ArticlePDF Available

Evolutionary histories determine DNA barcoding success in vascular plants: Seven case studies using intraspecific broad sampling of closely related species

Authors:

Abstract and Figures

Background: Four plastid regions, rpoB, rpoC1, matK, and trnH-psbA, have been recommended as DNA barcodes for plants. Their success in delimiting species boundaries depends on the existence of a clear-cut difference between inter- and intraspecific variability. We tested the ability of these regions to discriminate among closely related species in seven genera of flowering plants with different generation times (trees, perennials, and annuals). To ensure a maximum coverage of intraspecific diversity, and therefore to better evaluate the resolution power of each barcode, we applied a population genetics approach by sampling three to 45 individuals per species over a wide geographical range. Results: All possible combinations between loci were analysed, which showed that using more than one locus does not always improve the resolution power. The trnH-psbA locus was most effective at discriminating among closely related species (Acer, Lonicera, Geranium, and Veronica), singly or in combination. For Salix, Adenostyles, and Gentiana, the best results were obtained with the combination of matK, rpoB, and trnH-psbA. No barcoding gap was found within six genera analysed, excepting Lonicera. This is due to shared polymorphisms among species, combined with very divergent sequences within species. These genetic patterns reflect incomplete lineage sorting and hybridization events followed by chloroplast capture. Conclusions: Our results strongly suggest that adding trnH-psbA to the two obligate DNA barcodes proposed by the CBOL plant-working group (matK and rbcL) should be mandatory for closely related species. In our sampling, generation time had no influence on DNA barcoding success, as the best and worst identification successes were found for the two tree genera (Acer, 64 % success and Salix, 86 % failure). Evolutionary histories are the main factor influencing DNA barcoding success in the studied genera.
Content may be subject to copyright.
R E S E A R C H A R T I C L E Open Access
Evolutionary histories determine DNA
barcoding success in vascular plants: seven
case studies using intraspecific broad
sampling of closely related species
Sofia Caetano Wyler
1,2
and Yamama Naciri
1*
Abstract
Background: Four plastid regions, rpoB, rpoC1, matK, and trnH-psbA, have been recommended as DNA barcodes
for plants. Their success in delimiting species boundaries depends on the existence of a clear-cut difference between
inter- and intraspecific variability. We tested the ability of these regions to discriminate among closely related species in
seven genera of flowering plants with different generation times (trees, perennials, and annuals). To ensure a maximum
coverage of intraspecific diversity, and therefore to better evaluate the resolution power of each barcode, we applied a
population genetics approach by sampling three to 45 individuals per species over a wide geographical range.
Results: All possible combinations between loci were analysed, which showed that using more than one locus does
not always improve the resolution power. The trnH-psbA locus was most effective at discriminating among closely
related species (Acer,Lonicera,Geranium,andVeronica), singly or in combination. For Salix,Adenostyles, and Gentiana,
the best results were obtained with the combination of matK, rpoB, and trnH-psbA. No barcoding gap was found
within six genera analysed, excepting Lonicera. This is due to shared polymorphisms among species, combined with
very divergent sequences within species. These genetic patterns reflect incomplete lineage sorting and hybridization
events followed by chloroplast capture.
Conclusions: Our results strongly suggest that adding trnH-psbA to the two obligate DNA barcodes proposed by the
CBOL plant-working group (matKandrbcL) should be mandatory for closely related species. In our sampling, generation
time had no influence on DNA barcoding success, as the best and worst identification successes were found for the
two tree genera (Acer,64%successandSalix, 86 % failure). Evolutionary histories are the main factor influencing DNA
barcoding success in the studied genera.
Keywords: Acer,Adenostyles, Chloroplast capture, Incomplete lineage sorting, Interspecific hybridization, Gentiana,
Geranium,Lonicera,Salix,Veronica
Background
DNA barcoding uses a short DNA sequence from a
standard locus to identify the species to which a particu-
lar specimen belongs [1]. Since DNA barcoding was first
used in plants, several regions have been recommended
as universal barcodes [27]. Primarily located in the
chloroplast genome, these regions focus on coding and
non-coding loci. Kress and Erickson [2] proposed the
combined use of rbcL and trnH-psbA, but other combi-
nations have been suggested as well ([810]; among
others). More recently, the Plant Working Group of the
Consortium for the Barcode of Life adopted rbcLand
matK as the core DNA barcodes for plants [11], with
trnH-psbA as an additional marker. Other studies have
suggested the use of the nuclear ribosomal locus ITS
[4, 12, 13], but the aim of the present study was to test
for the accuracy of the chloroplast barcodes per se and
we therefore selected matK, rpoC1, rpoB, and trnH-
* Correspondence: Yamama.Naciri@ville-ge.ch
1
Laboratoire de Systématique Végétale et Biodiversité, Conservatoire et
Jardin botaniques & University of Geneva, Chemin de lImpératrice, 1, 1292,
Chambésy, Geneva, Switzerland
Full list of author information is available at the end of the article
© 2016 Caetano Wyler and Naciri. Open Access This article is distributed under the terms of the Creative Commons
Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution,
and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link
to the Creative Commons license, and indicate if changes were made. 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.
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103
DOI 10.1186/s12862-016-0678-0
psbA. The barcode studies published so far agree that
matKandtrnH-psbA are the two most promising
chloroplast regions for discriminating among closely re-
lated species, whereas other regions, such as rbcL, are
more suitable for identifications at the family and/or the
genus level [14]. This is the main reason why we decided
to discard rbcL, although it is one of the official barcodes.
Because the debate has long focused on which
marker(s) should be used to obtain the best assignment
to species [2, 4, 7, 10, 11, 1320], other fundamental is-
sues have received less attention, although they are of
high relevance for barcoding success. One of these issues
is how many individuals should be analysed within a
species to generate a reliable reference for an accurate
identification. Early studies that analysed the success of
DNA barcoding [2, 14, 16] did not use the closest spe-
cies when pairs of species were selected. Accordingly, a
higher identification success was usually obtained when
barcoding floras, for which closely related taxa are not
always included, versus taxonomical groups for which it
is usually the case [7, 8]. Meyer and Paulay [21] raised
the sample size concern but it has rarely been addressed
in barcoding studies (but see [7, 22]), although it is par-
ticularly critical when working with closely related spe-
cies, for which intra- and interspecific genetic variation
may overlap quite frequently. Therefore the methods by
which intraspecific variability is documented has a direct
influence on the accuracy with which a given DNA
sequence identifies species.
DNA barcoding success depends on the existence of a
clear cut-off between intraspecific variation and inter-
specific divergence, the so-called barcoding gap.The
barcoding gap is largely dependent on the studied
groups and species, which constitutes a second issue
that has hardly ever been addressed (but see [7]). Many
plant species evolved recently through adaptive radia-
tions and rapid speciation [3, 2326]. Recent speciation
with consecutive incomplete lineage sorting often results
in reduced sequence divergence between the newly spe-
ciating taxa [2729]. In the worst case, i.e. retention of
ancestral polymorphism(s) among species, the identifica-
tion of specimens is impossible [30]. Problematic identi-
fication of specimens also arises from hybridization
between species, which is very frequent [31, 32], and
polyploidization [29, 33, 34]. Therefore, the success of
DNA barcoding is expected to vary among groups
depending on their evolutionary history.
Still, a general prediction about DNA barcoding suc-
cess can be made based on life traits such as the gener-
ation time. The short generation times that characterize
annual plants are expected to lead to a rapid accumula-
tion of mutations and to prompt species differentiation.
Significant barcoding gaps are expected for such plants,
leading to high DNA barcoding assignment success. The
longer life spans and slower accumulation of mutations
in woody plants are expected to result in poorer species
delimitations [35, 36].
In this study, we analysed the impact of generation
times and large sample sizes on DNA barcoding success.
We addressed this question using four chloroplast loci
(matK, rpoB, rpoC1, and trnH-psbA) that have been pro-
posed as barcodes [14]. These markers were evaluated
for closely related species within seven genera that dis-
play different generation times: Acer and Salix (trees);
Adenostyles,Gentiana, and Lonicera (perennials); and
Geranium and Veronica (annuals). Within genera, we se-
lected species that have clear taxonomical status with
overlapping geographical distributions. We then sampled
as many populations as possible in order to assess intra-
specific and interspecific variation in the barcoding loci
to infer how well specimens could be assigned to species
with the selected chloroplast barcodes.
Results and discussion
Sampling
A total of 485 individuals were sampled for the 27 spe-
cies used in this study (Additional file 1). Differences in
sampling sizes per genus are explained by the relative
abundance of some species (Acer103 individuals) com-
pared to others (Geranium16 individuals) and by the
effort put into sampling Gentiana (137 individuals) for a
detailed study on the phylogeography of the Ciminalis
group [37]. Samples were collected in Austria, the Czech
Republic, France, Italy, Norway, Portugal, Switzerland,
the United Kingdom, Spain, and Sweden from 37.05° to
69.30° in latitude and from 8.38° to 22.48° in longitude.
