ArticlePDF Available

Complete chloroplast genome and comparison of herbicides toxicity on Aeschynomene indica (Leguminosae) in upland direct-seeding paddy field

Authors:

Abstract and Figures

Background Indian jointvetch (Aeschynomene indica) is a common and pernicious weed found in the upland direct-seeding rice fields in the lower reaches of the Yangtze River in China. However, there are few reports on the degree of harm, genetic characteristics, and management methods of this weed. The purpose of this study is to clarify the harm of Indian jointvetch to upland direct-seeding rice, analyze the genetic characteristics of this weed based on chloroplast genomics and identify its related species, and screen herbicides that are effective in managing this weed in upland direct-seeding rice fields. Results In a field investigation in upland direct-seeding rice paddies in Shanghai and Jiangsu, we determined that the plant height and maximum lateral distance of Indian jointvetch reached approximately 134.2 cm and 57.9 cm, respectively. With Indian jointvetch present at a density of 4/m² and 8/m², the yield of rice decreased by approximately 50% and 70%, respectively. We further obtained the first assembly of the complete chloroplast (cp.) genome sequence of Indian jointvetch (163,613 bp). There were 161 simple sequence repeats, 166 long repeats, and 83 protein-encoding genes. The phylogenetic tree and inverted repeat region expansion and contraction analysis based on cp. genomes demonstrated that species with closer affinity to A. indica included Glycine soja, Glycine max, and Sesbania cannabina. Moreover, a total of 3281, 3840, and 3838 single nucleotide polymorphisms were detected in the coding sequence regions of the cp. genomes of S. cannabina voucher IBSC, G. soja, and G. max compared with the A. indica sequence, respectively. A greenhouse pot experiment indicated that two pre-emergence herbicides, saflufenacil and oxyfluorfen, and two post-emergence herbicides, florpyrauxifen-benzyl and penoxsulam, can more effectively manage Indian jointvetch than other common herbicides in paddy fields. The combination of these two types of herbicides is recommended for managing Indian jointvetch throughout the entire growth period of upland direct-seeding rice. Conclusions This study provides molecular resources for future research focusing on the identification of the infrageneric taxa, phylogenetic resolution, and biodiversity of Leguminosae plants, along with recommendations for reliable management methods to control Indian jointvetch.
This content is subject to copyright. Terms and conditions apply.
RESEARCH Open Access
© The Author(s) 2024. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use,
sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and
the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this
article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included
in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The
Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available
in this article, unless otherwise stated in a credit line to the data.
Gao et al. BMC Genomics (2024) 25:277
https://doi.org/10.1186/s12864-024-10102-x BMC Genomics
*Correspondence:
Guohui Shen
zb5@saas.sh.cn
Zhihui Tian
tianzhihui@saas.sh.cn
Full list of author information is available at the end of the article
Abstract
Background Indian jointvetch (Aeschynomene indica) is a common and pernicious weed found in the upland direct-
seeding rice elds in the lower reaches of the Yangtze River in China. However, there are few reports on the degree
of harm, genetic characteristics, and management methods of this weed. The purpose of this study is to clarify the
harm of Indian jointvetch to upland direct-seeding rice, analyze the genetic characteristics of this weed based on
chloroplast genomics and identify its related species, and screen herbicides that are eective in managing this weed
in upland direct-seeding rice elds.
Results In a eld investigation in upland direct-seeding rice paddies in Shanghai and Jiangsu, we determined
that the plant height and maximum lateral distance of Indian jointvetch reached approximately 134.2cm and
57.9cm, respectively. With Indian jointvetch present at a density of 4/m2 and 8/m2, the yield of rice decreased by
approximately 50% and 70%, respectively. We further obtained the rst assembly of the complete chloroplast (cp.)
genome sequence of Indian jointvetch (163,613bp). There were 161 simple sequence repeats, 166 long repeats, and
83 protein-encoding genes. The phylogenetic tree and inverted repeat region expansion and contraction analysis
based on cp. genomes demonstrated that species with closer anity to A. indica included Glycine soja, Glycine max,
and Sesbania cannabina. Moreover, a total of 3281, 3840, and 3838 single nucleotide polymorphisms were detected
in the coding sequence regions of the cp. genomes of S. cannabina voucher IBSC, G. soja, and G. max compared with
the A. indica sequence, respectively. A greenhouse pot experiment indicated that two pre-emergence herbicides,
saufenacil and oxyuorfen, and two post-emergence herbicides, orpyrauxifen-benzyl and penoxsulam, can more
eectively manage Indian jointvetch than other common herbicides in paddy elds. The combination of these two
types of herbicides is recommended for managing Indian jointvetch throughout the entire growth period of upland
direct-seeding rice.
Complete chloroplast genome
and comparison of herbicides toxicity
on Aeschynomene indica (Leguminosae)
in upland direct-seeding paddy eld
YuanGao1, TianYuChen2, JiaqiLong2, GuohuiShen1* and ZhihuiTian1*
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 2 of 13
Gao et al. BMC Genomics (2024) 25:277
Background
e genus Aeschynomene in the family Leguminosae is
widely distributed in tropical and subtropical regions
of the world, mostly in America, Africa, and Asia, with
approximately 200 species identied to date [13].
Aeschynomene is an annual herbaceous plant that grows
as shrubs, with shield-shaped stipules, buttery-shaped
owers, and segmented leaves connected by diaphragms
[4]. Only one species of this genus has been found in
China to date, namely A. indica (Indian jointvetch) [2]. It
usually germinates in April and its seeds mature in Octo-
ber. e germination of Indian jointvetch seeds mainly
depends on the temperature during the day [5]. Research
has shown that the seeds of Indian jointvetch have dor-
mancy characteristics, which can survive for about six
months in the soil seed bank and mechanical scarica-
tion is an eective method to break dormancy [6].
Although this plant does not germinate under ooded
conditions, once it grows, Indian jointvetch indirectly
aects crop productivity by competing for resources,
hindering harvesting operations, and reducing grain
and seed quality [79]. Moreover, it has been conrmed
that mixing seeds of A. indica into rice as pig feed can
induce the development of pig diseases [1]. A. indica was
reported to be a weed in irrigated rice elds previously
[10, 11] and was considered the third most troublesome
weed after weedy rice and barnyard grass in certain rice-
producing areas [12]. However, A. indica was recently
found to impair the growth of upland direct-seeding
rice in the lower reaches of the Yangtze River in China
(Additional File 3, Figure S1). Because the planting area
of upland direct-seeding rice is increasing owing to the
global shortage of labor and water resources, more and
more attention has been paid to the serious weed prob-
lem [1316]. Distinguishing plants of the genus Aeschyn-
omene is very dicult, and managing these weeds in rice
elds is also very challenging presently [3]. erefore,
further research is required to elucidate the characteris-
tics of this species, including the genetic characteristics,
degree of harm induced, and responses to available her-
bicides for eective management and control.
Chloroplasts are the main organelles involved in pho-
tosynthesis in photosynthetic plants or algae [17]. Chlo-
roplasts have their own distinct genetic material with
characteristics of non-recombination, haploid, and
single-parent inheritance, which leads to a highly con-
served genome; therefore, analysis of the chloroplast (cp.)
genome can provide very rich evolutionary information
[1820]. Additionally, the cp. genome is small and easy to
obtain, oering unique research value in phylogeny, spe-
cies identication, and population genetics [21]. e cp.
genome has a typical tetrad structure, with large single
copy (LSC) and small single copy (SSC) regions sepa-
rated by inverted repeats (IRs) [2224]. Previous studies
have shown that in the context of network evolution (i.e.,
hybridization) and polyploidy, analysis of the cp. genome
is particularly useful for characterizing the phylogenetic
and historical aspects of most plant lineages [2527].
However, information on the composition, structure,
interspecic dierences, and evolutionary relationships
of A. indica based on its cp. genome is limited. erefore,
exploring the cp. genomic information of A. indica can
provide a theoretical basis for the management or utiliza-
tion of this species and predict the possibility of evolu-
tion of related species into paddy weeds.
Current weed control methods in upland direct-seed-
ing rice elds mainly include cultural weed management
practices and physical methods such as deep plowing,
germination, mechanical weeding, and hand-pulling
weeding, along with chemical methods using herbicides
[2830]. Chemical methods for weed control in upland
direct-seeding rice elds have specic advantages of
obtaining higher yields with relatively lower labor costs
[3032]. Aeschynomene weeds are often controlled using
a combination of imazapyr and imazapic herbicides [3].
However, this genus of weed has been found to exhibit
imazapic resistance [33]. In addition, this herbicide is
typically considered to be unsafe for rice growth. To our
knowledge, there have been no systematic studies on
the toxicity of common herbicides used for A. indica in
rice elds. erefore, screening for suitable herbicides
to eectively control A. indica is of great signicance to
achieve a high and optimal yield of upland direct-seeding
rice.
