Вы находитесь на странице: 1из 14

St/sl. Biol.

55(5)715-728,2006
Copyright Society of Systematic Biologists
ISSN: 1063-5157 print / 1076-836X online
DO1:10.1080/10635150600969864

DNA Barcoding and Taxonomy in Diptera: A Tale of High Intraspecific Variability


and Low Identification Success
RUDOLF MEIER, KWONG SHIYANG, GAURAV VAIDYA, AND PETER K. L. NG

Abstract.DNA barcoding and DNA taxonomy have recently been proposed as solutions to the crisis of taxonomy and
received significant attention from scientific journals, grant agencies, natural history museums, and mainstream media.
Here, we test two key claims of molecular taxonomy using 1333 mitochondrial COI sequences for 449 species of Diptera. We
investigate whether sequences can be used for species identification ("DNA barcoding") and find a relatively low success
rate (<70%) based on tree-based and newly proposed species identification criteria. Misidentifications are due to wide
overlap between intra- and interspecific genetic variability, which causes 6.5% of all query sequences to have allospecific
or a mixture of allo- and conspecific (3.6%) best-matching barcodes. Even when two COI sequences are identical, there is a
6% chance that they belong to different species. We also find that 21% of all species lack unique barcodes when consensus
sequences of all conspecific sequences are used. Lastly, we test whether DNA sequences yield an unambiguous species-level
taxonomy when sequence profiles are assembled based on pairwise distance thresholds. We find many sequence triplets for
which two of the three pairwise distances remain below the threshold, whereas the third exceeds it; i.e., it is impossible to
consistently delimit species based on pairwise distances. Furthermore, for species profiles based on a 3% threshold, only 47%
of all profiles are consistent with currently accepted species limits, 20% contain more than one species, and 33% only some
sequences from one species; i.e., adopting such a DNA taxonomy would require the redescription of a large proportion of
the known species, thus worsening the taxonomic impediment. We conclude with an outlook on the prospects of obtaining
complete barcode databases and the future use of DNA sequences in a modern integrative taxonomy. [Cytochrome oxidase
I; genetic distance; pairwise distance; species identification.]

Species description and identification are among the


most important tasks in biology, because biologists can
neither report empirical results nor access published
information on a study organism until it is correctly
named and/or identified. Descriptive taxonomy started
in earnest in the 18th century and after 250 years, 1.5
to 1.8 million species have been described with an estimated 5 to 100 million species awaiting discovery and/or
description (Wilson, 2003). Not surprisingly, taxonomic
character sources have changed over the past 250 years.
However, new techniques have been largely unable to
prevent a much lamented crisis of taxonomy, which is
now itself an endangered species (Godfray and Knapp,
2004; Wheeler, 2004). One of the most important problems for the future of alpha taxonomy is the slow rate
at which taxonomists can revise, describe, and identify
species. For example, it has been estimated that another
940 years will be required before all species are described
using traditional techniques (Seberg, 2004). Not surprisingly, biologists are thus looking for alternatives that
can speed up the process (Hogg and Hebert, 2004). Two
DNA sequence-based approaches are currently receiving much attention. According to their supporters, the
approaches have the potential to improve the prospects
of alpha-taxonomical research (Hebert et al., 2003a; Tautz
et al., 2003).
First, Hebert et al. (2003a) proposed an identification
system for specimens based on "DNA barcodes." The
"Barcoding Life" consortium (Hebert et al., 2004a) proposes to sequence approximately 650 bp of the mitochondrial COI gene for each species and argues that these
sequences can be used for species identification. In a
barcoded world, an unidentified specimen will be determined based on its COI sequence, which is matched

to an identified DNA barcode from a publicly available


database. Second, Tautz et al. (2002, 2003), envision a
"DNA taxonomy" in which DNA sequences will ultimately provide the main scaffold for a species-level taxonomy. Examples of DNA-based taxonomy can already
be found in numerous papers that use genetic distances
for revising species limits (e.g., Chu et al., 1999, 2003;
Shih et al., 2004; Sun et al., 2003; Tang et al., 2000, 2003;
Zhi et al., 1996). DNA barcoding and DNA taxonomy
have attracted much attention and discussion (e.g., Baker
et al., 2003; Barrett and Hebert, 2005; Blaxter and Floyd,
2003; Dunn, 2003; Ebach and Holdrege, 2005; Ferguson,
2002; Hebert and Barrett, 2005; Hebert and Gregory, 2005;
Hebert et al., 2003a, 2003b, 2004a, 2004b; Janzen, 2004;
Lee, 2004; Lipscomb et al., 2003; Marshall, 2005; Moritz
and Cicero, 2004; Pennisi, 2003; Prendini, 2005; Seberg,
2004; Seberg et al., 2003; Tautz et al., 2002,2003; Will and
Rubinoff, 2004; Will et al., 2005). However, the latter initially mostly focused on theoretical issues, whereas we
will here test empirically whether, regardless of all theoretical challenges, DNA taxonomy and DNA barcoding
can deliver reliable species-level classifications and/or
species identifications.
For this purpose, we are using a data set composed
of 1333 COI sequences for 449 species of Diptera, 127
of which are represented by multiple sequences. For
testing DNA barcoding, we pretend for each sequence
that we do not know its species identity. We then identify the query using different identification criteria (see
below) and count how many identifications yield the
name recorded in GenBank. For testing the feasibility
of a DNA taxonomy based on pairwise distances, we assemble DNA profiles and assess whether these profiles
correspond to currently accepted species.

715

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

Department of Biological Sciences, National University of Singapore, 14 Science Drive 4, Singapore 117543, Singapore;
E-mail: dbsmr@nus.edu. sg (R.M.)

716

SYSTEMATIC BIOLOGY

VOL. 55

Barrett and Hebert, 2005; Hebert et al., 2003a, 2003b,


Tautz et al.'s (2002, 2003) proposal of a DNA taxon- 2004b; Hogg and Hebert, 2004; Meyer and Paulay, 2005;
omy based on sequences has received less attention than Ward et al., 2005). However, most employed data sets
Hebert's work on DNA barcodes. This is conceivably due that included few or no species with multiple sequences
to the lack of detail on how a DNA taxonomy using se- and/or few closely related species (Moritz and Cicero,
quences as a "scaffold" can be constructed. Simplistic 2004; Prendini, 2005; Sperling, 2003; Will and Rubinoff,
approaches using pairwise distances were rejected by 2004). It therefore remains unclear whether intraspecific
the authors and synthetic approaches utilizing morpho- and interspecific variability are sufficiently distinct for
logical and molecular data espoused (Tautz et alv 2003). DNA barcoding to be a viable technique (Seberg, 2004;
Stoeckle, 2003). Identification was furthermore based on
However, it remained unclear how these data could be barcodes derived from individual organisms instead of
meaningfully combined and how taxonomists would re- using the information from all known sequences for a
solve conflict between data partitions. Meanwhile, many given species. We are here testing whether species retain
taxonomic publications using DNA sequences as the unique barcodes when we use the consensus sequence
main scaffold have been published. However, these ap- for all available conspecific sequences. Lastly, the accuproaches using "molecular operational taxonomic units" racy of the tree-based species identification techniques
(Blaxter, 2003; Blaxter and Floyd, 2003; Floyd et al., 2002) that were utilized by Hebert et al. (2003a) have been
generally use pairwise distances for clustering sequences questioned (DeSalle et al., 2005; Prendini, 2005; Will and
by similarity, although the choice of threshold value Rubinoff, 2004; Will et al., 2005). We here address all
for distinguishing intra- from interspecific distances is these issues by using a data set rich in closely related selargely arbitrary (DeSalle et al., 2005; Ferguson, 2002; Will quences (see above), testing the uniqueness of consensus
and Rubinoff, 2004).
species barcodes, and by establishing identification sucOne additional and largely neglected problem is that cess based on several different identification techniques.
the use of pairwise distances or other similarity measures can lead to logically inconsistent results. Three sequences can have two pairwise distances conforming to
Tree-Based Identification Techniques
and one exceeding a given threshold (Fig. 1). Under this
Techniques
for identifying sequences of unknown
circumstance, it remains unclear whether all three seprovenance
are
still in their infancy and generally fall
quences should be included in the same DNA species
into
two
categories.
Most studies use tree-based idengiven that one distance is violating the threshold. We are
tification
tools
based
on neighbor-joining trees (Barrett
here using our empirical data to test how frequently this
and
Hebert,
2005;
Floyd
et al., 2002; Hebert et al., 2003a;
situation is encountered in a real data set. We also assess, based on different thresholds, whether the limits Tautz et al., 2003). Queries are considered successfully
of "DNA species" based on distances correspond to the identified when they cluster with conspecific barcodes.
There are several problems with this approach (Will and
currently accepted species limits.
Rubinoff, 2004). Imagine a query clustering with a chimp
barcode.
Based on the query's position, one cannot deDNA Barcoding
cide whether it comes from Homo sapiens or another
Hebert et al. (2003a) proposed that unidentified speci- chimp; i.e., forming a cluster on a tree is logically insuffimens can be reliably identified to species based on DNA cient for identifying a sequence (Will and Rubinoff, 2004).
sequences (DNA barcodes). This proposal is imminently Yet, Barrett and Hebert (2005), Hebert (2003a), Hebert
testable and several tests have been carried out (e.g., et al. (2004b) have consistently used clustering as an indication for identification success. We would argue that
at best only those queries can be identified whose position unambiguously allow for the assignment of one
species name to the query sequence (Will and Rubinoff,
2004). These queries are found at least one node into a
clade consisting of only sequences from the same species
or are part of a polytomy formed by conspecific barcodes
4.79%
(see Table 1).
A second problem with tree-based identification as
currently practiced is the reliance on neighbor-joining
trees (Prendini, 2005; Will and Rubinoff, 2004). Data ambiguity is here difficult to detect, because most neighborjoining algorithms only generate a single tree even if
other trees have the same fit to the data ("tie trees"; see
2.87%
Takezaki, 1998). Tree choice then becomes dependent on
taxon entry order, which is hardly acceptable for a techFIGURE 1. Pairwise distances for three Anopheles sequences
nique
used in science (Backeljau et al., 1996). Simula(AF491693.1, AY258179.1 = A. maculipennis; AY258169.1 = A. messae).
All belong to the same 3% DNA profile, although one pairwise distance tion studies have furthermore revealed that "tie trees"
are particularly common in trees with short internal
exceeds the threshold.
DNA Taxonomy