Primer universality and amplification success
A DNA barcode must fulfil several requirements and
should optimally be universal (present in all taxa), easily
amplified (i.e., without species-specific PCR primers),
short enough (so that it can be easily sequenced, even
on degraded samples), informative at the species level
(with enough variation insuring a satisfactory identifica-
tion of species), and conserved or slightly polymorphic
at the intraspecific level (so that a barcode gap can be
observed).
Four candidate chloroplast regions were targeted in
the present study: matK, rpoC1, rpoB, and trnH-psbA.
Only 440 specimens were amplified and sequenced suc-
cessfully for the four loci (91 %). Loci were sequenced
with 100 % success, except for rpoC1 and rpoB in one
individual of Gentiana, and matKinAcer,Gentiana,
Lonicera, and Veronica (Table 1). We used four combi-
nations of five matK primers (one of them newly de-
signed in this study) to improve the results (Additional
file 2). Still, we were not able to obtain matK sequences
from 13 individuals of Veronica hederifolia (sequencing
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 2 of 11
success: 58.6 %; Table 1). This marker is known to have
a lower success rate of PCR amplification and sequen-
cing [11, 13] and our results emphasize the lack of
primer universality for this DNA barcode, even at the
genus level (Acer and Veronica; Additional file 2). More-
over, generating fully bidirectional sequences for matK
was sometimes challenging, a problem that has also been
reported in many families, including Asteraceae [27] and
Lemnaceae [10].
Sequence variation and discriminating power
Alignments, sequence variation analyses, and identifica-
tion of unique sequences were performed within each
genus separately. The alignment lengths for rpoC1 and
rpoB were conserved for all genera, while those of matK
and trnH-psbA ranged from 761 to 1228 bp and from
325 to 525 bp, respectively (Table 1). For the trnH-psbA
spacer, the differences in length are not surprising and
are easily explained by a high number of insertion/dele-
tion events. The use of different primer pairs for differ-
ent genera explains the range in matK product size.
Sequence variation was quantified using the number
of conserved and parsimony informative sites. The
percentage of conserved sites was high for each genus,
ranging from 80 % in Gentiana for trnH-psbA to 100 %
in Salix and Adenostyles for rpoC1 and rpoB, respectively
(Table 1). The percentage of congeneric species resolved
as monophyletic was accordingly very low for rpoC1 and
rpoB. This is not a surprising result given the slow evo-
lutionary rate of these two coding regions. These loci are
Table 1 Diversity measures for matK, rpoC1, rpoB, and trnH-psbA, given separately for the seven genera (n is the number of sampled
individuals)
matKrpoC1 rpoBtrnH-psbA
Acer (n= 103) Aligned length (bp) 849 508 349 512
Sequencing success (%) 95.1 100 100 100
Conserved sites (%) 98 99.2 99.1 93.8
Parsimony informative sites (%) 1.3 0.8 0.9 3.7
Salix (n= 69) Aligned length (bp) 855 508 349 325
Sequencing success (%) 100 100 100 100
Conserved sites (%) 96.4 100 99.7 91.4
Parsimony informative sites (%) 0 0 0.3 0.9
Adenostyles (n= 37) Aligned length (bp) 798 508 349 508
Sequencing success (%) 100 100 100 100
Conserved sites (%) 98.2 99.8 100 99
Parsimony informative sites (%) 0.1 0.2 0 0.8
Gentiana (n= 135) Aligned length (bp) 761 508 349 460
Sequencing success (%) 98.5 99.3 99.3 100
Conserved sites (%) 92.6 96.7 98.3 80
Parsimony informative sites (%) 6.6 1.6 1.7 19.6
Lonicera (n= 67) Aligned length (bp) 1190 508 340 525
Sequencing success (%) 85.1 100 100 100
Conserved sites (%) 96.7 99.8 98.8 97.1
Parsimony informative sites (%) 1.7 0.2 1.2 2.7
Geranium (n= 16) Aligned length (bp) 769 508 349 356
Sequencing success (%) 87.5 100 100 100
Conserved sites (%) 94.1 96.7 99.4 89
Parsimony informative sites (%) 5.6 1.6 0 2.5
Veronica (n= 58) Aligned length (bp) 1228 508 349 393
Sequencing success (%) 58.6 100 100 100
Conserved sites (%) 91.5 94.5 94.3 81.7
Parsimony informative sites (%) 5.7 5.3 5.2 15.5
Total aligned sequence length (bp), percentage of individuals successfully amplified and sequenced, percentage of conserved and parsimony informative
characters in the aligned sequences
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 3 of 11
therefore not suitable to distinguish closely related spe-
cies, as also reported in other studies (e.g., [10]). We
highlight that both loci have slightly lower resolution
powers compared to that of the recommended DNA
barcode rbcL [11]. Therefore the use of the latter region
would not have dramatically changed our results in the
present study.
The percentage of parsimony informative sites was low
for most markers in all genera, especially rpoC1 and
rpoB (mean = 1.4 and 1.3 %, respectively). The locus
trnH-psbA harbours the highest percentage of parsi-
mony informative sites, except in Geranium, for which
the highest value is found with matK (5.6 % instead of
2.5 % with trnH-psbA).
When considered separately, the locus with the highest
number of sequences private to a single species was found
with trnH-psbA (Table 2). Accordingly, the highest identifi-
cation success at the species level was also observed using
this locus. The ability of trnH-psbA to distinguish species is
generally well accepted [16]. Many studies have recom-
mended using this marker as a DNA barcode on a regular
basis [2, 3840]. Moreover, its use in intraspecific popula-
tion studies [41, 42] highlights its utility for discriminating
closely related species, which agrees with the results
obtained here. Intergenic spacers are generally difficult to
align across genera [43], but performing the analyses inde-
pendently within each genus can surpass this obstacle.
Monophyly tested by phylogenetic trees
For each barcode, we estimated the recovered species
monophyly using multiple individuals per species and
phylogenetic NJ trees (Additional file 3). It should be noted
that the main purpose of the trees was not to study evolu-
tionary relationships, but rather species identification.
The 103 Acer individuals were divided into three clades
for matK, rpoC1, and trnH-psbA: 1, A. campestre L. and A.
platanoides L., 2, A. opalus Mill. and A. monspessulanum
L., and 3, A. pseudoplatanus Falk. With rpoB, A. campestre,
A. platanoides,A. opalus,andA. monspessulanum grouped
together in a single clade. Adenostyles species did not clus-
ter into distinguishable clades with the four markers. For
Gentiana, the four loci separated the four species into two
main clades: 1, G. alpina Vill. and G. clusii E.P.Perrier &
Songeon, and 2, G. acaulis L. and G. angustifolia Vill. Still,
three G. alpina individuals were clustered in the second
clade. For Geranium,rpoC1 and trnH-psbAweretheonly
markers able to distinguish G. columbinum L. from the
other two species that clustered together. For Lonicera,only
rpoC1 failed to distinguish the four species into monophy-
letic clades (L. caerulea L., L. nigra L., and L. alpigena L.
clustered in a single clade). Salix species were indiscernible
with the four DNA barcodes. Veronica hederifolia L. indi-
viduals formed a monophyletic clade with three loci (matK
failed to amplify this species). With matK, two clades could
be observed, the first one comprising almost all V. arvensis
Table 2 Assignment success for matK, rpoC1, rpoB, and trnH-psbA given alone and arranged according to all possible combinations
Opt. for Option. Option1: matK+rpoC1; Option2: matK+rpoB; Option3: matK+ trnH-psbA; Option4: rpoC1 + rpoB; Option5: rpoC1 + trnH-psbA; Option6: rpoB+trnH-
psbA; Option7: matK+rpoC1 + rpoB; Option8: matK+rpoC1 + trnH-psbA; Option9: matK+rpoB+trnH-psbA; Option10: rpoC1 + rpoB+trnH-psbA; Option11: matK+
rpoC1 + rpoB+trnH-psbA. Number of individuals successfully amplified and sequenced, number of total different sequences within the genus and the ones that
are private to a single species, and number of individuals harbouring speciesprivate sequences. The statistics are given for the seven genera separately (n is the
number of sampled individuals). The best DNA barcode(s) are highlighted in grey for each genus
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 4 of 11
L. individuals and the second one grouping V. persica Poir.
and V. polita Fr. together. The four loci also agreed in clus-
tering two V. polita individuals within the V. arvensis clade
and two V. arvensis individuals in the persica-polita clade
(Additional file 3).
Therefore, monophyletic clades grouping conspecific
individuals were only observed in Lonicera with matK,
rpoB, and trnH-psbA. For the six remaining genera,
none of the chloroplast regions was successful in recon-
structing monophyletic species clades.