To address these questions, we performed an in-depth
analysis of the genetic characteristics and herbicide
response of A. indica growing in the upland direct-seed-
ing rice elds of Shanghai, China. We studied the mor-
phological characteristics of this weed in rice elds and
its eect on rice yield. e cp. genome of A. indica was
then sequenced and compared to those of other species
to analyze their genetic characteristics and evolution-
ary relationships. Finally, the toxicities of common her-
bicides currently used in rice elds in China against A.
indica were evaluated in a greenhouse pot experiment
to determine a reasonable chemical control method. Our
Conclusions This study provides molecular resources for future research focusing on the identication of the
infrageneric taxa, phylogenetic resolution, and biodiversity of Leguminosae plants, along with recommendations for
reliable management methods to control Indian jointvetch.
Keywords Indian jointvetch, Chloroplast genome, Related species, Upland direct-seeding rice, Herbicide
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 3 of 13
Gao et al. BMC Genomics (2024) 25:277
study thus provides a theoretical and practical founda-
tion for gaining a better understanding of the damage of
the troublesome dicotyledonous weed A. indica imposes
in upland direct-seeding rice elds and facilitating strate-
gies for its eective management.
Methods
Plant materials
Indian jointvetch seeds were collected from upland
direct-seeding rice elds in Jinshan District, Shanghai,
China (N 30.81°, E 121.18°) in October 2021. e seeds
were dried to constant weight after collection and stored
at 4°C until planted.
Determination of eld traits
Plant height and the maximum lateral distance of Indian
jointvetch growing in upland direct-seeding rice elds in
Shanghai and Jiangsu Province were measured in Octo-
ber 2021. More than 20 plants were randomly selected;
however, no more than three plants were sampled from
each paddy eld. Seeds were collected, dried to a con-
stant weight, and weighed. A total of 294 upland direct-
seeding rice elds were investigated, elds with Indian
jointvetch were recorded, and their frequency was
calculated.
Subsequently, 0, 1, 2, 3, and 4 Indian jointvetch seed-
lings grown in a greenhouse previously were trans-
planted into a 1 m2 upland direct-seeding rice planting
area and all other weeds were manually removed. When
the upland direct-seeding rice had matured, the eective
spike number, grains per spike, and 1000-seed weight of
upland direct-seeding rice seeds were measured. Each
treatment contains three biological replicates and the
entire experiment was repeated twice. Data of percent-
ages of fresh weight compared to the control (none trans-
planted Indian jointvetch) were subjected to ANOVA.
For comparison of the dierences in the rice yield indica-
tors among the ve groups, the Duncan’s Multiple Range
Test (P < 0.05) was used. ANOVA was conducted using
SPSS version 20 (SPSS, Chicago, IL, USA).
Construction of the A. indica cp. genome
DNA sequencing and genome assembly
Total genomic DNA from ten Indian jointvetch at the
seedling stage was extracted using a modied cetyltri-
methylammonium bromide method and applied to con-
struct a 500-bp paired-end library using the NEBNext
Ultra DNA Library Prep Kit (New England Biolabs,
Ipswich, MA, USA) for next-generation sequencing on
the Illumina NovaSeq 6000 platform (Berry Genom-
ics Co., Ltd., Beijing, China). Approximately 2.35 Gb of
raw data from Indian jointvetch were generated with
150-bp paired-end read lengths. De novo assembly was
performed using NOVOPlasty v4.2 software (https://
github.com/ndierckx/NOVOPlasty, accessed March 24,
2023). Detailed sequencing and assembly methods are
described in our previous study [34].
Genome component analysis and gene annotation
Genes encoding proteins, tRNAs, and rRNAs in the cp.
genome of Indian jointvetch were predicted using GeSeq
software (https://chlorobox. mpimp-golm. mpg. de/
geseq.com ml/, accessed March 24, 2023). e protein
sequences of cp. genes were compared with known pro-
teins in databases using BLASTP (https://ncbiinsights.
ncbi.nlm.nih.gov/tag/blastp/, accessed March 24, 2023)
(e-value < 1 × 10 5). Because there may be more than one
alignment result for each sequence, only one optimal
alignment result was reserved for database alignment for
the given gene to ensure its biological signicance. e
amino acid sequences of Indian jointvetch were aligned
with sequences included in the Non-Redundant Pro-
tein Sequence (http://www.ncbi.nlm.nih.gov/, accessed
March 24, 2023), Swiss-Prot (http://www.ebi.ac.uk/uni-
prot, accessed March 24, 2023), Clusters of Orthologous
Groups, Kyoto Encyclopedia of Genes and Genomes
(KEGG; http://www.genome.jp/kegg/, accessed March
24, 2023), and Gene Ontology (GO; http://geneontology.
org/, accessed March 24, 2023) databases to obtain func-
tional annotation information for the coding genes.
Analysis of genetic relationships and identication
characteristics
Phylogenetic analysis
Nineteen cp. genomes of common plants, including the
model plant of dicotyledons Arabidopsis thaliana, com-
mon plants growing in rice elds (Oryza sativa Indica, O.
sativa Tropical Japonica, Echinochloa oryzoides, Eclipta
prostrata, Ammannia arenaria, Ammannia multiora,
Cyperus iria, Cyperus diormis), and common Legumi-
nosae plants (Vicia sepium, Sesbania cannabina, Medi-
cago sativa, Medicago truncatula, Medicago polymorpha,
Glycine max, Glycine soja, Astragalus sinicus), were
downloaded from the National Center for Biotechnology
Information (NCBI) database (see Additional File 1, Table
S1 for accession numbers) and subject to phylogenetic
analysis to determine their evolutionary relationships
with Indian jointvetch. e sequences were aligned using
ClustalW (v2.0.12) (http://www.clustal.org/clustal2/,
accessed April 4, 2023) with default settings. e DNA
substitution model was utilized based on the Akaike
information criterion [35]. e phylogenetic tree was
constructed by the maximum-likelihood (ML) method
using PhyML v3.0 (htp://ww.atgc-montpeller. fr/phyml/,
accessed April 4, 2023) and bootstrap values were calcu-
lated for 1000 replicates [36, 37]. e tree building model
was nally evaluated using jModelTest 2.1.10 (https://
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 4 of 13
Gao et al. BMC Genomics (2024) 25:277
github.com/ddarriba/jmodeltest2, accessed April 4,
2023), with the best model “GTR + I + G .”
Contraction and expansion analysis of IR regions
We performed IR contraction and expansion analy-
sis for the cp. genomes of Indian jointvetch and species
with closer genetic relationships selected based on the
phylogenetic analysis. e four quadripartite struc-
tures (LSC, SSC, and two IR repeat regions) of each cp.
genome were compared, and changes in the copy num-
ber of related genes caused by contraction and expansion
of the IR or pseudogenes, resulting in boundary regions,
were analyzed. Genes that crossed or were adjacent to
these boundaries were identied. In addition, the length
and distance from the boundaries of these genes were
analyzed.
Single nucleotide polymorphism (SNP) analysis
Using MUMmer software (http://mummer.sourceforge.
net/, accessed April 4, 2023), the cp. genome sequences
of Indian jointvetch (as the reference) and closely related
species, based on phylogenetic and contraction/expan-
sion IR analyses, were completely aligned, and prelimi-
nary ltering was performed to detect potential SNP
sites. Sequences of 100bp on both sides of the candidate
SNP site of the reference sequence were extracted and
aligned with the assembly results using BLAT v35 soft-
ware (http://hgdownload.soe.ucsc.edu/admin/exe/linux.
x86_64/blat/, accessed April 4, 2023) to verify the SNP
site. If the alignment length was less than 101bp, it was
considered an unreliable SNP and was removed; if the
alignment was repeated multiple times, the SNP was
considered a repetitive region and was removed, result-
ing in only reliable SNPs.
Herbicide sensitivity assays
We determined the susceptibility of Indian jointvetch
to common herbicides applied in rice elds using a
pot experiment in the greenhouse. e soil was a mid-
dle loam obtained from a farmland in the suburbs of
Shanghai, China, where herbicides had not been used.
Five seeds were sown in each plastic cup (7 × 7 × 7 cm),
water was added until saturation, a ne layer of soil was
added, and the seeds were treated with herbicides (i.e.,
pre-emergence treatment) using a 3WP-2000 walking-
type spray system (Nanjing Agricultural Mechanization
Research Institute of the Ministry of Agriculture, China).