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

2006

MEIER ET AL.TEST OF MOLECULAR TAXONOMY

717

TABLE 1. Overview of query identification criteria used in this study.

Tree-based identification sensu


Hebert

(a) Query in polytomy with


conspecifics
OR
(b) Query at least one node into a
clade exclusively consisting of
conspecifics

'Best match"

Sequence(s) with smallest


distance to query all
conspecific

'Best close match"

Sequence(s) with smallest


distance to query all
conspecific and within the
95th percentile of all
intraspecific distances

'All species barcodes"

As in "best match," but with all


conspecific sequences topping
the list of best matches

branches (Takezaki, 1998) and these can be expected to be


common on identification trees given that they usually
have multiple sequences from the same or closely related species. We are here addressing these concerns by
not only using neighbor-joining, but also parsimony and
Bayesian analyses for reconstructing identification trees.
Lastly, there are a host of conceptional problems with
using trees for identification (Will and Rubinoff, 2004).
For example, the "clustering" requirement assumes that
species should be monophyletic and not only ignores
the literature that questions the theoretical justification
for this requirement (Wheeler and Meier, 2000), but also
the substantial empirical evidence indicating that a large
proportion of currently recognized species are "paraphyletic" on gene trees (Crisp and Chandler, 1996; Funk
and Omland, 2003).
Alternative Identification Criteria

As an alternative to tree-based identification, we also


use identifications based on direct sequence comparison,
where a wide variety of techniques and metrics could be
used and a number of new criteria and algorithms have
recently been proposed (Blaxter et al., 2005; DeSalle et al.,
2005; Pozhitkov and Tautz, 2002; Steinke et al., 2005).
We here use three criteria (Table 1). Our first and least
stringent identification criterion is "best match." Any
query is assigned the species name of its best-matching
barcode regardless of how similar the query and
barcode sequences are. Obviously, under this criterion

Ambiguous/unidentified
Query without conspecific
sequence in data set
("singleton")
(a) Query without conspecific
sequence in data set
OR
(b) Query sister group to
conspecifics
OR
(c) Query in polytomy with at
least one conspecific and one
allospecific sequence
Sequence(s) with smallest
distance to query a mixture of
allo- and conspecific
sequences
(a) Sequence(s) with smallest
distance to query a mixture of
allo- and conspecific
sequences and within the 95th
percentile of all infraspecific
distances
(b) No match within 95 percentile
of all infraspecific distances
As in "best match", but with at
least one allospecific sequence
being more similar to query
than the least similar
conspecific sequence

Misidentification
Query with conspecific sequences
in multiple clusters
(a) Query at least one node into
an allospecific clade
OR
(b) Query in polytomy with only
allospecific sequences

All sequence(s) with smallest


distance to query allospecific
All sequence(s) with smallest
distance to query allospecific
and within 95th percentile of
all infraspecific distances

Query with conspecific sequence


in data set is more similar to
all sequences from another
species

misidentifications are common and even unavoidable


for all queries that belong to species without conspecific
barcodes in the database (Will and Rubinoff, 2004). For
example, the first Homo query will be most similar to a
great ape barcode and thus automatically identified as
such. Such mistakes are avoided by using our second
identification strategy, "best close match." Here, we first
identify the best barcode match of a query, but then only
assign the species name of that barcode to the query if
the barcode is sufficiently similar. In all other cases the
query remains unidentified.
This strategy requires a threshold similarity value that
defines how similar a barcode match needs to be before it can be identified. This value can be estimated for
a given data set by obtaining a frequency distribution
of all intraspecific pairwise distances and determining
the threshold distance below which 95% of all intraspecific distances are found. If, for example, 95% of conspecific sequences have pairwise distances below 1%, then
a query can only be identified according to the best close
match criterion if the query has a match in the barcode
data set that falls into the 0% to 1% interval. All queries
without such a match would remain unidentified. The
main practical drawback of this approach is that a given
data set may not be representative for the taxon under
investigation. For example, a very biased sample of conspecific and/or congeneric sequences could push the
threshold value up or down. It also remains unclear why
one should believe that there would be common threshold across species. However, in the absence of better

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

Tree-based identification, revised


criteria

Successful identification
Query clusters with all
conspecific sequences

718

VOL. 55

SYSTEMATIC BIOLOGY

submit the sequence only identified to genus (e.g., Pons


and Vogler, 2005). It is important to remember that data
sets containing a large number of sequences from putatively cryptic species submitted under the original and
potentially incorrect name could have a negative effect
on the performance of DNA barcodes.
Regardless of these drawbacks of GenBank data, we
believe that sequences from this database currently provide the best test databases for molecular taxonomy because only GenBank can provide data that are rich in
congeneric sequences and have a similar submission profiles as future barcode databases. At the very least, they
allow us to test to what degree DNA-based identification
and identification based on traditional tools will come to
congruent results.

Use of GenBank Data

MATERIALS AND METHODS

Similar to other tests of DNA barcoding and DNA


taxonomy (e.g., Barrett and Hebert, 2005; Hebert et al,
2003b), we are here utilizing sequences from GenBank,
although it is known to include misidentified sequences
(e.g., Harris, 2003; Hebert et al., 2003b; Ruedas et al., 2000;
Seberg, 2004; Vilgalys, 2003). Could it be that our study
is thus underestimating the potential of DNA barcoding? For several reasons, we believe this is not the case.
First, any future barcode database will be similar to GenBank in that many researchers will submit sequences; i.e.,
misidentified sequences in the barcode database should
be as common as they are in GenBank (Seberg, 2004).
Even when better vouchering policies are implemented,
it is unrealistic to believe that most vouchers will ever
be reidentified by taxonomic experts and therefore only
the most glaring mistakes may be found. Second, it has
been proposed that DNA barcodes should be generated based on museum specimens and existing tissues
in cryo-collections, although the former contain many
misidentified specimens (e.g., Meier and Dikow, 2004)
and the latter frequently have poor vouchering policies;
i.e., DNA barcode databases based on these sources will
again be similar to GenBank in containing a significant
number of misidentified sequences. Third, we carefully
inspected the structure of our database and found that
the sequences that were misidentified using DNA barcodes had been submitted by researchers that on average
provided nine conspecific and 36 congeneric sequences
(Appendix 1; available online at the Society of Systematic Biology Web site, http://systematicbiology.org); i.e.,
the misidentified sequences came from laboratories that
were carrying out phylogeographic studies. Such laboratories were probably as careful about identifying their
specimens as can be expected from any future barcoder.
However, authors of phylogeographic studies have yet to
settle on a policy of how to submit sequences from specimens that yield unusually high intraspecific distances.
Some researchers submit under the name of the original
specimen identification (e.g., Frost et al., 1998), some indicate in their submission that such unusually divergent
sequence comes from a closely related species (e.g., by
using "cf" or "aff": Stahls et al., 2004), whereas others