Locus combination and barcode gaps
Combining markers improves the rate of correct species
identification [20, 27]. In the present study, all possible
combinations between loci were analysed and are re-
ported in Table 2. Our results clearly showed that com-
bining loci is not always an advantage. For instance,
option 11, which combines all four loci, did not result in
the highest identification rate, as one might expect if
each locus was informative. The highest success in dis-
criminating closely related species was always attained
with a combination involving trnH-psbA. We stress,
however, that it is not always the same combination of
loci that gave the best results. With two loci (options 1
to 6), option 6 (rpoB+trnH-psbA) performed well for
most genera in terms of private intraspecific diversity
and number of individuals unambiguously identified.
The exceptions were Salix,Adenostyles, and Gentiana,
for whom identical or better results were obtained with
option 3 (matK+trnH-psbA). For the combinations with
three loci (options 710), the same pattern was ob-
served: whenever the number of individuals sequenced
was the same among options, the combination of matK
and trnH-psbA performed slightly better in discriminat-
ing species. The barcoding success was enhanced when
these two loci were combined, but the lower sequencing
success of matK limited its utility in this dataset.
None of the loci or combinations of loci performed
equally for the seven genera in terms of sequencing
and identification successes and no locus or combin-
ation of loci proved to be ideal for DNA barcoding.
We selected trnH-psbA alone as the DNA barcode
for Acer,Lonicera,Geranium,andVeronica,asthe
addition of other loci did not improve discrimination
of species in these four genera. This is in line with
the original concept of DNA barcoding, which advo-
cates the use of a single sequence. For Adenostyles,
combining matKandtrnH-psbA (option 3) performed
equally or better than other options while minimizing
the number of loci involved. For Salix and Gentiana,
option 9, which combined matK, rpoB, and trnH-
psbA, gave the best discriminatory results.
Barcoding gaps were evaluated by comparing the
intra- and interspecific divergences within each genus
[21]. The Kimura 2-parameter (K2P) distances were
computed for the chosen locus/combination according
to the above chosen options: trnH-psbA for Acer,Loni-
cera,Geranium, and Veronica, option 3 for Adenostyles,
and option 9 for Salix and Gentiana. Lonicera was the
only genus with a clear barcoding gap (Fig. 1). The
expected cut-off between intra- and interspecific K2P
distances was not observed in all other genera. Acer,
Geranium,Veronica, and Gentiana also tend to have
higher inter- than intraspecific distances though there is
some overlap at frequencies ranging between 9 and
20 %. Conversely, intra- and interspecific distances over-
lap completely in Veronica and Geranium.
Analyses were performed separately in each genus, so
overlaps between intra- and interspecific variation are
expected when closely related taxa are included. In our
study, the overlap between the two distributions indi-
cated that DNA barcoding with the studied chloroplast
loci is not effective for the studied genera, except Loni-
cera. Indeed, the nearest-neighbour distance (minimum
average interspecific distance) was, with the exception of
Lonicera, lower than the maximum intraspecific distance
(Fig. 2). This type of result is associated to two main
population genetic factors, incomplete lineage sorting
and interspecific hybridization [21, 28, 44]. Recently
diverged species are likely to have a null or very low
average sequence distance to the most closely related
species. Moreover, hybridization events associated with
chloroplast captures tend to maximize the intraspecific
divergence, as divergent chloroplasts can be exchanged
and shared among species [28]. This seems to be the
case in Geranium,Gentiana, and Veronica.
Influence of sampling size
The extent to which large sample sizes influenced the cap-
ture of intraspecific variability was analysed. The correl-
ation between sampling sizes and number of unique
sequences was only found for Adenostyles (r=0.99, n=3,
P< 0.05). The lack of correlation was observed for the ma-
jority of the genera, within genera (n=35) and overall
(r=0.21, n=27, P> 0.31). We employed the rarefaction
method to quantify the average number of different se-
quences that would be recovered using a small sampling
size within species. For a sampling size of three individ-
uals, the sequence richness (Rs) ranged between one for
species with no intraspecific diversity for the studied loci
(Acer monspessulanum, A. platanoides, Adenostyles leuco-
phylla DC., Gentiana acaulis, Geranium columbinum,
Lonicera nigra,andVeronica persica)and2.5(Gentiana
clusii,Salix herbacea Schrenk, and S. reticulata L.). Inter-
estingly, the most variable species never reached Rs =3,
despite having six to nine sequences. Similarly, other spe-
cies that displayed moderate variation (two to three unique
sequences) had very low Rs values (Acer pseudoplatanus Rs
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 5 of 11
Fig. 1 Relative distribution of intra- (light grey) and interspecific (dark grey) divergence, as measured by the K2P distance, of the defined DNA
barcode alone or combined within each genus
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 6 of 11
=1.2 and Gentiana angustifolia Rs =1.1). Rs was calculated
using observed sequence frequencies, emphasizing the fact
that small samples will often miss rare sequences.
Median joining networks and life histories
Median joining networks were drawn with the selected
barcode for each genus separately (Fig. 3) and illustrate
why barcoding gaps were seldom observed. Sister species
shared the same sequences in six out of the seven genera.
Lonicera was the only genus for which complete lineage
sorting was observed. According to the most recent phyl-
ogeny of the genus, the four species analysed here belong
to separate subclades of the Lonicera clade [45]. However,
these four sections were poorly supported, so it would be
interesting to analyse the DNA barcoding performance if
one had considered species from the same subclade.
Among the other genera, three different processes can
explain the sharing of chloroplast sequences between spe-
cies. The first is incomplete lineage sorting among sister
species, which is observed in four genera. 1) Within Acer
this pattern occurs twice: between A. campestre and A.
platanoides (sequence B) and between A. monspessula-
num and A. opalus (sequence F). The latest Acer phyl-
ogeny [46] confirms that these two pairs are sister species.
2) Within Gentiana, the AAA sequences are shared be-
tween G. angustifolia and G. acaulis. According to Christe
et al. [37], who studied the phylogeographic patterns
within the Ciminalis group, these species have diverged
recently. 3) Within Veronica,V. persica and V. polita share
sequence A. The latest phylogeny, based on ITS, reported
that these are sister species within subgenus Pocilla [47].
4) Adenostyles alliariae Kern and A. leucophylla share se-
quence A, which reflects their status as sister species [48].
The second process that explains sequence sharing is pu-
tative hybridization between species. This is observed
within both genera of annual plants: Veronica polita is
characterized by sequences A, G, and C. Sequence C, which
is distinct by 66 mutations from the two others, is
shared with V. arvensis. Hybridization is recognized as
an important evolutionary force for some subgenera of
Veronica [9]. In published phylogenies, the species for
which hybridization is suspected are grouped together
in the ITS consensus tree and the cladogram based on
the ITS sequences, chromosome numbers, and iridoid
composition [47, 49]. Geranium pusillum L. harbours two
different sequences that are separated by 23 mutations;
one is shared with G. dissectum L. (sequence A). A third
case of hybridization was also observed within Gentiana.
Indeed, G. alpina possesses two sequences that are dis-
tinct at 63 positions, one of which (BAJ) is closely related
to the most frequent sequence (CBC) in G. clusii. Hybrid-
izations between Gentiana species have often been re-
ported [5052], and distinct events of chloroplast capture
involving these species have also been suggested [37].
The complete lack of structure observed within Salix
was surprising, but not new. The three most common se-
quences were shared among the four species analysed in
this study, and only 14.5 % of the specimens had private
sequences. Our results agreed with a recent study that
documented little variation in chloroplast loci among
Salix species, with most taxa sharing the same barcode se-
quence. Complex processes involving recent repeated
plastid capture events, aided by widespread hybridization
and long-range seed dispersal, but primarily propelled by
one or more trans-species selective sweepswere sug-
gested to explain the observed pattern [53].
In summary, our results illustrate the effect of species
evolutionary histories on DNA barcoding success. In this
study, evolutionary history refers to recent speciation
events with incomplete lineage sorting and retention of
ancestral sequences, interspecific hybridization events
Fig. 2 Minimum average interspecific distance (light grey) against the maximum intraspecific divergence (dark grey), as measured by the K2P
distance, of the defined DNA barcode within each genus
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 7 of 11
Fig. 3 Sequence networks drawn for selected DNA barcode(s) within each genus. Each unique sequence combination is represented by a circle,
with size proportional to the number of individuals sharing the sequence. Each branch segment represents a single mutation; substitutions are
coded as full lines and indels as double lines
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 8 of 11
with chloroplast capture, and spatial expansions with se-
quence surfing [54]. It is commonly acknowledged that
several processes underlying the evolutionary patterns in
plants cause a partial failure of DNA barcodes to track
species boundaries [13, 28, 29, 43], but this study shows
that the absence of a barcoding gap among closely re-
lated species is quite common, with extensive sharing of
diversity among species (49 %).