Each treatment was performed with 30 mL of liquid
(450 L ha 1 water) using a fan-shaped nozzle. Control
plants were sprayed with the same amount of water as
the herbicide for the treated plants. e seedlings were
then placed in a greenhouse for cultivation. When the
seeds of the control group germinated and the treatment
group showed a signicant gradient, the Indian joint-
vetch plants were cut and weighed. Five seeds were sown
and cultivated at the three-leaf stage for post-emergence
herbicide application using the same method. After 21
days, the above-ground Indian jointvetch in each cup
was cut and weighed. e dose and details of the herbi-
cides are listed in Table1. e experiment included three
to four biological replicates and the entire experiment
was repeated twice. e eective rate of each herbicide
causing 50% reduction in plant growth (GR50) was deter-
mined using the four-parameter logistic function with
the “drc” add-on package [38] in the R 3.1.3 Language
and Environment for Statistical Computing [39]. is
model is dened as follows:
Y=c+{(dc)/(1 +exp(b(logx loge )) )}
where parameter e represents GR50 as the dose produc-
ing a response halfway between the upper limit d and the
lower limit c, and parameter b denotes the relative slope
around e.
Table 1 Information of the herbicides used in this study
Type Herbicides Manufacturer Dose (g a.i. ha− 1)
Pre-emergence Bensulfuron-methyl Zhejiang Tianyi Biotechnology Co.,Ltd., Shaoxing, Zhejiang, China 0, 5.625, 11.25, 22.5, 45, 90, 180
Butralin Shield Corporation, Fuzhou, Jiangxi, China 0, 135,270, 540, 1080, 2160, 4320
Oxyuorofen Yifan Biotechnology Group Co., Ltd., Wenzhou, Zhejiang, China 0, 11.25, 22.5, 45, 90, 180, 360
Saufenacil BASF SE, Ludwigshafen, Rheinland-Pfalz, Germany 0, 0.984375, 1.96875, 3.9375,
7.875, 15.75, 31.5
Post-emergence Pyrazosulfuron-ethyl Liben Crop Technology Co., Ltd., Liangyungang, Jiangsu, China 0, 0.9375, 1.875, 3.75, 7.5, 15, 30
2-methyl-4-chloro-
phenoxyacetic acid
Jiangsu Jian Gu Chemical Industry Co., Ltd., Suqian, Jiangsu Prov-
ince, China
0, 24.375, 48.75, 97.5, 195, 390,
780
Penoxsulam Corteva Agriscience, Wilmington, DE, USA 0, 0.9375, 1.875, 3.75, 7.5, 15, 30
Quinclorac Jiangsu Repont Agrochemical Co., Ltd., Changzhou, Jiangsu, China 0, 18.75, 37.5, 75, 150, 300, 600
Florpyrauxifen-benzyl Corteva Agriscience, Wilmington, DE, USA 0, 0.28125, 0.5625, 1.125, 2.25,
4.5, 9
Pyraquinate Shandong Cynda (Chemical) Co., Ltd., Jinan, Shandong, China 0, 4.6875, 9.375, 18.75, 37.5, 75,
150
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 5 of 13
Gao et al. BMC Genomics (2024) 25:277
Results
Impact of A. indica on rice growth
We rst investigated the morphological features of Indian
jointvetch during seed maturity in upland paddy elds in
the lower reaches of the Yangtze River in China. e aver-
age height of Indian jointvetch was 134.2 ± 3.7 cm, which
was higher than that of the rice (approximately 90cm).
e mean maximum lateral distance was 57.9 ± 2.1 cm.
e seeds presented a kidney shape (Additional File 3,
Figure S2) with a 1000-seed weight of 10.39 ± 0.055g. e
frequency of occurrence of this weed in upland paddy
elds in our survey was 13.85 ± 2.22%. From the perspec-
tive of rice damage, as the number of Indian jointvetch
plants increased, the eective spike number and grains
per spike of upland direct-seeding rice exhibited a sig-
nicant downward trend. Rice seed weights also showed
an overall downward trend with an increasing number of
Indian jointvetch plants. Specically, when the number of
Indian jointvetch was 0, 2, 4, 6, and 8, the yields of upland
direct-seeding rice were 8019.3, 5534.4, 3783.2, 3611.4,
and 2092.5kg/ha, respectively (Fig.1).
Characteristics of the A. indica cp genome
Cp genome map
After trimming low-quality fragments from the raw data,
33,624,994 clean reads with a GC content of 35.93% were
mapped to the complete cp. genome of Indian joint-
vetch. De novo assembly resulted in a circular genome of
163,613bp in length (Fig.2). e raw reads were depos-
ited in the NCBI GenBank database (accession number:
PRJNA963187). e complete cp. genome displayed the
typical quadripartite structure of most angiosperms,
including an LSC region, SSC region, and pair of IRs (IRa
and IRb). e lengths of the LSC, SSC, and IR regions
were 88,562, 19,801, and 27,625 bp, respectively, and
the intergenic region length was 86,150bp. e Indian
jointvetch cp. genome contains 83 protein-coding genes
(Table2).
Cp genome components
e cp. genome of Indian jointvetch contained 39 tRNA
genes and eight rRNA genes (Table3). ere were 70
protein-coding and 26 tRNA genes located within the
LSC; 10 protein-coding genes, 10 tRNA-coding genes,
and four rRNA-coding genes located within IRb or IRa;
and 13 protein-coding genes and one tRNA gene located
within the SSC region (Fig. 2). All 83 genes encoding
proteins in the cp. genome of Indian jointvetch were
functionally annotated, which mainly belonged to the
photosynthesis and self-replication categories. e gene
names, groups, and categories are listed in Table4. e
genes were mainly associated with GO biological pro-
cesses (Additional File 3, Figure S3), and were associ-
ated with KEGG energy production and conversion,
translocation, ribosomal structure, and biogenesis path-
ways (Additional File 3, Figure S4). In total, 161 simple
sequence repeats (SSRs) were identied in the Indian
jointvetch cp. genome. ere were eight SSRs on IRa or
IRb, 114 on the LSC region, and 31 on the SSC region.
In total, 166 long repeats (LRs) were identied in the cp.
genome of Indian jointvetch (Table5).
Genetic similarity analysis
Phylogenetic analysis
Bayesian inference of 19 complete cp. genomes showing
the same topology demonstrated that Indian jointvetch
clustered into a single clade (Fig.3). Indian jointvetch,
A. sinicus, M. polymorpha, G. soja, and G. max have the
most recent common ancestor (MRCA) (BS = 100 for the
ML tree), and this group has an MRCA with S. cannabina
Fig. 1 Eect of the number of Indian jointvetch plants on rice yield in the eld. The standard errors of the means are described by vertical bars. ANOVA
signicance groupings are shown as a, b, c, and d (P < 0.05)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 6 of 13
Gao et al. BMC Genomics (2024) 25:277
voucher IBSC voucher (BS = 100 for the ML tree). e
closest relatives to these plants were E. prostrata, O.
sativa Indica, O. sativa Tropical Japonica, and E. oryzi-
cola (BS = 100 for the ML tree). In addition, the group
formed by S. cannabina-1, M. sativa, M. polymorpha, A.
thaliana, C. iria, C. diormis, A. arenaria, and A. multi-
ora was closely related to the Indian jointvetch group.
e most distantly related plant to Indian jointvetch was
V. sepium.
IR expansion and contraction
To further observe the potential genetic relationships
among the most closely related species to Indian joint-
vetch, namely S. cannabina voucher IBSC, G. soja, and
G. max, based on the contraction and expansion of
IR regions, the gene variations at the IR/SSC and IR/
LSC boundary regions of the four species were com-
pared (Fig.4); A. sinicus and M. polymorpha do not have
complete IR regions and were therefore excluded from
this analysis. e genes rps19/ rpl2, trnN/ ndhF, ycf1/
trnN, and rpl2/ trnH were identied at the junctions of
Fig. 2 Assembly, size, and features of the chloroplast genome of Indian jointvetch. The genes outside the circle are transcribed in the counterclockwise
direction and the genes inside the circle are transcribed in the clockwise direction. Dierent colors of genes represent dierent functions. The dark gray
area and light gray area of the inner circle represent the GC and AT content of the genome, respectively
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 7 of 13
Gao et al. BMC Genomics (2024) 25:277
the LSC/IRb, IRb/SSC, SSC/IRa, and IRa/LSC regions,
respectively, in Indian jointvetch. Among them, ndhF
crosses IRb/SSC and ycf1 crosses SSC/IRa. Among all
tested species, the closest to Indian jointvetch was S. can-
nabina voucher IBSC with respect to the expansion and
contraction of IR regions; these two species exhibited
identical boundary genes with the only dierence being
the length of the genes or the distance between the genes
and the boundary.
