Você está na página 1de 11

Available online at www.sciencedirect.

com

Forensic Science International 177 (2008) 66–76


www.elsevier.com/locate/forsciint

A global study of forensically significant calliphorids:


Implications for identification
M.L. Harvey a,*, S. Gaudieri a,b, M.H. Villet c, I.R. Dadour a
a
Centre for Forensic Science, M420, University of Western Australia, Nedlands 6907, Australia
b
School of Anatomy and Human Biology, University of Western Australia, Nedlands 6907, Australia
c
Department of Zoology and Entomology, Rhodes University, Grahamstown 6140, South Africa
Received 17 February 2006; received in revised form 12 September 2007; accepted 31 October 2007
Available online 4 March 2008

Abstract
A proliferation of molecular studies of the forensically significant Calliphoridae in the last decade has seen molecule-based identification of
immature and damaged specimens become a routine complement to traditional morphological identification as a preliminary to the accurate
estimation of post-mortem intervals (PMI), which depends on the use of species-specific developmental data. Published molecular studies have
tended to focus on generating data for geographically localised communities of species of importance, which has limited the consideration of
intraspecific variation in species of global distribution. This study used phylogenetic analysis to assess the species status of 27 forensically
important calliphorid species based on 1167 base pairs of the COI gene of 119 specimens from 22 countries, and confirmed the utility of the COI
gene in identifying most species. The species Lucilia cuprina, Chrysomya megacephala, Ch. saffranea, Ch. albifrontalis and Calliphora stygia
were unable to be monophyletically resolved based on these data. Identification of phylogenetically young species will require a faster-evolving
molecular marker, but most species could be unambiguously characterised by sampling relatively few conspecific individuals if they were from
distant localities. Intraspecific geographical variation was observed within Ch. rufifacies and L. cuprina, and is discussed with reference to
unrecognised species.
# 2008 Published by Elsevier Ireland Ltd.

Keywords: Calliphoridae; Forensic entomology; Blowflies; COI; Cox1; Intraspecific variation

1. Introduction majority of molecular studies, however, have used the


cytochrome oxidase I (COI or cox1) encoding region of
The advent of DNA-based identification techniques for use mitochondrial DNA (mtDNA) [1,2,4–6,9].
in forensic entomology in 1994 [1] saw the beginning of a The COI gene holds enormous utility for species identifica-
proliferation of molecular studies into the forensically tion. Lying within the mitochondrial genome, it has the
important Calliphoridae. The use of DNA to characterise advantages of easy isolation, higher copy number than its
morphologically indistinguishable immature calliphorids was nuclear counterparts, and conserved sequence and structure
recognised as a valuable molecular tool with enormous across taxa. COI has been well studied in the Insecta [10], with
practical utility. Numerous studies have since addressed the its utility for distinction between closely related species of
DNA-based identification of calliphorids [2–6]. A variety of Diptera demonstrated by the large number of COI studies of
regions of DNA have been suggested for study including the species complexes in the Culicidae (e.g. [11]).
nuclear internal transcribed spacers (ITS) [7], mitochondrial Calliphorid molecular taxonomic studies have focused
rRNA genes [8] and the mitochondrial control region [8]. The largely on sequencing of the COI gene and have illustrated
the ability to successfully distinguish between a wide variety of
forensically important species based largely on monophyly
[1,2,4–6,9]. The main limitation to the use of COI sequence
* Corresponding author at: School of Biological Sciences, University of
data has been the inability to distinguish between some closely
Portsmouth, King Henry Building, King Henry I Street, Portsmouth PO12DY,
UK. Tel.: +44 23 9284 5012. related species of the genus Calliphora, generally due to
E-mail address: Michelle.Harvey@port.ac.uk (M.L. Harvey). incidences of para- or polyphyly. Wallman et al. [4,12] found

0379-0738/$ – see front matter # 2008 Published by Elsevier Ireland Ltd.


doi:10.1016/j.forsciint.2007.10.009
M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76 67