Conclusions
The main factor that impacts DNA barcoding success is a
speciesevolutionary history. Sampling many specimens
from a wide geographical distribution within species was
shown to be important as it increases the likelihood of cap-
turing the intraspecific genetic variation. However, sam-
pling sizes were not correlated to the number of different
sequences found within a species, because variability is
mostly influenced by the speciesevolutionary history. Our
study shows that within the same genus, and even within
the same section, sequence variation can range from low to
high, depending on the species (for instance, Gentiana clu-
sii and G. acaulis 11 and 2 different sequences, respect-
ively, with similar sampling sizes collected from the whole
distribution range). Such diverse patterns were obtained
through different demographic regimes (bottlenecks, spatial
or demographic expansions) that shaped the diversity and
its structuring.
Life traits, such as generation time, do not influence
the DNA barcode success in our study. The best and
worst identification successes were indeed found for
the two tree genera (Acer, 64 % success and Salix,86%
failure). The annual plants analysed here showed, on
average, a higher number of mutations between se-
quences than was observed in perennials. This should,
theoretically, be an advantage for DNA barcoding suc-
cess, but the incidence of interspecific hybridization
within these genera highly shapes the observed genetic
pattern and results in specimen identification failures.
Therefore, our results underline the impact of species
evolutionary histories on the ability to successfully
identify a given specimen.
We found that the most useful combination of loci for
discriminating closely related species can differ from one
genus to another, and this agrees with other papers that
discussed the interest of different loci as DNA barcodes.
However, our results demonstrated that trnH-psbAisal-
most always the best DNA barcode locus. This supports
the proposal for trnH-psbA to be added to the two core
DNA chloroplast barcodes proposed by the CBOL plant
working group. Moreover, our results show that the K2P
metric is not the most appropriate, as it does not take into
account invertion/deletion events that are of high interest,
especially for trnH-psbA, to distinguish and document
sequence variation.
Methods
Sampling strategy
Genera and species were selected for the present study
based on the following criteria: generation times, geo-
graphic distribution ranges, clear taxonomical status,
and ease of recognition. In each case, all possible closely
related species were sampled except any rare or endan-
gered ones. Species of two genera are trees (Acer and
Salix), three genera include perennial herbaceous or
woody species (Adenostyles,Gentiana, and Lonicera),
whereas two genera include annual species (Geranium
and Veronica). For each species, as many localities as
possible were sampled, over the largest possible geo-
graphical range, to gather as much intraspecific variation
as possible (Additional file 1). For each individual, an
herbarium voucher was collected, identified by an ex-
pert, and deposited at the Geneva herbarium (G). For
protected Gentiana species, high-quality photos were
taken in lieu of herbarium specimens.
DNA extraction, amplification and sequencing
Total genomic DNA was extracted using the NucleoS-
pin© Plant II kit (Macherey-Nagel, GmbH & Co. KG,
Düren, Germany) following the suppliers instructions.
Three cpDNA coding regions (matK, rpoC1, and rpoB)
and one cpDNA spacer (trnH-psbA) were amplified and
sequenced. PCR was performed in 20 μL total volume
with 0.60 U Taq (Roche, Mannheim, Germany), 2 μLof
10X buffer containing 20 mM MgCl
2
, 0.8 μL of each pri-
mer (10 mM), 0.4 μl of a mix containing 10 mM of each
dNTP (Roche), and 0.85 μL of template DNA of un-
known concentration. The PCR program had an initial
heating step at 95 °C for 6 min, followed by 35 cycles of
denaturation at 95 °C for 30 s, annealing for 30 s at a
locus-specific temperature, elongation at 72 °C for 45 s,
and a final elongation step at 72 °C for 10 min. Anneal-
ing temperatures varied between 45 and 52 °C depend-
ing on locus and species (see Additional file 2 for
details). The primers used are also listed in Additional
file 2. PCR products were cleaned and bidirectionally
sequenced using the PCR primers on an ABI 377 auto-
mated sequencer (Applied Biosystems, Foster City, CA,
USA) following the manufacturers protocols.
Sequence alignment and data analyses
Contig assembly and sequence consensus were gener-
ated using Sequencher (GeneCodes Corporation, Ann
Arbor, Michigan, USA). Barcode sequences were aligned
in BIOEDIT 7.0.3.5 [55] and edited manually. Sequence
variation was then characterized using the percentage of
conserved sites, the percentage of parsimony informative
sites, and the number of unique sequences per species.
This last measure is the only one that takes into account
insertion/deletion and inversion events. Both events
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 9 of 11
were manually coded as single mutation steps [56]. Se-
quence variation analyses were then performed in MEGA
version 6 [57]. All sequences were deposited in GenBank
under accession numbers KU672731KU674305 and
KU672731KU674305 (Additional file 4).
In order to investigate how well the different markers
performed individually in identifying species within a
genus, the number of sequences that were private to a sin-
gle species was checked and the number of individuals
unambiguously identified was reported. We also per-
formed a comparison of all possible locus-combinations.
Species discrimination was evaluated using tree-
based analyses. The Neighbour-Joining tree recon-
struction recommended as the standard barcoding
method [1] was adopted and performed with SeaView
4.4.0, based on the K2P model and 100 replicates for
bootstrap analyses [58].
The presence of barcoding gaps was analysed by
graphing the distributions of intra- and interspecific gen-
etic distances for each genus. Sequence divergences were
calculated using pairwise distances with the Kimura 2-
parameter in MEGA [57].
The correlation between sampling size and the
number of unique sequences was computed overall
species and within genera for trnH-psbA, which was
the most diverse barcode within species and the only
one common to all genera. The sequence richness
(Rs) was computed for a sample size of three individ-
uals, using the rarefaction methods that takes into ac-
count sequence frequencies in each species [59]. Rs
was used to quantify the average number of different
sequences that would be recovered using a sampling
size of three individuals within species. Correlations
and Rs were computed in Excel and confidence inter-
vals for correlation coefficients were assessed in the
online program VassarStat (http://vassarstats.net/)
using the Fisher r-to-z transformation.
Median joining networks of the sequences were drawn
using the program Network [60]. These analyses were
performed, within each genus, on the defined DNA
barcode alone or combined: trnH-psbA alone for Acer,
Lonicera,Geranium,andVeronica; option 3 (matK and
trnH-psbA) for Adenostyles, and option 9 (matK, rpoB,
and trnH-psbA) for Salix and Gentiana. Site mutations
and indels were equally weighted and all the structural
mutations (inversions and insertions/deletions of more
than 1 bp) were treated as single-step events.
Availability of data and materials
The datasets supporting the conclusions of this article
are available in the Genbank repository, [accession num-
bers KU672731KU674305 and KU672731KU674305
at http://www.ncbi.nlm.nih.gov/genbank/].
Additional files
Additional file 1: List of individuals. For each individual, the country,
district, and absolute coordinates in decimal degrees are given.
(PDF 29 kb)
Additional file 2: Amplification conditions. For each species, annealing
temperature and the primer combination are given for matK, rpoC1, rpoB,
and trnH-psbA. (PDF 27 kb)
Additional file 3: Phylogenetic trees. For each genus and locus, a
neighbour joining tree is presented. Bootstrap values above 80 % are
shown above branches. Codes following species names are individual
numbers (see Additional file 1). (PDF 572 kb)
Additional file 4: Genbank accession numbers. For each sample,
accession numbers for the four loci are given. (PDF 71 kb)
Competing interests
The authors declare that they have no competing interests.
Authorscontributions
Both authors conceived the experiment, conducted field surveys and wrote
the manuscript, SCW performed molecular and statistical analyses. Both
authors read and approved the final manuscript.
Acknowledgements
We wish to thank David Aeschimann, Beat Baumler, Ariane Cailliau, Camille
Christe, Daniel Jeanmonod, Philippe Juillierat, Catherine Lambelet, and Pierre-
André Loizeau for their help in sampling; David Aeschimann for specimen
identifications; Régine Niba and Camille Christe for their help with lab work,
and Slim Chraiti for his help with figures. We are extremely grateful for the
assistance provided by one of the reviewers to improve this manuscript. The
Swiss National Foundation supported this project with grant n°3100A0-
120390 to YN. We are also grateful to the Société Académique de Genève
and to SwissBOL (Swiss Barcode of Life) for their financial support to YN and
SCW, respectively.
Author details
1
Laboratoire de Systématique Végétale et Biodiversité, Conservatoire et
Jardin botaniques & University of Geneva, Chemin de lImpératrice, 1, 1292,
Chambésy, Geneva, Switzerland.
2
SwissBOL, University of Geneva,
Department of Genetics and Evolution, Quai Ernest Ansermet 30, 1211
Geneva, Switzerland.
Received: 8 January 2016 Accepted: 5 May 2016
References
1. Hebert PDN, Cywinska A, Ball SL, DeWaard JR. Biological identifications
through DNA barcodes. Proc RoySoc B Biol Sci. 2003;270:31321.