SNP analysis
SNP analysis was performed to compare the dieren-
tiation among Indian jointvetch, S. cannabina voucher
IBSC, G. soja, and G. max. A total of 4659 SNPs were
detected in S. cannabina voucher IBSC, including 1378
SNPs in the intergenic spacer regions and 3281 SNPs in
the coding sequence regions (Table6; Additional File 2,
Table S2). e nonsynonymous to synonymous substitu-
tion ratio was 0.626. Two SNPs were detected at the start
codon and eight SNPs were detected at the stop codon.
e numbers of SNPs in the cp. genomes of G. soja and
G. max were much higher than that for S. cannabina
Table 2 Summary of chloroplast genome features in Indian
jointvetch
Genome Features Indian jointvetch
Genome size (bp) 163,613
LSC length (bp) 88,562
SSC length (bp) 19,801
IR length (bp) 27,625
Intergenic region Length (bp) 86,150
Overall GC content (%) 35.54
GC content of LSC (%) 32.77
GC content of SSC (%) 28.71
GC content of IR (%) 42.41
Number of Protein-coding genes 83
Table 3 Non-coding RNAs in the Indian jointvetch chloroplast
genome
Type ncRNA number Total length
(bp)
Average
length (bp)
Length/
Genome
(%)
tRNA 39 2943 75 1.8
rrn23 2 5630 2815 3.44
rrn4.5 2 208 104 0.13
rrn16 2 2982 1491 1.82
rrn5 2 242 121 0.15
Table 4 Annotated genes of the Indian jointvetch chloroplast genome
Category Groups Genes
Photosynthesis Subunits_of_photosystem_I psaA, psaB, psaC, psaI, psaJ
Subunits_of_photosystem_II psbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL,
psbM, psbN, psbT, psbZ
Subunits_of_NADH_dehydrogenase ndhA, ndhB, ndhB, ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ,
ndhK
Subunits_of_cytochrome_b/f_complex petA, petB, petD, petG, petL, petN
Subunits_of_ATP_synthase atpA, atpB, atpE, atpF, atpH, atpI
Large_subunit_of_Rubisco rbcL
Self-replication Large_subunits_of_ribosome rpl14, rpl16, rpl2, rpl2, rpl20, rpl23, rpl23, rpl32, rpl33, rpl36
Small_subunits_of_ribosome rps11, rps12, rps12, rps14, rps15, rps16, rps18, rps19, rps19, rps2, rps3,
rps4, rps7, rps7, rps8
DNA-dependent_RNA_polymerase rpoA, rpoB, rpoC1, rpoC2
Ribosomal_RNAs 8 rRNA
Transfer_RNAs 39 tRNAs
Other genes Maturase matK
Protease clpP1
Envelope_membrane_protein cemA
Acetyl-CoA_carboxylase accD
C-type_cytochrome_synthesis_gene ccsA
Translation_initiation_factor
protochlorophillide_reductase_subunit
Genes unknown Proteins_of_unknown_function Ycf1, ycf2, ycf3, ycf4
Table 5 Single sequence repeats (SSRs) and long repeats (LRs) in
the Indian jointvetch chloroplast genome
Type of repeats Distribution Number
SSR Genome 161
Coding 22
IRa/b 8
LSC 114
SSC 31
LR Hamming Distance = 0 22
Hamming Distance = 1 6
Hamming Distance = 2 28
Hamming Distance = 3 110
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 8 of 13
Gao et al. BMC Genomics (2024) 25:277
Fig. 4 Comparison of large sequence copy (LSC), inverted repeat (IRb, IRa), and small sequence copy (SSC) border regions of the chloroplast genomes
of Indian jointvetch and related species
Fig. 3 Phylogenetic tree constructed using the maximum-likelihood method based on alignments of complete chloroplast genome sequences. The
numbers at the nodes indicate bootstrap values from 1000 replicates. If the bootstrap value is 100, this number is not shown on the nodes. The species
newly sequenced in this study is shown in bold
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 9 of 13
Gao et al. BMC Genomics (2024) 25:277
voucher IBSC compared to the Indian jointvetch cp.
genome.
Chemical herbicide eects
Toxicity of pre-emergence herbicides
In pre-emergence application, as the doses of oxyuo-
rofen and saufenacil increased, Indian jointvetch was
signicantly inhibited on the 14th day after herbicide
treatment and showed a gradient decrease pattern there-
after. However, upon treatment with the other two herbi-
cides, bensulfuron-methyl and butralin, Indian jointvetch
did not show signicant growth inhibition until 21 days
after treatment (Fig.5). e toxicity calculation showed
that the GR50 value of oxyuorofen on Indian jointvetch
was 34.9g a.i. ha 1 and the GR90 value was 157.5g a.i.
ha 1; the GR50 value of saufenacil on Indian jointvetch
was 1.3g a.i. ha 1 and the GR90 value was 10.3g a.i. ha 1
(Fig.5).
Toxicity of post-emergence herbicides
With post-emergence application, with the increase in
the doses of orpyrauxifen-benzyl, pyrazosulfuron-ethyl,
and penoxsulam, Indian jointvetch was signicantly
inhibited on the 21st day after herbicide treatment and
showed a decreasing gradient thereafter; the other three
herbicides, quinclorac, 2-methyl-4-chlorophenoxy-
acetic acid, and pyraquinate, did not have a signicant
inhibitory eect on the growth of Indian jointvetch.
From the toxicity calculation, the GR50 value of orpy-
rauxifen-benzyl on Indian jointvetch was 0.3g a.i. ha 1
and the GR90 value was 1.3g a.i. ha 1. e GR50 value of
penoxsulam on Indian jointvetch was 4.0g a.i. ha 1 and
the GR90 value was 16.8 g a.i. ha 1. e GR50 value of
pyrazosulfuron-ethyl on Indian jointvetch was 11.9g a.i.
ha 1 and the GR90 value was g a.i. ha 1 737.5, but this far
exceeds its safe usage limit (45.0g a.i. ha 1) (Fig.6).
Discussion
According to the results of our on-site investigation,
Indian jointvetch causes serious harm to upland direct-
seeding rice elds in the lower reaches of the Yang-
tze River in China. Upland direct-seeding rice is an
important food crop grown in areas with limited water
resources presently. However, rampant weeds are an
obstacle to the cultivation of upland direct-seeding rice
[1416]. e rough management practice of the upland
direct-seeding rice eld system has further aggravated
weed ravages. Although studies have shown that Indian
jointvetch can be considered a green fertilizer that is
benecial for rice growth [40], we obtained contrasting
results. In fact, the tall Indian jointvetch plant type has
reduced the survival resources of rice, thereby seriously
restricting its yield. However, research on Indian joint-
vetch in rice elds is limited.
Table 6 Single nucleotide polymorphisms (SNPs) in S. cannabina voucher IBSC, G. soja, and G. max chloroplast genomes compared to
the Indian jointvetch chloroplast genome
Species Start Stop Synonymous Nonsynonymous CDS Intergenic Total SNPs
S. cannabina voucher IBSC 2 8 2000 1252 3281 1378 4659
G. soja 1 7 2344 1469 3840 1595 5435
G. max 1 7 2342 1470 3838 1594 5432
CDS, coding sequence
Fig. 5 Dose–response analyses for the response of Indian jointvetch to pre-emergence herbicides. The X-axis represents the dose (g a.i. ha− 1) and the
Y-axis represents the percentage of fresh weight to that of the untreated control. “Bm,“Sf, “Of,” and “Bl” represent the herbicides bensulfuron-methyl,
saufenacil, oxyuorofen, and butralin, respectively
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 10 of 13
Gao et al. BMC Genomics (2024) 25:277
Given the serious harm caused by Indian jointvetch in
rice elds, genetic research is important to gain a deeper
understanding of this weed and predict its evolution
in rice elds in the future. Because the cp. genome has
unique research value in phylogeny, species identica-
tion, and population genetics [21], we sequenced the
cp. genome of Indian jointvetch for the rst time in this
study, demonstrating a typical tetrad structure, including
genes encoding proteins, tRNAs, and rRNAs, similar to
the cp. genomes reported in other plants [22, 34, 41]. e
complete cp. genome of Indian jointvetch was 163,613bp
long, which is larger than that of other common plants
found in rice elds with reported cp. genomes, includ-
ing those of the genera Oryza [42], Echinochloa [41], and
Ammannia [34].
Repeating sequences are unique labels of a species,
including SSRs and LRs. SSRs, composed of 1–6 nucle-
otide repeat units, are widely distributed molecular
genetic markers in the cp. genome of plants [43, 44]. We
identied 161 SSRs in the cp. genome of Indian joint-
vetch, which can be used for population genetics analy-
ses and plant genotyping [4548]. We also identied 166
LRs in the Indian jointvetch cp. genome. LRs are a spe-
cial type of DNA repeat sequence that typically occupy
a large proportion of the genome [49]. ese repetitive
structures exhibit diversity in the cp. genome and pro-
mote molecular recombination [50], which has important
molecular and biological signicance in the study of plant
evolution [51]. us, the SSRs and LRs detected in this
study oer important biological information resources
for the identication of Indian jointvetch and to facilitate
genetic diversity and population structure research of
Aeschynomene.