Table 1
Individuals used in this study, listed with locality of origin and GenBank accession number, and publication data where identified from another publicationa
Species Locality Accession no. Source
Chrysomya saffranea Broome, Australia EU418533 New sequence
Broome, Australia EU418534 New sequence
Brisbane, Australia AB112841 [6]
Chrysomya megacephala Sydney, Australia EU418535 New sequence
Perth, Australia AB112846 [6]
Perth, Australia AB112847 [6]
Pretoria, South Africa AB112848 [6]
Kitwe, Zambia AB112861 [6]
Kitwe, Zambia AB112856 [6]
KwaZulu-Natal, South Africa AB112830 [6]
Hawaii, United States EU418536 New sequence
Papua New Guinea AF295551 [32]
Kuala Lumpur, Malaysia EU418537 New sequence
Malaysia AY909052 NCBI submission
Malaysia AY909053 NCBI Submission
Chrysomya pinguis Hsintien, Taipei County, Taiwan AY092759 [33]
Chrysomya bezziana Bogor, Indonesia AF295548 [32]
Chrysomya inclinata KwaZulu-Natal, South Africa AB112857 [6]
Chrysomya chloropyga Graaf-Reinet, South Africa EU418540 New sequence
Graaf-Reinet, South Africa EU418541 New sequence
Pretoria, South Africa EU418538 New sequence
KwaZulu-Natal, South Africa EU418539 New sequence
Chrysomya putoria Kitwe, Zambia AB112831 [6]
Kitwe, Zambia AB112860 [6]
Snake Island, Botswana AB112835 [6]
Snake Island, Botswana AB112855 [6]
Sao Joao da Boa Vista, Brazil EU418542 New sequence
near Chilbre, Panama AF295554 [32]
Chrysomya marginalis Pretoria, South Africa AB112838 [6]
Pretoria, South Africa AB112832 [6]
Karoo, South Africa AB112866 [6]
Karoo, South Africa AB112862 [6]
Karoo, South Africa EU418543 New sequence
KwaZulu-Natal, South Africa AB112837 [6]
KwaZulu-Natal, South Africa AB112834 [6]
Chrysomya varipes Gladstone, Australia EU418544 New sequence
Sydney, Australia EU418545 New sequence
Adelaide, Australia AF295556 [32]
Perth, Australia AB112868 [6]
Perth, Australia AB112869 [6]
Perth, Australia AB112867 [6]
Chrysomya norrisi Wau, Papua New Guinea AF295552 [32]
Chrysomya rufifacies Perth, Australia EU418546 New sequence
Perth, Australia AB112828 [6]
Perth, Australia AB112845 [6]
Campbell Town, Tasmania EU418547 New sequence
Florida, USA AF083658 [32]
Knoxville, USA EU418548 New sequence
Oahu, Hawaii, USA EU418549 New sequence
Chingmei, Taipei City, Taiwan AY092760 [33]
Malaysia AY909055 NCBI submission
Malaysia AY909054 NCBI submission
Chrysomya albiceps Alexandria, Egypt AF083657 [32]
Pretoria, South Africa AB112840 [6]
Pretoria, South Africa AB112839 [6]
KwaZulu-Natal, South Africa AB112836 [6]
KwaZulu-Natal, South Africa AB112842 [6]
Deka, Zimbabwe AB112849 [6]
Deka, Zimbabwe AB112858 [6]
Manzini, Swaziland AB112865 [6]
Manzini, Swaziland AB112854 [6]
Manzini, Swaziland AB112851 [6]
Cochliomyia hominivorax Alfenas, Brazil EU418550 New sequence