2. Kress WJ, Erickson DL. A two-Locus global DNA barcode for land plants: The
coding rbcL gene complements the non-coding trnH-psbA spacer region.
PLoS One. 2007;2, e508.
3. Spriggs EL, Christin P-A, Edwards EJ. C4 photosynthesis promoted species
diversification during the Miocene grassland expansion. PLoS One. 2014;
9(5), e97722.
4. Chen SL, Yao H, Han JP, Liu C, Song JY, Shi LC, et al. Validation of the ITS2
region as a novel DNA barcode for identifying medicinal plant species. PLoS
One. 2010;5, e8613.
5. Yao H, Chen SL, Song JY, Liu C, Ma XY, Luo K, et al. Testing candidate plant
barcode regions in the Dendrobium species. Planta Med. 2009;75:9300.
6. Song JY, Yao H, Li Y, Li XW, Lin YL, Liu C, et al. Authentication of the family
Polygonaceae in Chinese pharmacopoeia by DNA barcoding technique.
J Ethnopharm. 2009;124:4349.
7. Hollingsworth ML, Andra Clark A, Forrest LL, Richardson J, Pennington RT,
Long DG, et al. Selecting barcoding loci for plants: Evaluation of seven
candidate loci with species-level sampling in three divergent groups of land
plants. Mol Ecol Res. 2009;9:43957.
8. Lahaye R, Van der Bank M, Bogarin D, Warner J, Pupulin F, Gigot G, et al.
DNA barcoding the floras of biodiversity hotspots. Proc Natl Acad Sci U S A.
2008;105:29238.
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 10 of 11
9. Bardy KE, Schönswetter P, Schneeweiss GM, Fischer MA, Albach DC.
Extensive gene flow blurs species boundaries among Veronica barrelieri,V.
orchidea and V. spicata (Plantaginaceae) in southeastern Europe. Taxon.
2011;60:10821.
10. Wang W, Wu Y, Yan Y, Ermakova M, Kerstetter R, Messing J. DNA barcoding
of the Lemnaceae, a family of aquatic monocots. BMC Pl Biol. 2010;10:205.
11. CBOL Plant Working Group. A DNA barcode for land plants. Proc Natl Acad
Sci U S A. 2009;106:127947.
12. Wang M, Zhao HX, Wang L, Wang T, Yang RW, Wang XL, et al. Potential use
of DNA barcoding for the identification of Salvia based on cpDNA and
nrDNA sequences. Gene. 2013;528:20615.
13. Hollingsworth PM, Graham SW, Little DP. Choosing and using a plant DNA
barcode. PLoS One. 2011;6, e19254.
14. Chase MW, Salamin N, Wilkinson M, Dunwell JM, Kesanakurthi RP, Haidar N,
et al. Land plants and DNA barcodes: short-term and long-term goals.
Philos Trans R Soc Lond B Biol Sci. 2005;360:188995.
15. Chase MW, Fay MF. Barcoding of plants and fungi. Ecology. 2009;325:6823.
16. Kress WJ, Wurdack KJ, Zimmer EA, Weigt LA, Janzen DH. Use of DNA barcodes
to identify flowering plants. Proc Natl Acad Sci U S A. 2005;102:836974.
17. Luo K, Chen S, Chen K, Song J, Yao H, Ma X, et al. Assessment of candidate
plant DNA barcodes using the Rutaceae family. SciChina Life Sci. 2010;53:
7018.
18. Pang X, Song J, Zhu Y, Xie C, Chen S. Using DNA barcoding to identify
species within Euphorbiaceae. Planta Med. 2010;76:17846.
19. Sass C, Little DP, Stevenson DW, Specht CD. DNA barcoding in the
Cycadales: testing the potential of proposed barcoding markers for species
identification of cycads. PLoS One. 2007;2, e1154.
20. Fazekas AJ, Burgess KS, Kesanakurti PR, Graham SW, Newmaster SG,
Husband BC, et al. Multiple multilocus DNA barcodes from the plastid
genome discriminate plant species equally well. PLoS One. 2008;3:112.
21. Meyer CP, Paulay G. DNA barcoding: Error rates based on comprehensive
sampling. PLoS Biol. 2005;3:222938.
22. Zhang AB, He LJ, Crozier RH, Muster C, Zhu CD. Estimating sample sizes for
DNA barcoding. Mol Phyl Evol. 2010;54:10359.
23. Arakaki M, Christin P-A, Nyffeler R, Lendel A, Eggli U, Ogburn RM, et al.
Contemporaneous and recent radiations of the worlds major succulent
plant lineages. Proc Natl Acad Sci U S A. 2011;108:837984.
24. Christin PA, Spriggs E, Osborne CP, Stromberg CAE, Salamin N, Edwards EJ.
Molecular dating, evolutionary rates, and the age of the grasses. Sys Biol.
2014;63:15365.
25. Hughes C, Eastwood R. Island radiation on a continental scale: Exceptional
rates of plant diversification after uplift of the Andes. Proc Natl Acad Sci U S
A. 2006;103:103349.
26. de Vos JM, Hughes CE, Schneeweiss GM, Moore BR, Conti E. Heterostyly
accelerates diversification via reduced extinction in primroses. Proc R Soc B
Biol Sci. 2014;281:178490.
27. Gao T, Yao H, Song J, Zhu Y, Liu C, Chen S. Evaluating the feasability of
using candidate DNA barcodes in discriminating species of the large
Asteraceae family. BMC Evol Biol. 2010;10:324.
28. Naciri Y, Linder HP. Species delimitation and relationships: the dance of the
seven veils. Taxon. 2015;64:316.
29. Spooner DM. DNA barcoding will frequently fail in complicated groups: An
example in wild potatoes. Amer J Bot. 2009;96:117789.
30. Newmaster SG, Ragupathy S. Ethnobotany genomics - use of DNA
barcoding to explore cryptic diversity in economically important plants. Ind
J Sci Technol. 2009;2:28.
31. Arnold ML. Natural hybridization and evolution. New York: Oxford University
Press; 1997.
32. Soltis PS, Soltis DE. The rôle of hybridization in plant speciation. Annu Rev
Plant Biol. 2009;60:56188.
33. Adams KL, Wendel JF. Polyploidy and genome evolution in plants. Curr
Opin Plant Biol. 2005;8:13541.
34. Jiao N, Wickett NJ, Ayyampalayam S, Chanderbali AS, Landherr L, Ralph PE,
et al. Ancestral polyploidy in seed plants and angiosperms. Nature. 2011;
473:97100.
35. Petit RJ, Hampe A. Some evolutionary consequences of being a tree. Annu
Rev Ecol Evol Syst. 2006;37:187214.
36. Smith SA, Donoghue MJ. Rates of molecular evolution are linked to life
history in flowering plants. Science. 2008;322:869.
37. Christe C, Caetano S, Aeschimann D, Kropf M, Diadema K, Naciri Y. The
intraspecific genetic variability of siliceous and calcareous Gentiana species
is shaped by contrasting demographic and re-colonization processes. Mol
Phyl Evol. 2014;70:32336.
38. Gere J, Yessoufou K, Daru BH, Mankga LT, Maurin O, van der Bank M.
Incorporating trnH-psbA to the core DNA barcodes improves significantly
species discrimination within southern African Combretaceae. ZooKeys.
2013;365:12947.
39. Pang X, Luo H, Sun C. Assessing the potential of candidate DNA barcodes
for identifying non-flowering seed plants. Plant Biol. 2012;14:83944.
40. Telford A, OHare MT, Cavers S, Holmes N. Can genetic barcoding be used
to identify aquatic Ranunculus L. subgenus Batrachium (DC) A. Gray? A test
using some species from the British Isles. Aquatic Bot. 2011;95:6570.
41. Sun XQ, Bai MM, Yao H, Guo JL, Li MM, Hang YY. DNA barcoding of
populations of Fallopia multiflora, an indigenous herb in China. Genet Mol
Res. 2013;12:407889.
42. Storchova H, Olson MS. The architecture of the chloroplast psbA-trnH non-
coding region in angiosperms. Plant Syst Evol. 2007;268:23556.
43. Fazekas AJ, Kesankurti PR, Burgess KS, Percy DM, Graham SW, Barrett SC, et
al. Are plant species inherently harder to discriminate than animal species
using DNA barcoding markers? Mol Ecol Res. 2009;9:1309.
44. Moritz C, Cicero C. DNA barcoding: promise and pitfalls. PLoS Biol. 2004;2:
152931.
45. Smith SA. Taking into account phylogenetic and divergence-time
uncertainty in a parametric biogeographical analysis of the Northern
Hemisphere plant clade Caprifolieae. J Biogeog. 2009;36:232437.
46. Renner SS, Grimm GW, Schneeweiss GM, Stuessy TF, Ricklefs RE. Rooting and
dating maples (Acer) with an uncorrelated-rates molecular clock: Implications
for North American/Asian disjunctions. Syst Biol. 2008;57:795808.