In order to explore the closely related plants of Indian
jointvetch, we conducted phylogenetic tree analysis, IR
expansion and contraction comparisons, and SNP stud-
ies based on the cp. genome. Genome assembly and
characterization provides great value in dening species,
revealing phylogenetic relationships, and determining
taxonomic status [5255]. Phylogenetic analysis based on
the cp. genomes showed that species with a close genetic
relationship to Indian jointvetch included A. sinicus, M.
polymorpha, G. soja, G. max, and Sesbania cannabina
voucher IBSC. Notably, the genetic relationship between
the two species of S. cannabina registered in the NCBI
database is not close, suggesting that they may in fact
represent dierent species or diverged due to geographi-
cal isolation and local adaptation. IR expansion and con-
traction in the cp. genome are common phenomena in
plants [18], mainly occurring at the IR/SC junction [56],
which is a key driving force in plant evolution [5760].
e results of IR expansion and contraction analyses
further indicated that the genetic relationship between
Indian jointvetch and S. cannabina voucher IBSC was
closer. Because their boundary genes are completely
identical, except for dierences in length of the genes or
the distance between the genes and the boundary (Fig.4).
SNPs are important indicators of evolutionary dierences
between closely related species. ese direct molecular
markers can clearly display the exact nature and location
of allele variations [61]. Using the cp. genome of Indian
jointvetch as reference, we identied 3281, 3840, and
3838 SNPs in the cp. genomes of S. cannabina voucher
IBSC, G. soja, and G. max, respectively. e fewer SNPs
detected represents the fewer variations in single nucleo-
tides. is result further indicated that the Indian joint-
vetch and S. cannabina voucher IBSC have a relatively
close genetic relationship, which is consistent with the
phylogeny and IR expansion and contraction results.
ese SNPs can be used as important dierential nucleo-
tide databases to distinguish the species.
From the perspective of crop protection, it is neces-
sary to study reasonable measures to eliminate Indian
jointvetch from paddy elds. e advantages of chemical
methods for weed control in upland direct-seeding rice
elds have been reported for decades in terms of higher
Fig. 6 Dose –response analyses for response of Indian jointvetch to post-emergence herbicides. The X-axis represents the dose (g a.i. ha− 1) and the Y-axis
represents the percentage of fresh weight to that of the untreated control. “Fb,“Ps,” “MCPA, “Qc,“Pq, and “Pz” represent the herbicides orpyrauxifen-
benzyl, penoxsulam, 2-methyl-4-chlorophenoxyacetic acid, quinclorac, pyraquinate, and pyrazosulfuron-ethyl, respectively
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 11 of 13
Gao et al. BMC Genomics (2024) 25:277
yields and lower labor costs [3032]. However, the chem-
ical control of Indian jointvetch is currently very chal-
lenging [3]. Our results also indicated that the availability
of herbicides for Indian jointvetch control is limited. Her-
bicides commonly used to control broadleaf weeds in
rice elds in China, such as bensulfuron-methyl and
2-methyl-4-chlorophenoxyacetic acid [6264], had poor
eects on inhibiting the germination of Indian jointvetch.
Nevertheless, we found that among the 10 herbicides
tested, four met the requirements, including two pre-
emergence herbicides (saufenacil and oxyuorfen) and
two post-emergence herbicides (orpyrauxifen-benzyl
and penoxsulam). Given that mixing Indian jointvetch
seeds with rice seriously reduces the quality of the rice
[1], we strongly recommend using herbicides contain-
ing these four components to control Indian jointvetch,
which is rampant in rice elds; however, further research
is needed to establish eective control strategies.
It should be pointed out that the understanding of the
genetic characteristics of Indian jointvetch and the pre-
diction of the evolution of related plants into rice eld
weeds in this study are only based on the chloroplast
genome. Further research should be based on broader
morphological and whole-genome. Additionally, this
study only provides suggestions for chemical control of
Indian jointvetch. Further research is needed on control
strategies that are more conducive to biodiversity and
even the value of resource utilization of this weed.
Conclusions
In this study, we obtained the rst complete cp. genome
of Indian jointvetch (A. indica), a common weed found
in upland direct-seeding rice elds in the lower reaches
of the Yangtze River in China. Overall, our study indi-
cates that species with closer anity to Indian jointvetch
include (but are not limited to) Glycine soja, Glycine
max, and Sesbania cannabina based on the cp. genome.
erefore, it is necessary to be vigilant about these
plants becoming farmland weeds (especially S. canna-
bina), entering rice elds, and causing harm to rice. Note
that because A. sinicus and M. polymorpha do not have
complete IR regions, further analyses of these two spe-
cies were not conducted in this study, which may war-
rant further exploration of their potential harm in upland
direct-seeding rice elds. Toxicity evaluation of common
herbicides currently used in rice elds indicate the need
for the combination of herbicides to achieve eective
control of Indian jointvetch and improve the yield and
quality of upland direct-seeding rice.
Abbreviations
cp Chloroplast
GO Gene Ontology
GR50 The herbicide dose needed to cause 50% growth reduction of plant
IBSC South China Botanical Garden, Chinese Academy of Sciences
IR Inverted Repeat
KEGG Kyoto Encyclopedia of Genes and Genomes
LR Long repeat
LSC Large single copy
ML Maximum likelihood
MRCA Most recent common ancestor
NCBI National Center for Biotechnology Information
SSC Small single copy
SNP Single nucleotide polymorphism
SSR Simple sequence repeat
Supplementary Information
The online version contains supplementary material available at https://doi.
org/10.1186/s12864-024-10102-x.
Additional File 1
Additional File 2
Additional File 3
Acknowledgements
We thank Shanghai BIOZERON Biotechnology Co., Ltd. for performing the
high-throughput sequencing.
Author contributions
Study design, Y.G.; concept, Y.G., Z.T., and G.S.; data collection, Y.G., T.C., and J.L.;
analysis, Y.G., T.C., and J.L.; manuscript writing, Y.G; review, Z.T., and G.S.
Funding
This research was funded by Shanghai Agriculture Applied Technology
Development Program, China (Grant number T20210104).
Data availability
The datasets generated and/or analyzed during the current study are available
in the [National Center for Biotechnology Information] repository, [Accession
number: PRJNA963187].
Declarations
Ethics approval and consent to participate
The seeds of Aeschynomene indica collected in this study were approved by
the landowner.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
Author details
1Eco-Environmental Protection Research Institute, Shanghai Academy of
Agricultural Sciences, 201403 Shanghai, China
2School of Chemical and Environmental Engineering, Shanghai Institute
of Technology, 201418 Shanghai, China
Received: 7 December 2023 / Accepted: 8 February 2024
References
1. Riet-Correa F, Timm C, Barros S, Summers B. Symmetric focal degeneration in
the cerebellar and vestibular nuclei in swine caused by ingestion of Aeschyn-
omene indica seeds. Vet Pathol. 2003;40(3):311–6.
2. Flora Reipublicae Popularis Sinicae. Available online: http://www.iplant.cn/.
3. Martins MB, Agostinetto D, Fogliatto S, Vidotto F, Andres A. Aeschynomene
spp. Identication and Weed Management in Rice elds in Southern Brazil.
Agronomy. 2021;11(3):453.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 12 of 13
Gao et al. BMC Genomics (2024) 25:277
4. Hartmann LS, Rodrigues RS, Flores AS. O gênero Aeschynomene (Legumino-
sae-Papilionoideae) no estado de Roraima. Brasil Rodriguésia. 2019;70:1–14.
5. Ouyang W, Chen L, Yang Y. The eect of temperature on the germination rate
of Sesbania cannabina and Aeschynomene indica and the allelopathic eect
between them. South China Agric. 2021;15(20):179–80. (in Chinese).
6. Martins MB, Munhos TF, Schaedler CE, Agostinetto D, Andres A. Seed dynam-
ics of Aeschynomene denticulata and Aeschynomene indica. Weed Biol Manag.
2021;21(3):172–80.
7. Fleck NG, Lazaroto CA, Schaedler CE, Ferreira FB. Suscetibilidade De três
espécies de angiquinho (Aeschynomene spp.) a herbicidas de utilização em
pós-emergência em arroz irrigado. Curr Agricultural Sci Technol 2008;14(3).
8. Dornelles SHB, Sanchotene DM, Macedo L, Rodrigues SN. Controle pós-tardio
de angiquinho Aeschynomene denticulata na cultura do arroz.: Vivências.
2014;10(19):42–9.
9. Galon L, Guimarães S, Radünz AL, Lima AMD, Burg GM, Zandoná RR, Bastiani
MO, Belarmino JG, Perin GF. Competitividade relativa de cultivares de arroz
irrigado com Aeschynomene denticulata. Bragantia. 2015;74:67–74.