Sequences, Alignment, and Pairwise Distances

Similar to the approach of Barrett and Hebert (2005)


and Hebert et al. (2003b), we used 1443 sequences
from GenBank and aligned them using ClustalX. We
eliminated sequences that were (1) too short (<300
bp, 57 sequences), (2) not identified to species (49
sequences), (3) came from species-hybridization experiments (Drosophila subquinaria GL25990046), and
(4) could not be aligned and/or translated into
proteins or had >30% sequence divergence to all
other COI sequences for Diptera (Dyscritomyia robusta
GL19879668; Drosophila busckii GL27657151; Drosophila

ajfinis GL27657153). Sequences with GenBank names that


are synonyms according to the Biosystematic Database
of World Diptera (Thompson, 2005) were renamed
using the valid name (Anopheles arabiensis is junior
synonym of Anopheles gambiae; 3 sequences). The remaining 1333 COI sequences came from 449 species of
Diptera, of which 127 species were represented by 1011
sequences (see supplementary material; available online
at http://systematicbiology.org). Sequences not belonging to COI were removed and misplaced gaps were corrected to yield a 1539-bp gap-free alignment lacking stop
codons (see supplementary material; available online at
http://systematicbiology.org). Most analyses were carried out using a program developed for this purpose
("TaxonDNA" available at http://taxondna.sf.net/). We
carried out separate analyses for all sequences with a
minimum of 300, 400, 500, and 600 overlap. Due to
large interspecific distances in the very speciose genus
Drosophila, we treated the Drosophila subgenera as separate genera.
In order to test for overlap between intraspecific with
interspecific genetic variability, we plotted all uncorrected pairwise distances for conspecific sequences and
all distances for interspecific, congeneric sequences. In
order to test whether all species have unique DNA barcodes, we first tested whether identical sequences were
shared by individuals from different species. We then
constructed species barcodes as the consensus sequence
of all conspecific sequences and again tested for the
uniqueness of these species barcodes. The intraspecific

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

DNA-based identification approaches, this technique at


least provides a rigorously derived threshold value.
Our last criterion for identifying queries is a more rigorous application of the best close match strategy. Here
we utilize information from all conspecific barcodes in
the database instead of just focusing on the barcode that
is most similar to the query. Imagine that the closest
match for a query is a list of all known sequences for
a single species. Under this circumstance, the identifier
will be more confident about assigning this species name
to the query than in cases where multiple species names
are found on the list of best matches. Indeed, a conservative identifier would probably only assign a species
name if the query is followed by all known barcodes for
a particular species and insist that there are at least two
conspecific matches.

2006

719

MEIER ET AL.TEST OF MOLECULAR TAXONOMY

sequence variability was summarized using IUPAC


codes and the consensus sequence had to be based on
at least two sequences.
DNA Barcoding: Tree-Based Query Identification

DNA Barcoding: Identifying Species Based on Distances


(see Table 1)

1. "Best match." We used TaxonDNA to find for


each query its closest barcode match. If both se-

DNA Taxonomy: Profiles Based on Distance Thresholds

We tested the viability of threshold values for distinguishing intra- from interspecific variability for 2%, 3%,
4%, 5%, and 6% thresholds. For this purpose, TaxonDNA
finds for each query a set of barcodes for which each sequence in the set has at least one other sequence within
the threshold distance. For all clusters, we determined
whether the largest observed distance exceeded the
threshold distance and whether they correspond to currently accepted species (= contains all sequences for
one species). If not, we determined whether it contained
sequences for several species (error 1) and/or not all
sequences for the same species (error 2).
RESULTS

Intra- and Interspecific Variabilities

The following results apply to sequences with 300


bp overlap. Corresponding results for the subset of sequences with 400, 500, and 600 bp overlap are found
in Tables 2 and 3 and document that identification
success is largely unrelated to sequence overlap. We
find that intraspecific and interspecific variation overlap widely in Diptera COI sequences (0% to 15.5%)
with 99% of all congeneric distances falling into this
interval (Table 2, Fig. 2). When the largest 5% of the
intraspecific and the lowest 5% of the interspecific values are excluded, the overlap shrinks to 2.31% to 3.34%
which holds 5.24% of all congeneric distances (Table 2).
The pattern of intra- versus interspecific variability

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

Using PAUP* (Swofford, 2002; Kimura 2-parameter


model as recommended in Barrett and Hebert, 2005;
ties broken randomly), we computed neighbor-joining
trees and bootstrap trees for the largest sets of congeneric sequences with at least 300 bp overlap. The
same data sets were analyzed using parsimony as implemented TNT (Goloboff et al, 2003) using New Technology Search = 15; find min. length = 3; bootstrap
250 replicates), and Bayesian analyses as implemented
in MrBayes 3.1 (Huelsenbeck and Ronquist, 2003). All
Bayesian searches were initiated from random starting
trees. For all data sets with congeneric sequences, the
GTR+I+G model was favored by the Akaike information
criterion and hierarchical likelihood-ratio testing as implemented in MrModelTest version 2.2 (Nylander, 2004).
The data set was run for 3,000,000 generations and a tree
was sampled every 300 generations, resulting in 10,000
trees. Chain stationarity had been achieved for all data
sets after 1,200,000 generations (burn-in) and 4000 trees
were subsequently discarded. Three independently repeated analyses resulted in similar tree topologies and
comparable clade probabilities and substitution model
parameters.
For all trees, identification success was initially assessed as described in Hebert et al. (2003a) and Table 1;
i.e., sequences were considered successfully identified
as long as they formed species-specific clusters. Species
with sequences at multiple positions in the tree were
considered failures and those species with a single sequence were counted as ambiguous. Second, we used
the revised identification criteria described in the introduction. We only considered queries to be correctly identified if they were found in a species-specific polytomy
or at least one node into a clade exclusively consisting of
sequences from one species. Ambiguous were all queries
belonging to species with one or two sequences and those
that formed a sister group to a cluster of conspecific sequences. We counted those sequences as misidentihed
that were assigned a definite but incorrect name (e.g., a
query within an allospecific sequence cluster). A special
case is polytomies of sequences from two species. If the
query was from a different species than all remaining
sequences in the polytomy, we counted the query as a
misidentification because an identifier will assume that
the query is conspecific with the remaining sequences.
However, if the polytomy had at least two sequences
each from two different species, then the query in the
polytomy was considered ambiguous, because the identifier will be aware that a query in such a polytomy cannot
be unambiguously identified.

quences were from the same species, the identification was considered a success, whereas mismatched
names were counted as failures. Several equally good
best matches from different species were considered
ambiguous.
2. "Best close match." We used TaxonDNA to plot the
relative frequency of intraspecific distances in order
to determine the threshold value below 95% of all intraspecific distances are found. All queries without
barcode match below the threshold value remained
unidentified. For the remaining queries, their identity was compared to the species identity of their closest barcode. If the name was identical, the query was
considered an identification success. The identification was considered a failure when the names were
mismatched and considered ambiguous when several
equally good best matches were found that belonged
to a minimum of two species.
3. "All species barcodes." We assembled for each query a
list of all barcodes sorted by similarity to the query using the same threshold as for best close match. Queries
were considered a success when they were followed
by all conspecific barcodes as long as there were at
least two barcodes for the species. Queries were considered ambiguous when they were followed by only
one conspecific barcode or only some of the conspecific sequences. Queries followed by all conspecific sequences from the "wrong" species were considered
misidentified.

VOL. 55

SYSTEMATIC BIOLOGY

720

TABLE 2. Intra- and interspecific sequence variability and identification success based on "All species barcodes."

Sequence
overlap

Total overlap
(% observations)
0-15.5%
0-15.5%
0-15.5%
0-15.5%

987
975
664
406

(99)
(99)
(100)
(100)

is very inconsistent across genera, with some essentially lacking overlap whereas others have little separation. Species barcodes sensu stricto constructed as the
union-based consensus of all conspecific barcodes revealed that 34 species are involved in 22 cases of two
species sharing identical barcodes (Appendix 2; available online at http://systematicbiology.org). Twentyfive of these species belong to the 117 species in the
data set with multiple sequences that have a minimum overlap of 300 bp; i.e., when multiple individuals are sequenced 21% of the species lack unique
barcodes.

90% Overlap
(% observations)

Queries with
allospecific
identical match

"All species
barcodes"
success

2.31-3.34 (5.2)
2.29-3.34 (5.6)
1.81-3.25 (9.3)
0.84-5.14 (33.1)

6% (35)
6% (35)
7.9% (28)
16.9% (28)

40.6%
41.7%
34.3%
33.1%

all sequences for the species would fail because it is not


"monophyletic."
Success of Similarity-Based DNA Identification Techniques
(see supplementary material, available online at:
http://systematicbiology.org)

Success under "best match" is 67.7%, 79 queries are


ambiguous (5.3%), and 361 are misidentified (Table 3;
27.1%). The data set contains 588 sequences whose best
match is an identical sequence. Of these, 35 have an allospecific identical match (6.0%; see Table 2 for corresponding results with larger sequence overlap).
In order to be able to use best close match, we first deSuccess of Tree-Based DNA Identification Techniques
termined that 95% of all intraspecific distances fall into
According to Hebert et al.'s (2003a) criteria and the interval from 0% to 3.34%; i.e., the latter value was
depending on tree reconstruction techniques (NJ, used to decide whether a query had a close enough barparsimony, Bayesian), only 40% to 47% of all queries rep- code match for identification. Success under best close
resenting 22% to 23% of all species are successfully iden- match is 66.3%, 5.9% of all sequences are misidentified,
tified (Table 4; supplementary material, available online and the remaining queries (26.4%) remain unidentified
at http://systematicbiology.org). Many sequences (37% because they have no match below 3.34% (Table 3).
For the "all species barcodes" criterion the success rate
to 43%) and species (15% to 16%) fail to form speciesspecific clusters and are misidentified. Due to the lack is 40.6%, in 23.0% there is no match below the threshold,
of conspecific sequences in the data set, a large propor- for 35.2% the result is ambiguous (<2 conspecific bartion of species (62%) and sequences (16%) are ambiguous codes or mixture of con- and allospecific barcodes top
and remain unidentified. The highest identification suc- the list), and 1.2% of all queries are misidentified.
cess and lowest misidentification rate are observed for
DNA Taxonomy: Profiles Based on Distance Thresholds
neighbor-joining trees followed by Bayesian and parsimony trees. Compared to the first set of tree-based identiBecause the distances between three sequences do not
fication criteria, the second set based on the revised rules have to be equilateral, a fixed threshold value cannot be
described in the Introduction have a higher success rate maintained. The following increases in cutoff values are
(60% to 63%), whereas the proportion of ambiguously observed within the respective sets: 2% -> 4.8%; 3% -
placed sequences remains high (39%). Misidentifications 4.8%; 4% - 6.1%; 5% -* 8.5 (Table 5). All cutoff values
are rare using both criteria (1.2%; Table 4). Note that the lead to species-level classifications that are radically difsuccess rate for sequences is higher under the latter cri- ferent from accepted species limits. For example, for the
terion, because it does not require monophyly; i.e., con- popular 3% threshold, only 47% of the clusters agree with
specific sequences found in two different clades on the traditional species and one cluster contains eight species.
tree can be identified as long as the name assignment Corresponding results for other thresholds are found in
is unambiguous. Under Hebert et al.'s (2003a) criterion, Table 5.