47. Taskov RM, Albach DC, Grayer RJ. Phylogeny of Veronica - a combination of
molecular and chemical evidence. Plant Biol. 2004;6:67382.
48. Dillenberger MS, Kadereit JW. The Phylogeny of the European high
mountain genus Adenostyles (Asteraceae-Snecioneae) reveals that edaphic
shifts coincide with dispersal events. Amer J Bot. 2013;100:117183.
49. Muñoz LM, Albach DC, Sánchez-Agudo JA, Martínez-Ortega MM. Systematic
Significance of Seed Morphology in Veronica (Plantaginaceae): A
Phylogenetic Perspective. Ann Bot. 2006;98:33550.
50. Wenk B. Maintenance of species boundaries between Gentiana acaulis and
G. clusii: asymmetry in gametic barriers. Zurich, Switzerland: Institute of
Systematic Botany, University of Zürich; 2008.
51. Hungerer KB, Kadereit JW. The phylogeny and biogeography of Gentiana L.
sect. Ciminalis (Adans.) Dumort.: A historical interpretation of distribution
ranges in the European high mountains. Persp Plant Ecol Evol Syst. 1998;1:
12135.
52. von Hagen KB, Kadereit JW. Notes on the systematics and evolution of Gentiana
sect. Ciminalis Bot JahrbSyst Pflanzengesh Pflanzengeog. 2000;122:30539.
53. Percy DM, Argus GW, Cronk QC, Fazekas AJ, Kesanakurti PR, Burgess KS, et
al. Understanding the spectacular failure of DNA barcoding in willows
(Salix): Does this result from a trans-specific selective sweep? Mol Ecol. 2014;
23:473756.
54. Klopfstein S, Currat M, Excoffier L. The fate of mutations surfing on the wave
of a range expansion. Mol Biol Evol. 2006;23:48290.
55. Hall TA. BioEdit: a user-friendly biologicalsequence alignment editor and
analysis programfor Windows 95/98/NT. Nucl Acids Symp Ser. 1999;41:958.
56. Barriel V. Phylogénies moléculaires et insertions-délétions de nucléotides.
Comptes rendus lAcad Sci Paris Sci Vie. 1994;317:693701.
57. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6. Molecular
Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol. 2013;30:27259.
58. Gouy M, Guindon S, Gascuel O. SeaView version 4: a multiplatform graphical
user interface for sequence alignment and phylogenetic tree building. Mol
Biol Evol. 2010;27:2214.
59. Petit RJ, El Mousadik A, Pons O. Identifying populations for conservation on
the basis of genetic markers. Cons Biol. 1998;12:84455.
60. Bandelt H-J, Forster P, Röhl A. Median-joining networks for inferring
intraspecific phylogenies. Mol Biol Evol. 1999;16:3748.
Caetano Wyler and Naciri BMC Evolutionary Biology (2016) 16:103 Page 11 of 11
... Such authentication at the DNA level is reproducible, sensitive and specific enough to allow the determination and differentiation of plant species. The Consortium for the Barcode of Life (CBOL) plant working group recommended a two-locus combination of rbcL and matK as plant barcodes (Caetano Wyler and Naciri, 2016;Hollingsworth et al., 2011). Other research groups propose the use of additional markers (Caetano Wyler and Naciri, 2016). ...
... The Consortium for the Barcode of Life (CBOL) plant working group recommended a two-locus combination of rbcL and matK as plant barcodes (Caetano Wyler and Naciri, 2016;Hollingsworth et al., 2011). Other research groups propose the use of additional markers (Caetano Wyler and Naciri, 2016). In the present study, the molecular identification of mistletoe was performed using four loci, matK, rbcL, psbA-trnH, from the chloroplast genome and ITS (internal transcribed spacer), from the nuclear genome. ...
Article
Ethnopharmacological relevance Viscum album L. (European mistletoe), a member of the Santalaceae, is a hemiparasitic, evergreen shrub growing on deciduous and coniferous trees. In traditional and folk medicine, mistletoe was used for the treatment of central nervous system disorders such as epilepsy, hysteria, insomnia, nervous excitability, neuralgia, headache, dizziness and fatigue. However, relatively little is known of its neuropharmacological activity. Aim of the study: The aim of the present study was to evaluate the effect of treatment with aqueous and hydroethanolic extracts from Viscum album L. parasitizing birch, linden and pine, on MAO-A and MAO-B activity as well as serotonin, dopamine and serotonin receptor 5-HTR1A levels in Galleria mellonella (Lepidoptera) larvae. Materials and methods The phytochemical composition of the extracts was characterised using UPLC-DAD-ESI-MS/MS. To investigate the neuropharmacological activity of Viscum album L. extracts, Galleria mellonella (Lepidoptera) larvae were used as a model organism. The inhibitory potential of the extracts against MAO-A and MAO-B was determined by fluorometry. The serotonin, dopamine and serotonin receptor 5-HTR1A levels in larvae hemolymph after treatment were quantified by ELISA. Results UPLC-DAD-ESI-MS/MS analysis allowed the identification of 88 compounds, either full or in part. Most of the characterised phytochemicals were flavonoids, hydroxycinnamic acids and lignans. Screening found that aqueous and hydroethanolic mistletoe extracts inhibited the enzymatic activity of either MAO-A or MAO-B or both. Additionally, mistletoe extract administration increased the levels of serotonin and serotonin receptor 5-HTR1A. None of the tested extracts had any significant effect on dopamine level. Conclusions A key novel finding was that the aqueous and hydroethanolic extracts from Viscum album L. inhibited monoamine oxidase activity and increased the levels of serotonin and serotonin receptor 5-HTR1A in Galleria mellonella (Lepidoptera) larvae. These properties may be due to the presence of phenolic constituents, particularly flavonoids. Further research based on bioassay-guided fractionation of mistletoe is needed to identify CNS-active molecules.
... The current investigation discovered that matK, rbcL, and trnH-psbA are trustworthy identification markers and that the DNA identification technique successfully distinguished all samples up to the species level (Caetano Wyler and Naciri, 2016;Naim and Mahboob, 2020). DNA can be amplified and sequenced using this technique with great accuracy, and plastid DNA plays a key role in establishing evolutionary connections and defining plant species boundaries (Srivastava et al., 2022). ...
Article
Cannabis sativa possess substantial amount of omega-3 fatty acids and several medicinal benefits with potential secondary metabolites. This study employed rbcL, matK, and trnH-psbA barcode markers and RAPD marker to explore genetic diversity in C. sativa populations in Punjab, Pakistan. In population genetic study using DNA barcode markers, cladogram showed investigated samples were genetically similar; however, demonstrated significant genetic diversity with the available DNA sequences of this plant from other regions and countries. The nucleotide sequence variation (% divergence) indicated high intraspecific variation in plants of different geographical zones. However, the RAPD marker analysis displayed genetic diversity among collected Cannabis sativa samples. The principal component analysis reinforced these findings by providing a spatial representation of genetic distances among the samples that revealed a considerable genetic diversity among the collected material and significant intraspecific variation was observed, indicating diverse genetic profiles across geographical regions. This research highlights the potential of molecular markers in germplasm selection for breeding programs.
... There is no report available on DNA barcoding of SA, yet, but one report on Salix babylonica L. was published [11,21]. It has been reported that Salix species like Salix herbacea Schrenk are imperceptible with matK gene [30]. The combination of matK and rbcL gene may be a better tool to identify the Salix species [16]. ...
Article
Full-text available
... The efficacy of a marker in delineating species boundaries also depends on the existence of a clear-cut difference between inter-and intraspecific variability [39]. The data generated in this study demonstrated high intraspecific conservation of the ITS-1 marker among T. nativa and Trichinella T6 isolates from Canadian wildlife. ...