10. Kissmann KG, Groth D. Plantas infestantes e nocivas. Volume 1. Basf Brasileira
São Paulo; 1991.
11. Oliveira M. Aeschynomene indica L. no Brasil: Primeira citação. Inheringia.
2001;41:3–8.
12. Fruet BDL, Junior AM, Ulguim ADR. Survey of rice weed management and
public and private consultant characteristics in Southern Brazil. Weed Tech-
nol. 2020;34(3):351–6.
13. Mahajan G, Chauhan BS, Kumar V. Integrated weed management in rice.
Recent advances in weed management. New York: Springer; 2014. pp.
125–53.
14. Fofana, Rauber. Weed suppression ability of upland rice under low-input
conditions in West Africa. Weed Res. 2000;40(3):271–80.
15. Saito K. Weed pressure level and the correlation between weed competitive-
ness and rice yield without weed competition: an analysis of empirical data.
Field Crop Res. 2010;117(1):1–8.
16. Saito K, Futakuchi K. Improving estimation of weed suppressive ability of
upland rice varieties using substitute weeds. Field Crop Res. 2014;162:1–5.
17. Howe CJ, Barbrook AC, Koumandou VL, Nisbet RER, Symington HA, Wight-
man TF. Evolution of the chloroplast genome. Philosophical Trans Royal Soc
Lond Ser B: Biol Sci. 2003;358(1429):99–107.
18. Plunkett GM, Downie SR. Expansion and contraction of the chloroplast
inverted repeat in Apiaceae subfamily Apioideae. Syst Bot. 2000;25(4):648–67.
19. Wicke S, Schneeweiss GM, Depamphilis CW, Müller KF, Quandt D. The evolu-
tion of the plastid chromosome in land plants: gene content, gene order,
gene function. Plant Mol Biol. 2011;76(3):273–97.
20. Cosner ME, Raubeson LA, Jansen RK. Chloroplast DNA rearrangements in
Campanulaceae: phylogenetic utility of highly rearranged genomes. BMC
Evol Biol. 2004;4:27.
21. Hong Z, Wu Z, Zhao K, Yang Z, Zhang N, Guo J, Tembrock LR, Xu D. Compara-
tive analyses of ve complete chloroplast genomes from the genus Pterocar-
pus (Fabacaeae). Int J Mol Sci. 2020;21(11):3758.
22. Wu FH, Chan MT, Liao DC, Hsu CT, Lee YW, Daniell H, Duvall MR, Lin CS.
Complete chloroplast genome of Oncidium Gower Ramsey and evaluation of
molecular markers for identication and breeding in Oncidiinae. BMC Plant
Biol. 2010;10(1):68.
23. Li P, Lu RS, Xu WQ, Ohi-Toma T, Cai MQ, Qiu YX, Cameron KM, Fu CX. Compara-
tive genomics and phylogenomics of east Asian tulips (Amana, Liliaceae).
Front Plant Sci. 2017;8:451.
24. Bi Y, Zhang MF, Xue J, Dong R, Du YP, Zhang XH. Chloroplast genomic
resources for phylogeny and DNA barcoding: a case study on Fritillaria. Sci
Rep-UK. 2018;8(1):1184.
25. Kugita M, Kaneko A, Yamamoto Y, Takeya Y, Matsumoto T, Yoshinaga K. The
complete nucleotide sequence of the hornwort (Anthoceros Formosae)
chloroplast genome: insight into the earliest land plants. Nucleic Acids Res.
2003;31(2):716–21.
26. Henry RJ. Plant diversity and evolution: genotypic and phenotypic variation
in higher plants. Biologist. 2005;(1):121.
27. Yamane K, Yasui Y, Ohnishi O. Intraspecic cpDNA variations of diploid and
tetraploid perennial buckwheat, Fagopyrum Cymosum (Polygonaceae). Am J
Bot. 2003;90(3):339–46.
28. Singh R, Ghosh D. Eect of cultural practices on weed management in
rainfed upland rice. Int J Pest Manage. 1992;38(2):119–21.
29. Renu S, George T, Raharn C. Stale seedbed: A technique for non-chemical
weed control for direct seeded upland rice. In: Proceedings of the 19th Kerala
Science Congress, Kerala, India: 2007;29–31.
30. Adigun J, Kolo E, Adeyemi O, Daramola O, Badmus A, Osipitan O. Growth and
yield response of upland rice to nitrogen levels and weed control methods.
Int J Agron Agricultural Res. 2017;11(6):92–101.
31. Ogungbile A, Lagoke S. On-farm evaluation of the economics of chemical
weed control in oxen‐mechanized maize production in Nigerian savanna. Int
J Pest Manage. 1986;32(4):269–73.
32. Kolo E, Adigun JA, Adeyemi OR, Daramola OS, Bodunde JG. Eect of nitrogen
levels and weed management methods on weed abundance and yield of
upland rice (Oryza sativa L). J Agricultural Sci (Belgrade). 2020;65(2):137–50.
33. Concenço G, Andres A, Schreiber F, Moisinho I, Martins M. Control of
jointvetch (Aeschynomene spp.), establishment and productivity of rice as a
function of [imazapic + imazapyr] doses. J Agr Sci. 2018;10(4):287.
34. Gao Y, Li S, Yuan G, Fang J, Shen G, Tian Z. Comparison of Biological and
genetic characteristics between two most common broad-leaved weeds
in Paddy Fields: Ammannia Arenaria and A. multiora (Lythraceae). Biology.
2023;12(7):936.
35. Sakamoto Y, Ishiguro M, Kitagawa G. Akaike information criterion statistics.
Dordrecht, The Netherlands: D Reidel. 1986;81(10.5555):26853.
36. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-
analysis of large phylogenies. Bioinformatics. 2014;30(9):1312–13.
37. Felsenstein J. Condence limits on phylogenies: an approach using the
bootstrap. Evolution. 1985;39(4):783–91.
38. Jiao B, Wang K, Chang Y, Dong F, Pan X, Wu X, Xu J, Liu X, Zheng Y. Photodeg-
radation of the Novel Herbicide Pyraquinate in Aqueous Solution: kinetics,
photoproducts, mechanisms, and Toxicity Assessment. J Agric Food Chem.
2023;71(10):4249–57.
39. Shinozaki K, Ohme M, Tanaka M, Wakasugi T, Hayashida N, Matsubayashi T,
Zaita N, Chunwongse J, Obokata J, Yamaguchi-Shinozaki K. The complete
nucleotide sequence of the tobacco chloroplast genome: its gene organiza-
tion and expression. EMBO J. 1986;5(9):2043–9.
40. Nilanthi D, Inoka K, Madhushani P, Dissanayaka D, Kumarasinghe S. Eect of
weeds Diya Siyambala (Aeschynomene indica), Kudamatta (Fimbristylis mili-
acea) and Thunassa (Cyperus iria) on growth and yield of Rice (Oryza sativa L.)
variety Bg 379-2, ld 365, Bg 357. Int J Sci Res Publications. 2015;5(6):473–5.
41. Gao Y, Shen G, Yuan G, Tian Z. Comparative analysis of whole chloroplast
genomes of three common species of Echinochloa (Gramineae) in Paddy
Fields. Int J Mol Sci. 2022;23(22):13864.
42. National Center for Biotechnology Information. 2023; Available online:
https://www.ncbi.nlm.nih.gov/nuccore/?term=Oryza+sativa+chloroplast+co
mplete+genome (accessed on 4 April 2023).
43. Wu M, Li Q, Hu Z, Li X, Chen S. The complete Amomum Kravanh chloroplast
genome sequence and phylogenetic analysis of the commelinids. Molecules.
2017;22(11):1875.
44. Zhou J, Cui Y, Chen X, Li Y, Xu Z, Duan B, Li Y, Song J, Yao H. Complete chloro-
plast genomes of Papaver rhoeas and Papaver orientale: molecular structures,
comparative analysis, and phylogenetic analysis. Molecules. 2018;23(2):437.
45. Doorduin L, Gravendeel B, Lammers Y, Ariyurek Y, Chin-A-Woeng T, Vrieling K.
The complete chloroplast genome of 17 individuals of pest species Jacobaea
vulgaris: SNPs, microsatellites and barcoding markers for population and
phylogenetic studies. DNA Res. 2011;18(2):93–105.
46. He S, Wang Y, Volis S, Li D, Yi T. Genetic diversity and Population structure:
implications for conservation of wild soybean (Glycine soja Sieb. Et zucc)
based on nuclear and chloroplast microsatellite variation. Int J Mol Sci.
2012;13:12608–28.
47. Yang AH, Zhang JJ, Yao XH, Huang HW. Chloroplast microsatellite markers in
Liriodendron tulipifera (Magnoliaceae) and cross-species amplication in L.
Chinense. Am J Bot. 2011;98(5):e123–6.