TABLE 3. Identification success based on "best match" and "best close match."
"Best match"

"Best close match"

Sequence overlap

Success

Ambiguous

Misidentification

300
400
500
600

67.7%
68.1%
65.0%
55.9%

5.3%
5.4%
5.9%
8.5%

27.1%
26.5%
29.0%
35.7%

Success
66.9%
67.5%
64.1%
55.1%

Ambiguous

Misidentification

No match

3.6%
3.7%
4.2%
6.3%

6.5%
6.0%
5.7%
11.2%

23.0%
22.8%
26.0%
27.4%

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

300
400
500
600

Number of
conspecific
sequences

2006

721

MEIER ET AL.TEST OF MOLECULAR TAXONOMY

interspecific
intraspecific

2%

4%

6%

8%

10% 12% 14%

16% 18%

Pairwise Distances
FIGURE 2. Overlap between intra- and interspecific genetic variability for congeneric sequences.

to 15.5%) and that many of the pairwise distances for


congeneric sequences fall into the area of overlap (99%;
Fig. 2). The overlap remains considerable even when the
extreme 5% of all intra- and interspecific distances are
removed (2.31% to 3.34%: 5.2% of distances). These results are not unexpected given that numerous phylogeographic studies have repeatedly contradicted Hebert's
contention that "the gene's [COI] sequence doesn't appear to vary among individuals of the same species"
(see Pennisi, 2003). Given the variability of COI, it is obviously insufficient to generate a single DNA barcode
per species. Instead, several individuals from different
parts of a species' range have to be included in serious
DNA barcoding projects. Our results also highlight that
the "barcode" analogy between species barcodes and the
kinds of barcodes used in industry is problematic given
that variable barcodes would never be tolerated in the
product barcodes used for commercial purposes (Lee,
2004; Moritz and Cicero, 2004).
Also frequently discussed in the theoretical literature
is the problem of identical barcodes shared by several
species (Ferguson, 2002; Floyd et al., 2002; Quicke, 2004;
Tautz et al., 2003). The existence of snared barcodes is beyond doubt, but for a technique like DNA barcoding, it is
more important to know how common they are. We find
that the incidence at the level of individual barcodes is
moderate. Of the queries with identical barcode matches,
only 6% share their barcode with an allospecific species.
Intra- and Interspecific Variabilities
However, this result nevertheless implies that it is imposMuch of the literature on DNA barcoding and DNA sible at the 5% level to support the intuitive conclusion
taxonomy consists of theoretical arguments in favor of that identical barcodes also come from the same species.
and against the use of DNA sequences in taxonomy. We
We find it more worrisome that there are so many
believe that these arguments are necessary and useful, species with identical consensus barcodes (21 % of species
but we also believe that more empirical information is with multiple sequences). Consensus sequences have nuneeded. One issue that has been discussed from a theoret- merous drawbacks in the context of phylogenetic reconical point of view but only partially addressed with em- struction and study of molecular evolution (e.g., Page
pirical data (but see Avise, 2000; Avise and Walker, 1999; and Holmes, 1998), but they are of interest in evaluating
Johns and Avise, 1998) is the overlap between intraspe- the performance of DNA barcodes for diagnostic purcific and interspecies genetic variabilities (Stoeckle, 2003; poses. Ideally, a character used for species identification
Ward et al., 2005; Will and Rubinoff, 2004). We are here and description should be diagnostic in the sense that
revealing that the overlap is extensive in Diptera (0% it unambiguously distinguishes all individuals of one
DISCUSSION

Many biologists have argued that the future of descriptive taxonomy will depend on successfully embracing new techniques. Many ideas have been proposed
(Godfray, 2002) and much progress has been achieved
by, for example, digitizing taxonomic information, using new microscopic techniques, and developing interactive identification tools (De Ley et al., 2005; Dunn, 2003;
Klaus et al., 2003; Marshall, 2003; Prendini, 2005; Sperling, 2003; Thacker, 2003; Wheeler, 2004; Will et al., 2005).
It is only natural that the discussion has now turned to the
contribution of DNA sequences. Uncontroversial is their
great promise for associating morphologically disparate
life history stages, associating males and females, solving species limits for polymorphic and cryptic species,
and identifying artefacts made of materials derived from
endangered species (Birstein et al., 1998; Hebert et al.,
2003a; Marshall, 2003; Paquin and Hedin, 2004; Palumbi
and Cipriano, 1998; Tautz et al., 2003). However, controversial is the extent to which DNA sequences and biologists only versed in molecular techniques should and
can replace the morphological tools and retiring traditional taxonomists. Most systematists would argue that
a more straightforward solution to employing molecular
techniques is hiring new taxonomists (Lee, 2004; Minelli,
2003; Will and Rubinoff, 2004).

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

0%

Parsimony
Bayes

NJ

Aedes
Anastrepha
Anopheles
Asphondylia
Bactrocera
Calliphora
Chiastocheta
Chironomus
Chrysomya
Culicoides
Liriomyza
Lucilia
Paragus
Phytomyza
Sapromyza
Scathophaga
Subgenus
Drosophila
Sophophora
Total

Genus

31/3
16/5

4/2
10/4

34/3
11/4

93/33
111/47
915/243

430 (47.0%)/55(22.6%)
368 (40.2%)/53 (21.8%)
410 (44.8%)/55 (22.6%)

22/6
2/1
153/5
11/2
27/2
15/2
22/4
2/1
10/3
18/4
18/3
16/2
0/0
9/4
14/3
24/5

Bayes

22/6
2/1
155/6
11/2
27/2
15/2
22/4
2/1
10/3
18/4
18/3
5/1
0/0
9/4
14/3
24/5

Par

22/6
2/1
155/6
42/3
27/2
15/2
22/4
2/1
10/3
18/4
18/3
5/1
0/0
9/4
14/3
24/5

NJ

32/14
45/15
188/11
42/3
27/2
15/2
50/9
43/34
16/9
86/6
19/4
50/7
12/7
19/14
32/11
35/15

No. of
seq./spp.

61/3
64/6

3/1
35/6
29/1
31/1
0/0
0/0
28/5
10/2
0/0
67/1
0/0
44/5
8/3
0/0
14/4
2/1

Par

34/2
61/6

3/1
35/6
31/2
31/1
0/0
0/0
28/5
10/2
0/0
67/1
0/0
34/4
8/3
0/0
14/4
2/1

Bayes

337 (36.8%)/37 (15.2%)


396 (43.3%)/39 (16.0%)
361 (39.5%)/38 (15.6%)

31/2
63/6

3/1
35/6
29/1
0/0
0/0
0/0
28/5
10/2
0/0
67/1
0/0
44/5
8/3
0/0
14/4
2/1

NJ

Failed seq./spp.

28
37

7
8
4
0
0
0
0
31
6
1
1
1
4
10
4
9

Singletons
(= ambiguous)

151 (16.5)/151 (62.1%)


151 (16.5)/151 (62.1%)
151 (16.5)/151 (62.1%)

Hebertet al.'i5 identification criteria

55
28

13
10
179
40
27
13
34
3
5
81
14
35
0
2
13
23

NJ

35
31

13
13
:177
38
24
15
27
2
6
82
16
39
0
2
14
24

Bayes

575 (62.8%)
550 (60.1%)
558 (61.0%)

32
28

14
11
178
39
26
14
28
3
5
81
15
39
0
2
11
24

Par

Successful sequences

38
80

19
33
8
2
0
2
16
39
11
4
5
14
12
17
17
12

NJ

329 (36.0%)
354 (38.7%)
346 (37.8%)

60
79

18
32
9
3
1
1
22
39
11
4
4
10
12
17
21
11

Par

57
76

19
31
10
4
3
0
23
41
10
3
3
10
12
17
17
10

Bayes

Ambiguous identification

Revised identification criteria

Success rate for tree-based species identification. Singleton = sequence without conspecific representation in data set.