Article
Full-text available
Background Amplicon-based next-generation sequencing (NGS) has rapidly gained popularity as a powerful method for delineating taxa in complex communities, including helminths. Here, we applied this approach to identify species and genotypes of zoonotic nematodes of the Trichinella genus. A known limitation of the current multiplex PCR (mPCR) assay recommended by the International Commission on Trichinellosis is that it does not differentiate Trichinella nativa from T. chanchalensis. Methods The new assay entails deep sequencing of an amplified variable fragment of the ribosomal cistron's (rDNA) internal transcribed spacer 1 using the Illumina platform. The assay was evaluated using first-stage larvae (L1) of select laboratory strains of various Trichinella taxa mixed in known proportions and then validated using archived L1 from 109 wildlife hosts. The species/genotypes of these L1 isolates from wildlife were previously determined using mPCR. Results NGS data analysis for Trichinella laboratory strains selected as representative of North American fauna revealed a sequence representation bias. Trichinella pseudospiralis, a non-encapsulated species, was the most underrepresented when mixed with T. spiralis, T. murrelli, T. nativa and Trichinella T6 in equal quantities. However, five L1 of T. pseudospiralis were readily revealed by NGS in a mix with 2000 L1 of T. nativa (1:400 ratio). From naturally infected wildlife, all Trichinella taxa revealed by mPCR were also identified by NGS in 103 of 107 (96.3%) samples amplified on both assays. NGS identified additional taxa in 11 (10.3%) samples, whereas additional taxa were revealed by mPCR in only four (3.7%) samples. Most isolates comprised single or mixed infections of T. nativa and Trichinella T6. On NGS, T. chanchalensis (T13) was detected in combination with Trichinella T6 in a wolverine (Gulo gulo) and in combination with T. nativa and Trichinella T6 in a marten (Martes americana) from the Northwest Territories, Canada. Conclusions This new NGS assay demonstrates strong potential as a single assay for identifying all recognised Trichinella taxa as well as improved sensitivity for detecting under-represented and novel genotypes in mixed infections. In addition, we report a new host record for T. chanchalensis in American marten. Graphical Abstract
... There is no report available on DNA barcoding of SA, yet, but one report on Salix babylonica L. was published [11,21]. It has been reported that Salix species like Salix herbacea Schrenk are imperceptible with matK gene [30]. The combination of matK and rbcL gene may be a better tool to identify the Salix species [16]. ...
Article
There is need to search substitution of stem bark with leaf of Salix alba L, which is yet not reported. Comparative phytochemicals viz. Salicin, Procyanidin B1 and Catechin were quantified in the heart wood (SA-HW), stem bark (SA-SB) and leaves (SA-L) of Salix alba L.by using newly developed HPLC method. Essential metal ions of all three parts were analysed using newly developed ICP-OES method, where SA-L were founded as rich source of micronutrients and essential metal ions as compared to SA-SB and SA-HW. GC-MS/MS analysis has shown the presence of fatty acids and volatile compounds. The observed TPC and TFC values for all three parts were ranged from 2.69 to 32.30 mg GAE/ g of wt. and 37.57 to 220.76 mg QCE/ g of wt. respectively. In DPPH assay the IC50 values of SA-SB, SA-HW, and SA-L were 1.09 (± 0.02), 5.42 (± 0.08), and 8.82 (± 0.10) mg/mL, respectively. The order of antibacterial activities against E. coli, S. aureus, P. aeruginosa, and B. subtilis strains was SA-L > SA-HW > SA-SB with strong antibacterial activities against S. aureus, and B. subtilis strains. The antacid activities order was SA-L > SA-SB > SA-HW.
... Adequately identifying different lineages in the fraterculus group, and especially in the A. fraterculus complex, is particularly challenging due to ILS and introgression (Liu et al., 2017;Lopez-Vaamonde et al., 2021). In such cases, multi-locus DNA barcoding has been shown to be effective in species identification even with some levels of interspecific gene flow (Liu et al., 2017 (Cardeñosa et al., 2020;Gonçalves et al., 2015;Hawlitschek et al., 2016;Serite et al., 2021), as well as plants such as the Solanum section Petota (wild potatoes) and species of the genera Salix (trees) and Veronica (speedwell) (Spooner, 2009;Wyler & Naciri, 2016). ...
Article
Full-text available
Insect pests cause tremendous impact to agriculture worldwide. Species identification is crucial for implementing appropriate measures of pest control but can be challenging in closely related species. True fruit flies of the genus Anastrepha Schiner (Diptera: Tephritidae) include some of the most serious agricultural pests in the Americas, with the Anastrepha fraterculus (Wiedemann) complex being one of the most important due to its extreme polyphagy and wide distribution across most of the New World tropics and subtropics. The eight morphotypes described for this complex as well as other closely related species are classified in the fraterculus species group, whose evolutionary relationships are unresolved due to incomplete lineage sorting and introgression. We performed multifaceted phylogenomic approaches using thousands of genes to unravel the evolutionary relationships within the A. fraterculus complex to provide a baseline for molecular diagnosis of these pests. We used a methodology that accommodates variable sources of data (transcriptome, genome, and whole‐genome shotgun sequencing) and developed a tool to align and filter orthologs, generating reliable datasets for phylogenetic studies. We inferred 3031 gene trees that displayed high levels of discordance. Nevertheless, the topologies of the inferred coalescent species trees were consistent across methods and datasets, except for one lineage in the A. fraterculus complex. Furthermore, network analysis indicated introgression across lineages in the fraterculus group. We present a robust phylogeny of the group that provides insights into the intricate patterns of evolution of the A. fraterculus complex supporting the hypothesis that this complex is an assemblage of closely related cryptic lineages that have evolved under interspecific gene flow. Despite this complex evolutionary scenario, our subsampling analysis revealed that a set of as few as 80 loci has a similar phylogenetic resolution as the genome‐scale dataset, offering a foundation to develop more efficient diagnostic tools in this species group.
... This gap marks the limit between minimum interspecific and maximum intraspecific divergence. Thus, it is crucial to ensure the distance-based method's effectiveness [51,81]. Our results showed that the recursive partitions in ABGD recognized more OTUs than primary ones, exhibiting a higher accuracy in species resolution under the analysis, which corroborates with previous observations [51,82,83]. ...
Article
Full-text available
The need for herbal medicinal plants is steadily increasing. Hence, the accurate identification of plant material has become vital for safe usage, avoiding adulteration, and medicinal plant trading. DNA barcoding has shown to be a valuable molecular identification tool for medicinal plants, ensuring the safety and efficacy of plant materials of therapeutic significance. Using morphological characters in genera with closely related species, species delimitation is often difficult. Here, we evaluated the capability of the nuclear barcode ITS2 and plastid DNA barcodes rbcL and matK to identify 20 medicinally important plant species of Caryophyllales. In our analysis, we applied an integrative approach for species discrimination using pairwise distance-based unsupervised operational taxonomic unit “OTU picking” methods, viz., ABGD (Automated Barcode Gap Analysis) and ASAP (Assemble Species by Automatic Partitioning). Along with the unsupervised OTU picking methods, Supervised Machine Learning methods (SML) were also implemented to recognize divergent taxa. Our results indicated that ITS2 was more successful in distinguishing between examined species, implying that it could be used to detect the contamination and adulteration of these medicinally important plants. Moreover, this study suggests that the combination of more than one method could assist in the resolution of morphologically similar or closely related taxa.
... In contrast, three plastid loci (ycf1, matK-trnG-GCC, and rpoC1) showed better resolution for identifying the species in Fritillaria. Among these regions, ycf1 has been proposed as the most potential DNA barcode from chloroplast genome for land plants (Dong et al., 2015), but rpoC1 provided low resolution in many groups due to short DNA sequences (CBOL Plant Working Group, 2009;Wyler and Naciri, 2016;Xie et al., 2019). Moreover, matK-trnG-GCC included more variable sites ( Table 2). ...
Article
Full-text available
The Fritillaria is an extremely complicated genus in taxonomy and phylogeny, which contains numerous medicinal species in China. Both traditional characteristic-based taxonomy and universal DNA barcodes (ITS, trnH-psbA, and rbcL) are difficult to effectively identify the species. Here, we generated a large dataset of chloroplast genomes from multiple accessions per species of Fritillaria to evaluate their effectiveness in species discrimination. Moreover, phylogeny of species in China was explored based on the complete chloroplast genomes, and then divergence times of each node were estimated. The results showed that all 21 species in Fritillaria here (including two suspicious species) could be correctly discriminated using cpDNA genomes except F. cirrhosa, which suggested that DNA super-barcode could greatly enhance species discriminatory resolution for complicated genera. Furthermore, four regions (ycf1, matK-trnG-GCC, rpoC1, and matK) gained remarkably higher resolution than that of other plastid regions, but only matK might be suitable to identify Fritillaria species in consideration of its lengths. Phylogenomic analysis showed that the subgenus Fritillaria in China was divided into four major clades with obvious geographic structure. Among them, Clade I, mainly distributed in southwest China, was a young and complicated group. Moreover, according to the analysis, taxonomic treatments of the two suspicious species, namely “F. omeiensis” and “F. hupehensis” in Flora of China (2000) are questionable and might need further revision. Molecular dating revealed that both origin and divergence of subgenus Fritillaria, as well as its four major clades, were significantly associated with geological and climatic fluctuations during the Middle to Late Miocene. This study would enrich case studies of DNA super-barcode and provide new insights on speciation, lineage diversification, and biogeography of the Fritillaria in China.