Cochliomyia macellaria Salvador, Brazil EU418551 New sequence
68 M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76
Table 1 (Continued )
Species Locality Accession no. Source
Gainesville, Florida AF295555 [32]
Calliphora dubia Geraldton, Western Australia EU418552 New sequence
Perth, Australia EU418553 New sequence
20km north New Norcia, Western Australia EU418554 New sequence
Ravensthorpe, Australia EU418555 New sequence
Toodyay, Australia EU418556 New sequence
Calliphora augur Sydney, Australia EU418557 New sequence
Sydney, Australia EU418558 New sequence
Calliphora hilli Gladstone, Tasmania EU418559 New sequence
Calliphora varifrons Boddington, Australia EU418560 New sequence
Calliphora ochracea Sydney, Australia EU418561 New sequence
Sydney, Australia EU418562 New sequence
Calliphora stygia Wallaceville, New Zealand EU418563 New sequence
Kaitoke, New Zealand EU418564 New sequence
Kempton, Tasmania EU418565 New sequence
Calliphora albifrontalis 20km north New Norcia, Australia EU418566 New sequence
Perth, Australia EU418567 New sequence
Perth, Australia EU418568 New sequence
Calliphora vomitoria Montferrier-Sur-Lez, France EU418569 New sequence
Calliphora vicina Montferrier-Sur-Lez, France EU418570 New sequence
Kempton, Tasmania EU418571 New sequence
London, UK EU418572 New sequence
London, UK EU418573 New sequence
Bristol University Colony, UK AJ417702 [32]
Lucilia illustris Montferrier-Sur-Lez, France EU418574 New sequence
Langford, UK AJ551445 [34]
Lucilia ampullacea Montferrier-Sur-Lez, France EU418575 New sequence
Lucilia cuprina Gladstone, Tasmania EU418576 New sequence
Perth, Australia AB112863 [6]
Perth, Australia AB112852 [6]
Perth, Australia AB112853 [6]
Townsville, Australia AJ417710 [34]
Dorie, New Zealand AJ417706 [34]
Chiang Mai University Lab Colony, Thailand EU418577 New sequence
Tororo, Uganda AJ417711 [34]
Dakar, Senegal AJ417708 [34]
Chingmei, Taipei City, Taiwan AY097335 [33]
Honolulu, Hawaii AJ417704 [34]
Waianae, Hawaii AJ417705 [34]
Lucilia sericata Montferrier-Sur-Lez, France EU418577 New sequence
Montferrier-Sur-Lez, France EU418578 New sequence
Perth, Australia AB112833 [6]
Graaf-Reinet, South Africa AB112850 [6]
Graaf-Reinet, South Africa AB112843 [6]
Pretoria, South Africa AB112864 [6]
Pretoria, South Africa AB112859 [6]
Harare, Zimbabwe AB112844 [6]
Harare, Zimbabwe AJ417717 [34]
Nerja, Spain AJ417716 [34]
Hilerod, Denmark AJ417712 [34]
Langford, Somerset, UK AJ417714 [34]
Dorie, New Zealand AJ417713 [34]
Los Angeles, USA AJ417715 [34]
Hydrotaea rostrata Perth, Australia AB112829 [6]
a
‘‘New sequence’’ indicates sequences have not been published elsewhere and have been submitted to the public databases, with release pending publication of this
study.