Successful seq./spp.

TABLE 4.

0
1
1
0
0
0
0
0
0
1
0
1
0
0
1
1
1
4

0
2
1
0
0
0
0
1
0
1
0
1
0
0
0
0
1
4

0
2
1
0
0
0
0
1
0
0
1
0
0
2
0
0
3

11 (1.2%)
11 (1.2%)
11 (1.2%)

Bayes

Par
NJ

Misident. sequences

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

I-H

2006

723

MEIER ET AL.TEST OF MOLECULAR TAXONOMY

TABLE 5. DNA taxonomy: clusters based on fixed pairwise-distance thresholds for all sequences from species with intraspecific sequences
(>300 bp overlap: 987 sequences, 117 species).
Threshold
pairwise distance

125
106
95
82
75

Profiles with
threshold violations
15
14
13
14
9

(12%)
(13%)
(14%)
(17%)
(12%)

Maximum
pairwise distance

No. of profiles compatible with trad,


species/profiles with only one species

4.8%
4.8%
6.1%
8.5%
12.3%

55 (44%)/105
50 (47%)/85
47 (49%)/73
37 (45%)/59
37 (49%)/54

species from the individuals of all other species (e.g.,


DeSalle et al., 2005; Panchen 1992: "monothetic" versus
"polythetic" species; Winston, 1999). In traditional taxonomy as well as species described based on DNA sequence
evidence, such characters are provided in the differential
diagnosis for the new species (Bond, 2004; Bond and
Sierwald, 2003; Winston, 1999). Our results suggest that
COI is not a very good diagnostic tool in Diptera because 21% of all species for which multiple sequences
are available have identical consensus sequences. Our
finding does not imply that COI cannot be used for
species identification, but for many species determinations will have to rely on indirect techniques such as
tree-based assignment of specimens to sequence clusters or probability-based statements about the presence
of nucleotides at particular sites. The lack of speciesspecific barcodes also bodes poorly for attempts to
develop short oligonucleotide probes for species identification (Gibbs et al., 2005; Summerbell et al., 2005). Several authors had envisioned microarray or DNA chip
approaches to species identification, but no probe will
be able to distinguish between species that share COI
consensus barcodes. We believe that the high frequency
of shared consensus barcodes indicates that DNA barcoding will ultimately only be useful for identifying
species to species groups. For many applied purposes of
species identification (e.g., biosecurity: Armstrong and
Ball, 2005; identification of medically important fungi:
Summerbell et al., 2005), this will be satisfactory, but
the real challenge in taxonomy is to find new and better

Max. no.
of species per profile

techniques for delimiting and identifying closely related


species (Sperling, 2003).
Identification Success Rates

Overall, we find that for our data set the identification success rates for DNA barcoding are unacceptably low and never exceed 70% (Fig. 3). Furthermore,
this low success rate does not improve even if sequences with greater sequence overlap are used (Tables
2 and 3). However, larger data sets will be needed to
rigorously test this prediction. In Diptera identification,
success actually declines with increasing sequence overlap, which we believe to be due to the relatively small
number of sequences with more than 500 base pairs of
overlap. Whether these success rates are high enough to
justify the considerable expense for barcoding all species
and obtaining sequences for unidentified specimens remains a judgement call for the user. However, we doubt
that DNA barcoding has a bright future unless the identification success rates can be increased considerably. Incidentally, our best identification rate of 68% is in line
with the rates reported in Will and Rubinoff (2004) and
decades of research on intraspecific sequence variability in a wide range of taxa (summarized in Funk and
Omland, 2003). Funk and Omland (2003) found that 23%
of the 2319 assayed species did not form monophyletic
groups and that in two-thirds of these cases, the "polyphyly" was supported by bootstrap values above 70%.
For arthropods, the rate was even higher (26.5% of 702

100%

ffl no match
E3 misidentified
ambiguous
success

300 bp

400 bp
500 bp
Basepair Overlap

600 bp

FIGURE 3. Sequence overlap and identification success based on best match (1st bar), best close match (2nd bar), and all species barcodes
(3rd bar).

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

2%
3%
4%
5%
6%

No. of
DNA profiles

724

SYSTEMATIC BIOLOGY

2005). Given the problems with tie-trees and the problematic assumption of species monophyly, we do not believe that trees are a promising tool (see also DeSalle
et al., 2005; Pozhitkov and Tautz, 2002; Steinke et al.,
2005). This is particularly so because each tree has many
nested clades and it is thus impossible to decide based
on tree topology alone which node on the tree delimits a
population, species, or supraspecific taxon (see our Homo
sapiens-great ape example in the introduction). Instead,
some kind of similarity criterion has to be used in conjunction with the tree before one can decide whether a
particular query is conspecific or allospecific to its closest match. Given that a similarity metric will also be
needed for tree-based identification, we would argue
that one may as well abandon the tree-based approaches
and instead directly use the similarity metric as has been
implemented in several recent algorithms (e.g., Blaxter
et al., 2005; Steinke et al., 2005). We furthermore find
that for our data set, similarity-based techniques outperform tree-based identification. For example, the success
rate of tree-based identification is below 50% (Table 4),
whereas best close match yields a much higher success
rate (67%) and has a relatively low incidence of misidentification (6%; Table 3). Therefore, if we had to use DNA
sequences for species identification, we would prefer this
technique, although a large proportion of the queries remain unidentified because they lack close matches in the
barcode database. We nevertheless consider this result
preferable over using an identification technique like
best match that yields marginally higher success rates
(68%) but also a large number of misidentifications (27%).
All species barcodes is the most conservative identification technique, but given its low success rate of 39% it
will probably only be justifiable for forensic purposes.
Is a Complete Barcode Database Obtainable?

Best match would perform much better if it was applied to a data set from which single-sequence species
have been removed. However, in a real-life situation
it is impossible to know whether a new query is coming from a "new" species or from a species that is already represented in the data set (Scotland et al., 2003;
Seberg, 2004; Will and Rubinoff, 2004). Of course, the
other solution would be using a database that has barcodes for all species on the planet. Creating such a
database may alleviate some of the problems with DNA
barcoding and is thus the goal of the Barcoding Life
Consortium. However, we agree with many systematists that this goal is completely unrealistic (Marshall,
2003). The existing strategies for creating a complete
barcode database mostly rely on tissues from existing
cryo-collections and museum specimens. However, the
former will only yield barcodes for relatively few, mostly
common, and/or widespread species. The proposed use
of tissues from identified specimens in natural history
museums is equally inadequate and problematic.
First, museums can at best provide identified specimens for described speciesfor poorly known groups

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

spp. surveyed). Furthermore, Moritz and Cicero (2004)


found that 74% of 29 surveyed sister species pairs of
birds would not be recognized as species if Hebert et al.'s
(2003b) rule for delimiting species was applied; i.e., our
results appear more in line with Will and Rubinoff's
(2004), Funk and Omland's (2003), and Moritz and Cicero's (2004) assessment than the empirical studies that
have been published by proponents of DNA barcoding.
The reader of the DNA barcoding literature will be
surprised by our low success rates, which contrast
sharply with the perfect or near perfect success rates reported elsewhere (Barrett and Hebert, 2005; Hebert and
Gregory, 2005; Hebert et al., 2003a, 2004b; Hogg and
Hebert, 2004). We believe that the main reason for this
discrepancy is study design. Previously, two different
kinds of tests of barcoding had been performed. The first
was based on extensive compilations of intra- and interspecific COI distances across a wide range of taxa (Barrett
and Hebert, 2005; Floyd et al., 2002; Hebert et al., 2003a;
Hogg and Hebert, 2004) and they revealed that the average distances between congeneric species is relatively
large. However, average distances between species are
of little relevance when it comes to identifying closely
related species. This is familiar to all taxonomists who
know that it may be easy to separate a species from
the "average" species in its genus but that it can still be
very challenging to distinguish sister species. Contrary
to suggestions in the literature (e.g., Barrett and Hebert,
2005; Hebert et al., 2003a; Hogg and Hebert, 2004), average similarities to other species in a genus are irrelevant
for predicting identification success for closely related
species.
The second kind of empirical test of DNA barcoding was conducted based on specimens collected at one
or a few sites. The problem with these studies is the
lack of consideration for geographic variation (Moritz
and Cicero, 2004; Prendini, 2005; Sperling, 2003; Will
and Rubinoff, 2004; Ward et al., 2005), although it is
well known that species with a wide geographic distribution often contain a considerable amount of genetic variability. By only sampling individuals from a
single locality, this variability is not considered and the
distinctness of species barcodes is easily overestimated.
Sparse taxon sampling at the species level also makes
it likely that mostly distantly related species are included in the test, thus increasing the chance of finding distinct species-specific sequence differences (e.g.,
Will and Rubinoff, 2004). We believe that the dense
taxon sampling in our data set accounts for some of the
identification problems encountered in our study, because some genera and/or species groups had clearly
been sampled very extensively by experts carrying out
phylogeographic studies within closely related species
and/or populations (Appendix 1; available online at
http: / / sy sterna ticbiology. org).
Techniques for identifying sequences to species using DNA barcoding are still in their infancy, but new
methods and algorithms are rapidly appearing (e.g.,
Blaxter et al., 2005; Nilsson et al., 2004; Steinke et al.,