Article
Full-text available
Many emerging invasive weeds display rapid adaptation against different stressful environments compared to their natives. Rapid adaptation and dispersal habits helped invasive populations have strong diversity within the population compared to their natives. Advances in molecular marker techniques may lead to an in-depth understanding of the genetic diversity of invasive weeds. The use of molecular techniques is rapidly growing, and their implications in invasive weed studies are considered powerful tools for genome purposes. Here, we review different approach used multi-omics by invasive weed studies to understand the functional structural and genomic changes in these species under different environmental fluctuations, particularly, to check the accessibility of advance-sequencing techniques used by researchers in genome sequence projects. In this review-based study, we also examine the importance and efficiency of different molecular techniques in identifying and characterizing different genes, associated markers, proteins, metabolites, and key metabolic pathways in invasive and native weeds. Use of these techniques could help weed scientists to further reduce the knowledge gaps in understanding invasive weeds traits. Although these techniques can provide robust insights about the molecular functioning, employing a single omics platform can rarely elucidate the gene-level regulation and the associated real-time expression of weedy traits due to the complex and overlapping nature of biological interactions. We conclude that different multi-omic techniques will provide long-term benefits in launching new genome projects to enhance the understanding of invasive weeds' invasion process.
Chapter
Botanicals and derived herbal products have been long used in meeting primary healthcare needs around the world. They have gained increasing popularity in industrialized countries as complementary and alternative therapies to synthetic pharmaceuticals. Botanical preparations are however susceptible to contamination, adulteration, and substitution, that raise serious quality and public health safety concerns. The deliberate or accidental use of undeclared ingredients may occur throughout the entire value chain of the botanical product, from the cultivation or wild harvesting of the raw botanical material to the final packaged and marketed product. The current identification and quality inspection use tests to establish the identity, purity, and constituents of the herbal products. These tests most often utilize sensory and/or phytochemical inspection to detect species-specific characteristics or compounds, along with assays for toxic constituents such as heavy metals and toxins. However, marketed herbal products are very often highly processed and sold as complex mixtures. Classical analytical methods are often limited in their ability to identify the target plant species that should be present, and it is even more difficult to detect non-targeted species using these methods. This chapter presents a review of DNA barcoding and metabarcoding approaches to assess species diversity in herbal products and the use of these approaches in their quality control.
Article
Full-text available
[ Phil. Trans. R. Soc. B 360 , 1889–1895 (2005; Published 29 October 2005) ([doi:10.1098/rstb.2005.1720][2])][2] Author Nadia Haider's last name was spelt incorrectly as Nadia Haidar. The corrected author list is as follows: Mark W. Chase, Nicolas Salamin, Mike Wilkinson, James M. Dunwell, Rao
Article
Full-text available
Although much biological research depends upon species diagnoses, taxonomic expertise is collapsing. We are convinced that the sole prospect for a sustainable identification capability lies in the construction of systems that employ DNA sequences as taxon 'barcodes'. We establish that the mitochondrial gene cytochrome c oxidase I (COI) can serve as the core of a global bioidentification system for animals. First, we demonstrate that COI profiles, derived from the low-density sampling of higher taxonomic categories, ordinarily assign newly analysed taxa to the appropriate phylum or order. Second, we demonstrate that species-level assignments can be obtained by creating comprehensive COI profiles. A model COI profile, based upon the analysis of a single individual from each of 200 closely allied species of lepidopterans, was 100% successful in correctly identifying subsequent specimens. When fully developed, a COI identification system will provide a reliable, cost-effective and accessible solution to the current problem of species identification. Its assembly will also generate important new insights into the diversification of life and the rules of molecular evolution.
Article
Full-text available
An increasing number of studies using genetic data at the generic and species levels reveal complex patterns of relationships among populations and species. Incomplete lineage sorting and reticulation are, for instance, commonly observed in low-rank phylogenies. These two processes falsify the simplifying assumptions often used when reconstructing phylogenetic relationships or when assigning specimens to species using barcodes, i.e., the absence of ancestral polymorphism or the obligate dichotomous relationship among species. We recognize seven processes that act as veils to obscure species boundaries and relationships and lead to erroneous interpretations if not diagnosed and analysed properly. These processes include the transfer of genetic material from plastids or mitochondria to the nuclear genome (NuPt, NuMt), hybridization, the stochasticity of lineage sorting within and among species, genome organization, species demography, selection, genetic and geographic structure. Most of these processes have a direct impact on species effective population sizes, a central parameter to be considered. Failure to take the seven processes into account can affect the accuracy of barcode-based identification, in particular between closely related species. It can also bias the delimitation of species, result in inaccurate estimations of species relationships and lead to incorrect estimation of species ages. We suggest that these problems can be identified and, to some extent, mitigated by ensuring that the full spatial and genetic ranges of the species are sampled, and by the use of information from all genomes of the species. By analysing numerous loci, each as a separate entity in a coalescence framework, it is possible to take into account stochasticity as well as variation in genome effective population sizes. Such a critical approach would allow the use of genetic data to make realistic assessments of species delimitation, affinities and ages, and should promote a more biologically realistic view of species and speciation.
Article
Full-text available
Identifying how organismal attributes and environmental change affect lineage diversification is essential to our understanding of biodiversity. With the largest phylogeny yet compiled for grasses, we present an example of a key physiological innovation that promoted high diversification rates. C4 photosynthesis, a complex suite of traits that improves photosynthetic efficiency under conditions of drought, high temperatures, and low atmospheric CO2, has evolved repeatedly in one lineage of grasses and was consistently associated with elevated diversification rates. In most cases there was a significant lag time between the origin of the pathway and subsequent radiations, suggesting that the 'C4 effect' is complex and derives from the interplay of the C4 syndrome with other factors. We also identified comparable radiations occurring during the same time period in C3 Pooid grasses, a diverse, cold-adapted grassland lineage that has never evolved C4 photosynthesis. The mid to late Miocene was an especially important period of both C3 and C4 grass diversification, coincident with the global development of extensive, open biomes in both warm and cool climates. As is likely true for most "key innovations", the C4 effect is context dependent and only relevant within a particular organismal background and when particular ecological opportunities became available.
Conference Paper
There is uncertainty in the number of species of Physalis L. occurring in the Indian subcontinent, which may be due to (i) the introduction of weeds and cultivated species, and (ii) cryptic diversity due to hybridization amongst closely related species. These factors may allow these species to inhabit areas outside of their native range and habitats. Cultivation and breeding may also influence the distribution and diversity of Physalis species in India. Physalis lagascae (Physalis lagascae Roem. & Schult. = (P. minima auct non L.) is locally called ciyattini, erumaittakkali, koticceni, koticcenittakkali, kutavayacceti; this is a traditional medicinal plant used for treating chronic ailments such as stomach ulcers and several forms of cancer. Another closely related species, Physalis peruviana, is referred locally as cirutakkali, cirutakkali, itayakali, nattuttakkali, and it is known to inhabit the same areas as P. lagascae. The two species are often confused due to their morphological similarity. However, P. peruviana has different medicinal properties from P. lagascae. Our research has shown that traditional healers have often misidentified these species and their respective medicinal properties. This is unfortunate because P. peruviana is invasive and widely abundant, while P. lagascae is a local species with reduced populations due to over exploitation for medicinal use and has even been extirpated in some areas of India. Hence, the authentication of this species is crucial for its conservation and for maintaining sufficient populations for the herbal industry. Our research utilized rbcL and ITS sequence data to clearly differentiate samples of Physalis peruviana and Physalis lagasae. DNA barcoding using the tiered approach may provide a useful authentication tool for differentiating these species, which could be used to authenticate medicinal materials and support conservation initiatives that support the preservation of small populations.
Article
Willows (Salix: Salicaceae) form a major ecological component of Holarctic floras, and consequently are an obvious target for a DNA-based identification system. We surveyed two to seven plastid genome regions (~3.8 kb; ~3% of the genome) from 71 Salix species across all five subgenera, to assess their performance as DNA barcode markers. Although Salix has a relatively high level of interspecific hybridization, this may not sufficiently explain the near complete failure of barcoding that we observed: only one species had a unique barcode. We recovered 39 unique haplotypes, from more than 500 specimens, that could be partitioned into six major haplotype groups. A unique variant of group I (haplotype 1*) was shared by 53 species in three of five Salix subgenera. This unusual pattern of haplotype sharing across infrageneric taxa is suggestive of either a massive non-random coalescence failure (incomplete lineage sorting), or of repeated plastid capture events, possibly including a historical selective sweep of haplotype 1* across taxonomic sections. The former is unlikely as molecular dating indicates that haplotype 1* originated recently, and is nested in the oldest major haplotype group in the genus. Further, we detected significant non-neutrality in the frequency spectrum of mutations in group I, but not outside group I, and demonstrated a striking absence of geographical (isolation-by-distance) effects in the haplotype distributions of this group. The most likely explanation for the patterns we observed involves recent repeated plastid capture events, aided by widespread hybridization and long-range seed dispersal, but primarily propelled by one or more trans-species selective sweeps. This article is protected by copyright. All rights reserved.
Article
... by careful choice of markers and appropriate sampling; however, overall fine-scale species discrimination in plants relative to animals may be inherently more difficult ... Indeed, in plant barcoding studies that have included multiple samples per species (Fazekas et al. ...