Partial chromatograms of the neutral fraction of the lipid extract of sample M04-S3 (113 cm). (<b>a</b>) Total Ion Current chromatogram showing the peak of friedelin (0); (<b>b</b>) <span class="html-italic">m/z</span> 425 + 440 ion-specific chromatogram revealing the peaks of pentacyclic triterpene methyl ethers (1—crusgallin; 2—miliacin); (<b>c</b>) <span class="html-italic">m/z</span> 453 + 468 ion-specific chromatogram illustrating the peaks of pentacyclic triterpenyl acetates (3—taraxeryl acetate; 4—β-amyrenyl acetate; 5—germanicyl acetate; 6—isobauerenyl acetate; 7—bauerenyl acetate; 8—isopichierenyl acetate; 9—pichierenyl acetate).</p>
Full article ">Figure 5
<p>Near-site cores (N-S3, C-S1, and F-S2) of the Terramara S. Rosa di Poviglio: concentration of miliacin (ng/g) and <span class="html-italic">Panicum</span> s.l. pollen percentages from the 31 samples studied for lipid biomarkers. The selected samples come from the Holocene levels; the orange bands mark the Bronze Age phase in each core.</p>
Full article ">Figure 6
<p>Near-site cores (N-S3, C-S1, and F-S2) of the Terramara S. Rosa di Poviglio: percentage pollen diagram of selected sums for palaeoenvironmental reconstruction (AP/NAP, hygrophilous herbs, API group [<a href="#B58-sustainability-15-10408" class="html-bibr">58</a>], wild and cultivated Poaceae). Enhanced curves × 10.</p>
Full article ">
Back to TopTop