VOL. 55

2006

725

MEIER ET AL.TEST OF MOLECULAR TAXONOMY

DNA Taxonomy

Supporters of DNA taxonomy have suggested that a


functional taxonomy can be based wholly on DNA sequences. The most popular approach involves the use of
fixed sequence-divergence values for delimiting species
(e.g., 3%; Barrett and Hebert, 2005; Hebert et al., 2003a).
One problem with this approach is that the choice of
threshold value is arbitrary and renders species borders a matter of opinion (Ferguson, 2002; Prendini, 2005;
Will and Rubinoff, 2004). However, an equally serious
and more fundamental theoretical challenge is that it is
logically impossible to maintain such threshold values
(Fig. 1). We find numerous cases in Diptera where two
out of the three pairwise distances for three sequences remain below a given threshold while the third sequence
exceeds the threshold (Fig. 1). For example, for 3% sequence clusters, 13% of the 106 DNA profiles have pairwise distances in excess of the threshold. The largest
observed distance in a 3% cluster is 4.8%, which implies
that even if a query sequence differs by 4.5% from another sequence the query sequence may still belong to
the same 3% DNA species (Table 5); i.e., species based on
thresholds only lead to a convenient taxonomic system
as long as few sequences are known. As taxon sampling
improves, the maze of pairwise distances becomes impenetrable and convenient fixed values are unenforceable unless one is willing to accept that sequence input
order influences the composition of the sequence clusters (see Blaxter et al., 2005). Note that these results are
largely independent of which pairwise distance is used
and whether the sequence overlap is small (300 bp) or
large (600 bp; Fig. 4).
Furthermore, for the Diptera data set, the majority
(53%) of the 3% profiles contradict traditionally recognized species with some having sequences in six different profiles (Table 5). Instead of speeding up taxonomic
work, a sequence-based taxonomy would thus have to
start with the redescription of most species. This will
increase the taxonomic impediment and further slow

49%

47%

1 45%
U
3 43%
300 bp overlap

39%
37%

400 bp overlap
500 bp overlap
600 bp overlap

35%
2%

3%

4%

5%

6%

7%

Pairwise Distance Threshold


FIGURE 4. Relationship between sequence overlap and proportion of clusters that are compatible with currently accepted species
boundaries.

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

this is a minute fraction of the actual diversity. It is in


this context revealing that current work on barcoding
concentrates on taxonomically well-known groups such
as birds, fish, and Lepidoptera (Hebert and Gregory,
2005; Hebert et al., 2003a, 2004b). However, if barcoding
is restricted to these kinds of taxa, then the very groups
that are most in need of new taxonomic techniques will
be left out. For example, only 50,000 of the estimated 1
million species of nematodes have been described (see
Seberg, 2004) and recent revisions of invertebrates routinely double species counts (Ponder and Lunney, 1999).
This indicates mat prerevision collections contain at best
identified material for 50% of all species. Species richness
estimation furthermore clearly indicates that many additional species remain uncollected and are thus not even
represented in the collections (Marshall, 2003; Seberg,
2004; Meier and Dikow, 2004). All these species will have
to be formally described before a reference sequence can
be deposited in a barcode database. Barcoding can thus
never be faster than traditional taxonomy. We would
even argue that barcoding is currently even exasperating the taxonomic crisis by submitting large numbers of
poorly identified sequences for putatively cryptic species
to GenBank. Of the 5426 sequences with the keyword
"barcode" that had been submitted by February 2006,
2459 (45%) were only identified to genus. At the same
time, Hebert and Gregory (2005) explicitly state DNA sequences should not be used for describing new species so
that these sequences will probably remain unidentified
for the foreseeable future.
Second, much of the identified material in museums
is either unsuitable and/or too valuable for molecular
work (Quicke, 2004; Will and Rubinoff, 2004). A good
example are the approximately 40% of known beetle
species that have only been collected at a single locality (see Seberg, 2004). Prior to pinning, most specimens
in all likelihood went through the conventional entomological pinning procedures, which involves softening in
a moist chamber for several days. Such treatment will degrade much of the DNA and probably explains the generally low gene-amplification success for DNA extracted
from pinned insect specimens (e.g., 31% for "archival
moths": Hajibabaei et al., 2005). The material for other
groups is even less accessible. For example, nematodes, mites, and fish were in the past mostly preserved
in formaldehyde and /or slide-mounted and especially
many nematode specimens have been lost (De Ley et al.,
2005).
Third, even if museum specimens are well preserved
and available for DNA extraction, they are often not
a good source for generating barcodes because a large
proportion are misidentified (see examples in Meier and
Dikow, 2004; Ward et al., 2005). In conclusion, no strategy
is in sight that will yield an even approximately complete barcode database for those groups that urgently require new identification techniques. However, our and
one other recent empirical study (e.g., Meyer and Paulay,
2005) document that barcoding will yield low identification success if the barcode database is incomplete.

726

VOL. 55

SYSTEMATIC BIOLOGY

valuable for many purposes and should be extensively


used where needed (Armstrong and Ball, 2005; Birstein
et al., 1998; Hebert et al., 2003a; Palumbi and Cipriano,
1998; Paquin and Hedin, 2004; Tautz et al., 2003; Will
et al., 2005). There is also evidence that DNA barcoding can be successful in some taxa and/or at a regional
scale. For example, we find that Aedes and Anopheles
mosquitoes identify relatively well based on COI and
expect the same for species samples lacking large numbers of closely related species (Armstrong and Ball, 2005;
Moritz and Cicero, 2004; Sperling, 2003; Stoeckle, 2003;
Summerbell et al., 2005). DNA sequences also allow for
the identification of genetic diversity and unusual patterns of genetic variability that require further study
(e.g., Hebert and Gregory, 2005; Paquin and Hedin, 2004;
Quicke, 2004). Over the past 250 years the character basis for describing and identifying species has continually
broadened, and DNA sequences are a very valuable new
tool. Sequences certainly improve our ability to recognize
and describe species (DeSalle et al., 2005; Paquin and
Hedin, 2004; Sperling, 2003; Will et al., 2005). For some
groups of organisms and life history stages, the old tools
have been struggling and here DNA sequences will turn
out to be the character source of choice (e.g., Armstrong
and Ball, 2005; Paquin and Hedin, 2004). For other clades,
DNA sequences will only play a minor role. Barcoding
all of life may be a bold proposal that has recently been
skilfully promoted (Sperling, 2003), but it is really quite
unnecessary for taxa like birds (Dunn, 2003) and can be
misleading in other taxa like Diptera. Systematists have
not failed to describe all species because of a lack of effort,
but rather, because of the overwhelming species diversity in combination with the complexities involved in
recognizing closely related species. Simplistic proposals
based on DNA sequences only create pseudosolutions to
real biological problems and what is really needed is an
integrative approach to taxonomy embracing all available evidence (DeSalle et al., 2005; Paquin and Hedin,
2004; Will etal., 2005).
ACKNOWLEDGMENTS
We gratefully acknowledge comments on the manuscripts by numerous colleagues, including two anonymous reviewers, Roderic Page,
and Marshal Hedin. Financial support came from the Academic Research Fund grants R-154-000-256-112 and R154-000-270-112 from the
Ministry of Education, Singapore.

CONCLUSIONS

Our empirical test of DNA taxonomy reveals that


an identification system exclusively relying on COI for
Diptera successfully determines less than 70% of all sequences. Misidentification rates are low, but for a significant number of species the sequences are not divergent
enough or too divergent to allow identification. Given
the prevalence of violations of pairwise-distance thresholds, an alpha-taxonomy based on DNA sequences is
similarly problematic. However, it is important to stress
that these results do not question the value of DNA sequences in taxonomy. DNA sequences are already in-

REFERENCES
Armstrong, K. F., and S. L. Ball. 2005. DNA barcodes for biosecurity:
Invasive species identification. Philos. Trans. R. Soc. Lond. B Biol.
Sci. 360:1813-1823.
Avise, J. C. 2000. Phylogeography: The history and formation of species.
Harvard University Press, Cambridge, Massachusetts.
Avise, J. C, and D. Walker. 1999. Species realities and numbers in sexual vertebrates: Perspectives from an asexually transmitted genome.
Proc. Natl. Art. Sci. USA 96:992-995.
Backeljau, T., L. De Bruyn, H. De Wolf, K. Jordaens, S. Van Dongen, and
B. Winnepennincks. 1996. Multiple UPGMA and neighbor-joining
trees and the performance of some computer packages. Mol. Biol.
Evol. 13:309-313.