48. Xue J, Wang S, Zhou SL. Polymorphic chloroplast microsatellite loci in
Nelumbo (Nelumbonaceae). Am J Bot. 2012;99(6):e240–4.
49. Han Y, Gao Y, Zhai X, Zhou H, Ding Q, Ma L. Assembly and comparative analy-
sis of chloroplast genome of wheat K-CMS line and maintainer line. BMC
Genom. 2020.
50. Guo H, Liu J, Luo L, Wei X, Zhang J, Qi Y, Zhang B, Liu H, Xiao P. Complete
chloroplast genome sequences of Schisandra chinensis: genome structure,
comparative analysis, and phylogenetic relationship of basal angiosperms.
Sci China Life Sci. 2017;60:1286–90.
51. Cavalier-Smith T. Chloroplast Evolution: secondary Symbiogenesis and mul-
tiple losses. Curr Biol. 2002;12(2):R62–4.
52. Yang JB, Yang SX, Li HT, Yang J, Li DZ. Comparative chloroplast genomes of
Camellia species. PLoS ONE. 2013;8(8):e73053.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Page 13 of 13
Gao et al. BMC Genomics (2024) 25:277
53. Li HT, Yi TS, Gao LM, Ma PF, Zhang T, Yang JB, Gitzendanner MA, Fritsch PW, Cai
J, Luo Y. Origin of angiosperms and the puzzle of the jurassic gap. Nat Plants.
2019;5(5):461–70.
54. Li X, Yang Y, Henry RJ, Rossetto M, Wang Y, Chen S. Plant DNA barcoding: from
gene to genome. Biol Rev. 2015;90(1):157–66.
55. Zhang Z, Zhang Y, Song M, Guan Y, Ma X. Species identication of Dracaena
using the complete chloroplast genome as a super-barcode. Front Pharma-
col. 2019;10:1441.
56. Wang W, Messing J. High-throughput sequencing of three Lemnoi-
deae (duckweeds) chloroplast genomes from total DNA. PLoS ONE.
2011;6(9):e24670.
57. Raubeson LA, Peery R, Chumley TW, Dziubek C, Fourcade HM, Boore JL,
Jansen RK. Comparative chloroplast genomics: analyses including new
sequences from the angiosperms Nuphar advena and Ranunculus macran-
thus. BMC Genomics. 2007;8(1):174.
58. Yang M, Zhang X, Liu G, Yin Y, Chen K, Yun Q, Zhao D, Al-Mssallem IS, Yu J. The
complete chloroplast genome sequence of date palm (Phoenix dactylifera L).
PLoS ONE. 2010;5(9):e12762.
59. Daniell H, Lin CS, Yu M, Chang WJ. Chloroplast genomes: diversity, evolution,
and applications in genetic engineering. Genome Biol. 2016;17(1):1–29.
60. Logacheva MD, Krinitsina AA, Belenikin MS, Khazov K, Konorov EA, Kuptsov
SV, Speranskaya AS. Comparative analysis of inverted repeats of polypod fern
(Polypodiales) plastomes reveals two hypervariable regions. BMC Plant Biol.
2017;17(2):61–73.
61. Germano J, Klein AS. Species-specic nuclear and chloroplast single nucleo-
tide polymorphisms to distinguish Picea glauca, P. mariana and P. rubens.
Theor Appl Genet. 1999;99(1):37–49.
62. Shi XR, Zhao HY, Gong B, Cao WS. Study on the Control Eect of Dierent
Herbicide Treatments in Direct Water Rice Fields. Shanghai Agricultural Sci
Technol 2013;(4):132–3. (in Chinese).
63. Zhang JL, Qiao JL, Meng ZH, Du JL, Li CG, Liu YW. Experiment on the control
of weeds in Rice Transplant elds with 10% bensulfuron WP. Heilongjiang
Agricultural Sci 2009;(2):76. (in Chinese).
64. Jin DJ, Chi XC, Gu XJ, Yang XY, Zhang YH, Fu Y. Ecacy and selectivity of 65%
of MCPA-Dimethylamine Salt (AS) to Rice in transplanted Rice Fields. J Weed
Sci. 2017;035(004):25–9. (in Chinese).
Publisher’s Note
Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional aliations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
Article
Full-text available
Ammannia arenaria and A. multifloras, morphologically similar at the seedling stage, are the most common broad-leaved weeds in paddy fields. Our study showed that A. arenaria occupied more space than A. multifloras when competing with rice. However, A. multifloras germination has lower temperature adaptability. No difference in sensitivity to common herbicides between two Ammannia species was observed. Chloroplast (cp) genomes could be conducive to clarify their genetic relationship. The complete cp genome sequences of A. arenaria (158,401 bp) and A. multiflora (157,900 bp) were assembled for the first time. In A. arenaria, there were 91 simple sequence repeats, 115 long repeats, and 86 protein-encoding genes, one, sixteen, and thirty more than those in A. multiflora. Inverted repeats regions expansion and contraction and the phylogenetic tree based on cp genomes demonstrated the closely relationship between the two species. However, in A. arenaria, 20 single nucleotide polymorphisms in the CDS region were detected compared to A. multiflora, which can be used to distinguish the two species. Moreover, there was one unique gene, infA, only in A. arenaria. This study provides reliable molecular resources for future research focusing on the infrageneric taxa identification, phylogenetic resolution, population structure, and biodiversity of Ammannia species.
Article
Full-text available
Echinochloa crus-galli var. crus-galli, E. crus-galli var. zelayensis, and E. glabrescens, morphologically similar at the seedling stage, are the most pernicious barnyard grass species in paddy fields worldwide. Chloroplast (cp) genomes could be conducive to their identification. In this study, we assembled the complete cp genome sequences of Echinochloa crus-galli var. crus-galli (139,856 bp), E. crus-galli var. zelayensis (139,874 bp), and E. glabrescens (139,874 bp), which exhibited a typical circular tetramerous structure, large and small single-copy regions, and a pair of inverted repeats. In Echinochloa crus-galli var. crus-galli, there were 136 simple sequence (SSRs) and 62 long (LRs) repeats, and in the other two species, 139 SSRs and 68 LRs. Each cp genome contains 92 protein-encoding genes. In Echinochloa crus-galli var. crus-galli and E. glabrescens, 321 and 1 single-nucleotide polymorphisms were detected compared to Echinochloa crus-galli var. zelayensis. IR expansion and contraction revealed small differences between the three species. The phylogenetic tree based on cp genomes demonstrated the phylogenetic relationship between ten barnyard grass species and other common Gramineae plants, showing new genetic relationships of the genus Echinochloa. This study provides valuable information on cp genomes, useful for identifying and classifying the genus Echinochloa and studying its phylogenetic relationships and evolution.
Article
Full-text available
In 2002, a survey carried out in rice paddies in the Rio Grande do Sul (RS) state reported the occurrence of nine species of jointvetch (Aeschynomene). Due to their semi-aquatic habit, some species adapted to irrigated rice fields, which led to their being considered the worst broadleaf weed in RS. Although farmers have successfully implemented weed management practices, Aeschynomene plants have reportedly escaped chemical control. This study aims to identify the species of Aeschynomene that occur in rice fields in RS and to evaluate the reasons why escapes are occurring. A survey was carried out by collecting mature seeds from individual adult plants. A questionnaire on the management practices employed in each field was administered to 54 farmers and 18 extension agents, each of whom was responsible for one of the surveyed rice fields. This survey found four species of Aeschynomene are present in rice fields in RS: A. denticulata, A. indica, A. rudis, and A. sensitiva. The results suggest that the explanation for escapes may lie in the management practices adopted by farmers, which are focused on the control of weedy grasses. Escapes are also associated with problems such as the lack of irrigation uniformity and out-of-stage, late herbicide applications.