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

taxonomic progress. Because new sequences can change


the pairwise-distance profiles within DNA species by
fusing two clusters, we also dispute that a DNA taxonomy would be more stable than traditional taxonomy as
has been claimed. We tested multiple threshold values,
but all lead to major revisions of existing species-level
classifications and all would cause complete taxonomic
chaos in Diptera (Table 5). Given these problems it is
surprising that distance-based approaches to taxonomy
remain so popular (Blaxter, 2003; Chu et al., 1999, 2003;
Shih et al., 2004; Sun et al., 2003; Tang et al., 2000, 2003;
Zhi et al., 1996).
Proponents of a DNA-based taxonomy may argue
that species paraphyly, large intraspecific variability,
and small interspecific variability are reasons to revise
the borders of traditionally recognized species (Barrett
and Hebert, 2005; Blaxter, 2003; Hebert et al., 2003a,
2004b; Hogg and Hebert, 2004; Tautz et al., 2003; Zhi
et al., 1996). However, there are compelling reasons why
species boundaries cannot be mechanically modified in
response to genetic distances (Ferguson, 2002; Lee, 2004).
The genes used in molecular taxonomy vary predominantly in selectively largely neutral nucleotide positions
and therefore distances among closely related sequences
reflect to a large extent time of divergence; i.e., if we were
to break up species with large intraspecific variability, we
would effectively deny that species can be ancient. Similarly, given standard rates of COI evolution (Pestano
et al., 2003), lumping species with low intraspecific variability is equivalent to refusing species status to species
younger than 1.5 million years regardless of reproductive
isolation and morphological distinctiveness. Mechanically modifying species borders in response to small
distances also ignores all research indicating that reproductive isolation can evolve rapidly via sexual selection
(e.g., Ferguson, 2002; Lee, 2004; Moritz and Cicero, 2004)
and that mitochondrial genes can be misleading due to
phenomena such as lineage sorting, male-biased gene
flow, introgression following hybridization, and numts
(e.g., Moritz and Cicero, 2004). One should also remember that in contrast to the characters used in traditional
taxonomy (e.g., genitalia, mating calls, coloration, etc.;
e.g., Lee, 2004), the genes used in molecular taxonomy are
not functionally correlated with speciation (Ferguson,
2002; Lee, 2004).

2006

MEIER ET AL.TEST OF MOLECULAR TAXONOMY

727

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

Baker, C. S., M. L. Dalebout, S. Lavery, and H. A. Ross. 2003. www.DNA- Hebert, P. D. N., A. Cywinska, S. L. Ball, and J. R. deWaard. 2003a.
surveillance: Applied molecular taxonomy for species conservation
Biological identifications through DNA barcodes. Proc. Roy. Soc. Ser.
and discovery. TREE 18:271-272.
B 270:313-321.
Barrett, R. D. H., and P. D. N. Hebert. 2005. Identifying spiders through Hebert, P. D. N., and T. R. Gregory. 2005. The promise of DNA barcoding
DNA barcodes. Can. J. Zool. 83:481^191.
for taxonomy. Syst. Biol. 54:852-859.
Birstein, V. J., P. Doukakis, B. Sorkin, and R. Desalle. 1998. Population Hebert, P. D. N., S. Ratnasingham, and J. R. deWaard. 2003b. Barcodaggregation analysis of three caviar-producing species of sturgeons
ing animal life: Cytochrome c oxidase subunit 1 divergences among
closely related species. Proc. Roy. Soc. Ser. B 270:S96-S99.
and implications for the species identification of black caviar. Cons.
Biol. 12:766-775.
Hebert, P. D. N., S. Ratnasingham, and R. Dooh. 2004a. Barcodes of
Life, http://www.barcodinglife.com/
Blaxter, M. 2003. Counting angels with DNA. Nature 421:122-124.
Blaxter, M., and R. Floyd. 2003. Molecular taxonomies for biodiversity Hebert, P. D. N., M. Stoeckle, T. S. Zemlak, and C. M. Francis. 2004b.
surveys: Already a reality. TREE 18:268-269.
Identification of birds through DNA barcodes. PLoS Biol. 2:1657Blaxter, M., J. Mann, T. Chapman, F. Thomas, C. Whitton, R. Floyd, and
1663.
E. Abebe. 2005. Defining operational taxonomic units using DNA Hogg, I. D., and P. D. N. Hebert. 2004. Biological identification of springbarcode data. Philos. Trans. R. Soc. Lond. B Biol. Sci. 360:1935-1943.
tails (Hexapoda: Collembola) from the Canadian Arctic, using mitoBond, J. E. 2004. Systematics of the Californian euctenizine spider genus
chondrial DNA barcodes. Can. J. Zool. 82:749-754.
Apomastus (Araneae: Mygalomorphae: Cyrtaucheniidae): The rela- Janzen, D. 2004. Now is the time. Philos. Trans. R. Soc. Lond. B Biol.
tionship between molecular and morphological taxonomy. Invert.
Sci. 359:731-732.
Syst. 18:361-376.
Johns, G. C, and J. C. Avise. 1998. A comparative summary of genetic
Bond, J. E., and P. Sierwald. 2003. Molecular taxonomy of the Anadenobo- distances in the vertebrates from the mitochondrial cytochrome b
lus excisus (Diplopoda: Spirobolida: Rhinocricidae) species-group on
gene. Mol. Biol. Evol. 15:1481-1490.
the Caribbean island of Jamaica. Invert. Syst. 17:515-528.
Klaus, A. V., V. L. Kulasekera, and V. Schawaroch. 2003. ThreeChu, K. H., H. Y. Ho, C. P. Li, and T. Y. Chan. 2003. Molecular phyloge- dimensional visualization of insect morphology using confocal laser
netics of the mitten crab species in Eriocheir, sensu lato (Brachyura:
scanning microscopy. J. Microsc. 212:107-121.
Grapsidae). J. Crust. Zool. 23:738-746.
Lee, M. S. Y. 2004. The molecularisation of taxonomy. Inv. Syst. 18:1-6.
Chu, K. H., J. Tong, and T. Y. Chan. 1999. Mitochondrial cytochrome Lipscomb, D., N. Platnick, and Q. Wheeler. 2003. The intellectual
content of taxonomy: A comment on DNA taxonomy. TREE 18:65-66.
oxidase I sequence divergence in some Chinese species of Charybdis (Crustacea: Decapoda: Portunidae). Biochem. Syst. Ecol. 27:461- Marshall, E. 2005. Taxonomy: Will DNA Bar Codes Breathe Life Into
468.
Classification? Science 307:1037.
Crisp, M. D., and G. T. Chandler. 1996. Paraphyletic species. Telopea Marshall, S. 2003. The real costs of insect identification [opinion page].
6:813-844.
Newsl. Biol. Surv. Can. (Terrestrial Arthropods) 22:15-18.
De Ley, P., I. T. De Ley, K. Morris, E. Abebe, M. Mundo-Ocampo, Meier, R., and T. Dikow. 2004. Significance of specimen databases from
M. Yoder, J. Heras, D. Waumann, A. Rocha-Olivares, A. H. J. Burr,
taxonomic revisions for estimating and mapping the global species
J. G. Baldwin, and W. K. Thomas. 2005. An integrated approach to
diversity of invertebrates and repatriating reliable and complete
fast and informative morphological vouchering of nematodes for apspecimen data. Cons. Biol. 18:478-488.
plications in molecular barcoding. Philos. Trans. R. Soc. Lond. B Biol. Meyer, C. P., and G. Paulay. 2005. DNA barcoding: Error rates based
Sci. 360:1945-1958.
on comprehensive sampling. PLoS Biol. 3:2229-2238.
DeSalle, R., M. G. Egan, and M. Siddall. 2005. The unholy trinity: Tax- Moritz, C, and C. Cicero. 2004. DNA barcoding: Promise and pitfalls.
onomy, species delimitation and DNA barcoding. Philos. Trans. R.
PLoS Biol. 2:1529-1531.
Soc. Lond. B Biol. Sci. 360:1905-1916.
Nilsson, R. H., K. H. Larsson, and B. M. Ursing. 2004. GalaxieCGI
Dunn, C. P. 2003. Keeping taxonomy based in morphology. TREE
scripts for sequence identification through automated phylogenetic
18:270-271.
analysis. Bioinformatics 20:1447-1452.
Ebach, M. C, and C. Holdrege. 2005. DNA barcoding is no substitute Nylander, J. A. A. 2004. MrModelTest, version 2. Evolutionary Biology
for taxonomy. Nature 434:697.
Centre, Uppsala University.
Ferguson, J. W. H. 2002. On the use of genetic divergence for identifying Page, R. D. M., and E. C. Holmes. 1998. Molecular evolution: A phylospecies. Biol. J. Linn. Soc. 75:509-516.
genetic approach. Blackwell Science, Oxford, UK.
Floyd, R., E. Abebe, A. Papert, and M. Blaxter. 2002. Molecular barcodes Palumbi, S. R., and F. Cipriano. 1998. Species identification using gefor soil nematode identification. Mol. Ecol. 11:839-850.
netic tools: The value of nuclear and mitochondrial gene sequences
Frost, D. R., H. M. Crafts, L. A. Fitzgerald, and T. A. Titus. 1998. Gein whale conservation. J. Hered. 89:459-464.
ographic variation, species recognition, and molecular evolution of Panchen, A. L. 1992. Classification, evolution, and the nature of biology.
cytochrome oxidase I in the Tropidurus spinulosus complex (Iguania: Cambridge University Press, Cambridge, UK.
Tropiduridae). Copeia 1998:839-851.
Paquin, P., and M. Hedin, M. 2004. The power and perils of "molecFunk, D. J., and K. E. Omland. 2003. Species-level paraphyly and polyular taxonomy": A case study of eyeless and endangered Cicurina
phyly: Frequency, causes, and consequences, with insights from an(Araneae: Dictynidae) from Texas caves. Mol. Ecol. 13: 3239-3255.
Pons, J., and A. P. Vogler. 2005. Complex pattern of coalescence and fast
imal mitochondrial DNA. Ann. Rev. Ecol. Syst. 34:397-423.
Gibbs, M. J., J. S. Armstrong, and A. J. Gibbs. 2005. Individual sequences evolution of a mitochondrial rRNA pseudogene in a recent radiation
in large sets of gene sequences may be distinguished efficiently by
of tiger beetles. Mol. Biol. Evol. 22:991-1000.
combinations of shared sub-sequences. Bmc Bioinformatics 6.
Pennisi, E. 2003. Modernizing the tree of life. Science 300:1692-1697.
Godfray, H. C. J. 2002. Challenges for taxonomy. Nature 417:17-19.
Pestano, J., R. P. Brown, N. M. Suarez, and M. Baez. 2003. DiversificaGodfray, H. C. J., and S. Rnapp. 2004. Taxonomy for the twentytion of sympatric Sapromyza (Diptera: Lauxaniidae) from Madeira:
first centuryIntroduction. Philos. Trans. R. Soc. Lond. B Biol. Sci.
Six morphological species but only four mtDNA lineages. Mol. Phy359:559-569.
logenet. Evol. 27:422-428.
Goloboff, P., J. S. Farris, and K. Nixon. 2003. T.N.T.: Tree analysis using Ponder, W., and D. Lunney. 1999. The other 99%The conservation
new technology. Program and documentation, available from the
and biodiversity of invertebrates. Trans. Roy. Zool. Soc. NSW.
authors, and at www.zmuc.dk/public/phylogeny.
Pozhitkov, A. E., and D. Tautz. 2002. An algorithm and program for
Hajibabaei, M., J. R. DeWaard, N. V. Ivanova, S. Ratnasingham, R. T.
finding sequence specific oligonucleotide probes for species identification. BMC Bioinformatics 3.
Dooh, S. L. Kirk, P. M. Mackie, and P. D. N. Hebert. 2005. Critical
factors for assembling a high volume of DNA barcodes. Philos. Trans. Prendini, L. 2005. Comment on "Identifying spiders through DNA barR. Soc. Lond. B Biol. Sci. 360:1959-1967.
codes." Can. J. Zool. 83:498-504.
Quicke, D. L. J. 2004. The world of DNA barcoding and morphology
Harris, D. J. 2003. Can you bank on GenBank? TREE 18:317-319.
Collision or synergism and what of the future? Systematist 23: 8-12.
Hebert, P. D. N., and R. D. H. Barrett. 2005. Reply to the comment by
L. Prendini on "Identifying spiders through DNA barcodes." Can. J. Ronquist, F., and J. P. Huelsenbeck. 2003. MrBayes 3: Bayesian phyloZool. 83:481-491.
genetic inference under mixed models. Bioinformatics 19:1572-1574.