Article
Full-text available
Weed infestation and inherent low soil fertility are among the major factors attributed to the low yield of rice in Nigeria. Field trials were therefore conducted to evaluate the effect of nitrogen application levels and weed control methods on growth and yield of upland rice (var. NERICA 2) at the Teaching and Research Farm of the Federal University of Agriculture, Abeokuta (07 o 15'N, 03 o 25'E) during 2015 and 2016 cropping seasons. Three nitrogen (N) levels (0, 60 and 90 kg/ha) were evaluated and they constituted the main plot treatments, while three weed control treatments, viz: pre-emergence application of Orizo Plus ® (propanil plus 2, 4-D) at 2.0 kg a.i ha-1 , Orizo Plus ® at 2.0 kg a.i ha-1 followed by supplementary hoe weeding (SHW) at 6 weeks after sowing (WAS) and three hoe-weeding regimes at 3, 6 and 9 WAS, and a weedy check constituted the sub-plot treatments. All the treatments in different combinations were laid out in a randomized complete block design with a split-plot arrangement with three replicates. Results indicated a significant (p≤0.05) increase in weed density and dry matter with an increase in N application level from 0 to 90 kg ha-1. Similarly, crop vigour and plant height increased significantly (p≤0.05) with increasing N application levels up to 90 kg ha-1. However, 60 and 90 kg N ha-1 were at par in increasing the number of tillers, leaf area index and yield attributes of rice. All the weed control methods resulted in a significant (p≤0.05) reduction in weed density and dry matter with subsequent increase in rice growth and yield than the weedy check. Pre-emergence application of Orizo Plus ® followed by SHW at 6 WAS and three hoe-weeding regimes resulted in significantly (p≤0.05) lower weed density and dry matter, and a higher number of tillers, panicle weight and grain yield than a sole application of Orizo Plus ®. With Orizo Plus ® followed by one SHW or three Emmanuel Kolo et al. 138 hoe-weeding regimes, increasing N application levels resulted in a significant (p≤0.05) increase in grain yield of rice. However, with Orizo Plus ® applied alone, increasing N application levels did not increase rice grain yield. These results suggest that Orizo Plus ® at 2.0 kg a.iha-1 followed by one SHW at 6 WAS integrated with N application at 90 kg ha-1 is adequate to effectively control weeds and increase rice yield in the rainforest-savannah transition zone of Nigeria.
Article
Full-text available
Pterocarpus is a genus of trees mainly distributed in tropical Asia, Africa, and South America. Some species of Pterocarpus are rosewood tree species, having important economic value for timber, and for some species, medicinal value as well. Up to now, information about this genus with regard to the genomic characteristics of the chloroplasts has been limited. Based on a combination of next-generation sequencing (Illumina Hiseq) and long-read sequencing (PacBio), the whole chloroplast genomes (cp genomes) of five species (rosewoods) in Pterocarpus (Pterocarpus macrocarpus, P. santalinus, P. indicus, P. pedatus, P. marsupium) have been assembled. The cp genomes of five species in Pterocarpus have similar structural characteristics, gene content, and sequence to other flowering plants. The cp genomes have a typical four-part structure, containing 110 unique genes (77 protein coding genes, 4 rRNAs, 29 tRNAs). Through comparative genomic analysis, abundant simple sequence repeat (SSR)loci (333–349) were detected in Pterocarpus, among which A /T single nucleotide repeats accounted for the highest proportion (72.8–76.4%). In the five cp genomes of Pterocarpus, eight hypervariable regions, including trnH-GUG_psbA, trnS-UGA_psbC, accD-psaI, ndhI-exon2_ndhI-exon1, ndhG_ndhi-exon2, rpoC2-exon2, ccsA, and trnfM-CAU, are proposed for use as DNA barcode regions. In the comparison of gene selection pressures (P. santalinus as the reference genome), purifying selection was inferred as the primary mode of selection in maintaining important biological functions. Phylogenetic analysis shows that Pterocarpus is a monophyletic group. The species P. tinctorius is resolved as early diverging in the genus. Pterocarpus was resolved as sister to the genus Tipuana.
Preprint
Full-text available
The sterile line is vital for the heterosis utilization of wheat (Triticum aestivum L.), and the cytoplasmic male sterility (CMS) line has more practical significance in the utilization of heterosis. To reveal the sterile mechanism of wheat K-CMS line K519A from the perspective of the chloroplast genome, the chloroplast genomes of the K519A and its same nuclear genotype maintainer line 519B were sequenced using second-generation high-throughput technology and assembled. Then the chloroplast genomic contents, SSR sequences, long repeat fragments, and qPCR were analyzed. The results showed that the total length of K519A and 519B were 136,996 bp and 136,235 bp, respectively. Both chloroplast genomes encode 126 genes, including 89 protein-coding genes, 8 rRNA genes, and 39 tRNA genes. There were 186 and 188 SSRs in K519A and 519B, respectively. And a developed SSR maker, which from atpF, can be used to distinguish the cytoplasmic type of Aegilops kotschyi and Triticum aestivum, respectively. There were 50 and 52 long repeat fragments, which come from the gene sequences, in K519A and 519B, respectively. A total of 107 mutations between K519A and 519B, which from 45 genes, were identified, of which, 16 genes (matK, rps16, rpoB, rpoC2, atpI, ndhK, atpB, rbcL, psbB, psbH, rpl14, ndhH, ndhF, rpl32, ccsA, ndhA) contained non-synonymous mutations. Further, the qPCR analysis was performed, and the gene expression levels of selected six genes for K519A compared to 519B were mostly downregulated at the binucleate and trinucleate stages of pollen development, thus, these non-synonymous mutant genes might affect the fertility of K519A.
Article
Full-text available
The taxonomy and nomenclature of Dracaena plants are much disputed, particularly for several Dracaena species in Asia. However, neither morphological features nor common DNA regions are ideal for identification of Dracaena spp. Meanwhile, although multiple Dracaena spp. are sources of the rare traditional medicine dragon’s blood, the Pharmacopoeia of the People’s Republic of China has defined Dracaena cochinchinensis as the only source plant. The inaccurate identification of Dracaena spp. will inevitably affect the clinical efficacy of dragon’s blood. It is therefore important to find a better method to distinguish these species. Here, we report the complete chloroplast (CP) genomes of six Dracaena spp., D. cochinchinensis, D. cambodiana, D. angustifolia, D. terniflora, D. hokouensis, and D. elliptica, obtained through high-throughput Illumina sequencing. These CP genomes exhibited typical circular tetramerous structure, and their sizes ranged from 155,055 (D. elliptica) to 155,449 bp (D. cochinchinensis). The GC content of each CP genome was 37.5%. Furthermore, each CP genome contained 130 genes, including 84 protein-coding genes, 38 tRNA genes, and 8 rRNA genes. There were no potential coding or non-coding regions to distinguish these six species, but the maximum likelihood tree of the six Dracaena spp. and other related species revealed that the whole CP genome can be used as a super-barcode to identify these Dracaena spp. This study provides not only invaluable data for species identification and safe medical application of Dracaena but also an important reference and foundation for species identification and phylogeny of Liliaceae plants.
Article
Pyraquinate, a newly developed 4-hydroxyphenylpyruvate dioxygenase class herbicide, has shown excellent control of resistant weeds in paddy fields. However, its environmental degradation products and corresponding ecotoxicological risks after field application remain ambiguous. In this study, we systematically investigate the photolytic behaviors of pyraquinate in aqueous solutions and in response to xenon lamp irradiation. The degradation follows first-order kinetics, and its rate depends on pH and the amount of organic matter. No vulnerability to light radiation is indicated. Ultrahigh-performance liquid chromatography with quadrupole-time-of-flight mass spectrometry and UNIFI software analysis reveals six photoproducts generated by methyl oxidation, demethylation, oxidative dechlorination, and ester hydrolysis. Gaussian calculation suggests that activities due to hydroxyl radicals or aquatic oxygen atoms caused these reactions on the premise of obeying thermodynamic criteria. Practical toxicity test results show that the toxicity of pyraquinate to zebrafish embryos is low but increases when the compound is combined with its photoproducts.
Article
Jointvetch (Aeschynomene spp., Fabaceae) is considered a problematic broadleaf weed for rice production in Brazil, the United States, and Asia. Seed dormancy is particularly important for the perpetuation of annual weeds, representing the only link between different generations and thus the source of all future weed infestations. Seed dynamics in annual species are complex and related to many biotic and abiotic factors. The objectives of this study were to determine whether dormancy occurs in Aeschynomene denticulata and Aeschynomene indica seeds, and, if so, what is the most efficient method to break it. In addition, the persistence of seeds in soil also was investigated for both species. Three experiments were conducted, two in lab conditions and one in field conditions. The first experiment consisted of germination tests conducted in controlled conditions to determine whether seeds have dormancy. The second experiment tested the effectiveness of preheating, precooling, potassium nitrate, gibberellic acid, water soaking, chemical scarification, and mechanical scarification as dormancy breaking methods. Finally, the third experiment evaluated seedbank longevity for 1‐year period in the rice‐cultivated soil during which seeds were buried and recovered every 3 months. Variations in physical dormancy are present at different levels in both species due to seed‐coat impermeability to water absorption. Meanwhile, A. denticulata can persist over a year in soil seed bank but A. indica seeds only persisted for half a year.
Chapter
This book contains 15 chapters that review the aspects of plant evolution as it relates to variation in plant genomes (gene level) and associated variations in plant phenomes (trait level). Topics covered include chloroplast and mitochondrial genomes, reticulate evolution, polyploidy, population genetics within a species, the evolution of the flower, diversity in plant cell walls and in secondary metabolism, diversity and evolution of gymnosperms, and the importance of plant diversity in ecology and agriculture. This book is intended for those working within the areas of plant genetics, evolution, biodiversity, botany and agriculture.