728

SYSTEMATIC BIOLOGY

Tautz, D., P. Arctander, A. Minelli, R. H. Thomas, and A. P. Vogler. 2002.


DNA points the way ahead in taxonomy. Nature 418:479.
Tautz, D., P. Arctander, A. Minelli, R. H. Thomas, and A. P. Vogler. 2003.
A plea for DNA taxonomy. TREE 18:70-74.
Thacker, P. D. 2003. Morphology: The shape of things to come. BioScience 53:544-549.
Thompson, F. C. 2005. Biosystematic Database of World Diptera.
http://www.sel.barc.usda.gov/Diptera/biosys.htm
Tong, J. G., T. Y. Chan, and K. H. Chu. 2000. A preliminary phylogenetic analysis of Metapenaeopsis (Decapoda: Penaeidae) based on mitochondrial DNA sequences of selected species from the Indo West
Pacific. J. Crust. Biol. 20:541-549.
Vilgalys, R. 2003. Taxonomic misidentification in public DNA
databases. New Phytol. 160:4-5.
Ward, R. D., T. S. Zemlak, B. H. Innes, P. R. Last, and P. D. N. Hebert.
2005. DNA barcoding Australia's fish species. Philos. Trans. R. Soc.
Lond. B Biol. Sci. 360:1847-1857.
Wheeler, Q. D. 2004. Taxonomic triage and the poverty of phylogeny.
Philosophical Philos. Trans. R. Soc. Lond. B Biol. Sci. 359:571583.
Wheeler, Q. D., and R. Meier. 2000. Species concepts and phylogenetic
theory. A debate. Columbia University Press, New York.
Will, K. W., B. D. Mishler, and Q. D. Wheeler. 2005. The perils of DNA
barcoding and the need for integrative taxonomy. Syst. Biol. 54:844851.
Will, K. W, and D. Rubinoff. 2004. Myth of the molecule: DNA barcodes for species cannot replace morphology for identification and
classification. Cladistics 20:47-55.
Wilson, E. O. 2003. The encyclopedia of life. TREE 18:77-80.
Winston, J. E. 1999. Describing species: Practical taxonomic procedure
for biologists. Columbia University Press, New York.
Zhi, L., W. B. Karesh, D. N. Janczewski, H. Frazier-Taylor, D. Sajuthi, F.
Gombek, M. Andau, J. S. Martenson, and S. J. O'Brien. 1996. Genomic
differentiation among natural populations of orang-utan (Pongo pygmaeus). Curr. Biol. 6:1326-1336.

First submitted 6 September 2005; reviews returned 9 December 2005;


final acceptance 7 April 2006
Associate Editor: Marshal Hedin

Downloaded from http://sysbio.oxfordjournals.org at BIBLIOTECA CENTRAL UNIV ESTADUAL CAMPINAS on March 11, 2010

Ruedas, L. A., J. Salazar-Bravo, J. W. Dragoo, and T. L. Yates. 2000.


The importance of being earnest: What, if anything, constitutes a
specimen examined? Mol. Phylogenet. Evol. 17:129-132.
Scotland, R., C. Hughes, B. Donovan, and A. Wortley. 2003. The Big
Machine and the much-maligned taxonomist. Syst. and Biodiv. 1:139
143.
Seberg, 0.2004. The future of systematics: Assembling the Tree of Life.
The Systematist 23: 2-8.
Seberg, O., C. J. Humphries, S. Knapp, D. W. Stevenson, G. Petersen,
N. Scharff, and N. M. Andersen. 2003. Shortcuts in systematics? A
commentary on DNA-based taxonomy. TREE 18:63-65.
Shih, H.-T., P. K. L. Ng, and H.-W. Chang. 2004. The systematics of the
genus Ceothelphusa (Crustacea, Decapoda, Brachyura, Potamidae)
from southern Taiwan: A molecular appraisal. Zool. Stud. 43:561570.
Sperling, F. 2003. DNA barcoding. Deus et machina [opinion page].
Newsl. Biol. Surv. Can. (Terrestrial Arthropods) 22:50-53.
Stahls, G., J.-H. Shake, A. Vujic, D. Doczkal, and J. Muona. 2004. Phylogenetic relationships of the genus Cheilosia and the tribe Rhingiini
(Diptera, Syrphidae) based on morphological and molecular characters. Cladistics 20:105-122.
Steinke, D., M. Vences, W. Salzburger, and A. Meyer. 2005. Taxi: A
software tool for DNA barcoding using distance methods. Philos.
Trans. R. Soc. Lond. B Biol. Sci. 360:1975-1980.
Stoeckle, M. 2003. Taxonomy, DNA, and the Bar Code of Life. BioScience 53:796-797.
Summerbell, R. C, C. A. Levesque, K. A. Seifert, M. Bovers, J. W. Fell,
M. R. Diaz, T. Boekhout, G. S. de Hoog, J. Stalpers, and P. W. Crous.
2005. Microcoding: the second step in DNA barcoding. Philos. Trans.
R. Soc. Lond. B Biol. Sci. 360:1897-1903.
Sun, H. Y, K. Zhou, and X. J. Yang. 2003. Phylogenetic relationships
of the mitten crabs inferred from mitochondrial 16S rDNA partial
sequences (Crustacean, Decapoda). Acta Zool. Sin. 49:592-599.
Swofford, D. L. 2002. *PAUP*: Phylogenetic analysis using parsimony
(and other methods), version 4.0bl0. Sinauer Associates, Sunderland, Massachusetts.
Takezaki, N. 1998. Tie trees generated by distance methods of phylogenetic reconstruction. Mol. Biol. Evol. 15:727-737.
Tang, B., K. Zhou, D. Song, G. Yang, and A. Dai. 2003. Molecular
systematics of the Asian mitten crabs, genus Eriocheir (Crustacea:
Brachyura). Mol. Phylogenet. Evol. 29:309-316.

VOL. 55

Вам также может понравиться