difficulty in separating the C. augur/C. dubia, C. stygia/C. Ch. megacephala and C. vicina. However, forensic entomol-
albifrontalis and C. hilli/C. varifrons species pairs, creating ogy is a locality-specific science and molecular studies are
difficulty in geographical regions where these sister species generally directed to the specific fauna found in a region [e.g.
overlap. Our first aim was to contribute to solving such 4–6]. Specimens from new localities may not exactly match
problems by enlarging the available data set. published DNA sequences, raising questions regarding
A number of species of forensic utility are relatively acceptable levels of variation for distinction. Our second
cosmopolitan, such as L. cuprina, L. sericata, Ch. rufifacies, aim was therefore to use conspecific specimens from
M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76 69

geographically distant localities to estimate the range of 2.3. Amplification


variation found in various recognised species, and thus
Approximately 1270 bp of the COI gene was amplified using the primers
evaluate the conservation of the COI gene over geographic
C1-J-1718 (50 –30 GGAGGATTTGGAAATTGATTAGTTCC) and TL2-N-3014
distances and implications for necessary sample size in (50 –30 TCCAATGCACTAATCTGCCATATTA) [14]. For the amplification of
taxonomic studies of these species. some species, TL2-N-3014 proved problematic and therefore a degenerate
Specimens that are similar to, but do not lie within the known primer TL2-N-3014MOD (50 –30 TCCATTGCACTAATCTGCCATATTA) was
range of genetic variation of, a recognised species may designed based on the sequence of Chrysomya chloropyga (accession number
represent previously-unsampled geographical variation of that AF352790), and used to amplify a number of individuals for which amplifica-
tion was not achieved with the original reverse primer.
species, another species recognised by systematics but not yet The PCR reaction mix composed of: 1X PCR buffer (Biotools; Fisher
sampled for DNA, or taxonomically unrecognised cryptic Biotec), 200 mM dNTPs (Biotools; Fisher Biotec), 1.5 mM MgCl2, 25 pM each
species. The occurrence of cryptic species, which appear primer, 1 unit of Taq polymerase (Biotools; Fisher Biotec), 10–150 ng of
morphologically the same as named species but differ in their template DNA, and water added to a total volume of 50 mL. Reactions were
behaviour, development or other biology, may contribute to performed on Perkin Elmer GeneAmp PCR System 2400 and Applied Bio-
systems GeneAmp PCR System 2700 thermocyclers. Cycling conditions were:
error in PMI estimates. For example, Ch. rufifacies is a fly with 90 s 94 8C denaturation, followed by 36 cycles of: 94 8C for 22 s, 48 8C for 30 s
a widespread distribution displaying variable behaviour in and 72 8C for 80 s. A final extension period of 1 min at 72 8C was used,
different localities [6], and Wallman et al. [12] have recently followed by holding at 4 8C. Products were visualised using 1.5% agarose gels
indicated the possibility of cryptic species within Ch. rufifacies with ethidium bromide staining and UV transillumination.
PCR products were purified using the QiaQuick PCR Purification Kit
in Australia.
(Qiagen), according to manufacturer’s instructions.
In molecular phylogenies, discrimination is commonly
based on the separation of monophyletic clades, or 2.4. COI sequencing
alternately, DNA barcoding studies have led to the suggested
application of heuristic thresholds or expected percent DNA Sequencing reactions were performed using the ABI PRISM Big Dye
sequence divergence in discriminating species and identify- Terminator 3.0 or 3.1 Sequencing Kit (Perkin Elmer), according to the
ing novel taxa [13]. The sampling of data over a target region manufacturer’s protocol. Cycling conditions for the sequencing reactions were
as per the manufacturer’s recommendations, but the annealing temperature was
of DNA provides data for the definition of prescribed levels
lowered to 48 8C. Individuals were sequenced using the external primers C1-J-
of inter and intraspecific variability, and therefore identifying 1718 and TL2-N-3014 or (TL2-N-3014MOD, where appropriate), and the
novel taxa. However, numerous authors have cautioned the internal primers C1-J-2183 (50 –30 CAACATTTATTTTGATTTTTTGG) and
use of such an approach, particularly within undersampled C1-N-2329 (50 –30 ACTGTAAATATATGATGAGCTCA).
taxa [23,24], where levels of intra and interspecific variation
may overlap and prevent accuracy. Our third aim was to 2.5. Sequence analysis
estimate the amount of genetic variation found within and
Sequences were visualised using Chromas v1.43 (http://trishul.sci.gu.e-
between recognised species, with a view to commenting on du.au/conor/chromas.html), and alignments and editing conducted using
the use of the threshold approach in the designation of novel DAPSA [15]. Sequences were submitted to DDBJ (accession numbers in
species and identification of species affinity of forensically Table 1). Additional COI sequences of some relevant calliphorids were obtained
important calliphorids. from the publicly available DNA database (Genbank) at www.ncbi.nlm.nih.gov
(Table 1).
This study gathered data on a variety of forensically
Phylogenetic analyses were performed using MEGA 3.0 [16], PAUP*4.0
significant calliphorids across their geographical distributions [17] and MrBayes v.3.0b4 [18]. MEGA was used to calculate pairwise distances
and assessed the potential for the COI gene to provide and was employed in the distance analysis using the neighbour-joining method
distinction between the species. Sequencing of 1167 base pairs with the Tamura-Nei model of substitution and 500 bootstrap replications. Base
of the mitochondrial COI gene was conducted and phylogenetic frequencies, transition/transversion ratio and the gamma shape parameter were
analysis used to represent the relationships between the taxa. estimated from the data using PAUP, and analyses were performed using
MrBayes v.3.0b4. These Bayesian inference analyses were conducted using
This study considered 119 flies from 28 species and 22 one cold and three hot chains, and the INVGAMMA model. Analyses were run
countries, including 47 new sequences. for 1,500,000 generations, sampling every 100 generations. The likelihood
scores from every 100 generations was plotted to evaluate when stationarity had
been reached. From the plots, it appeared that the burn-in phase was complete
2. Materials and methods by 50,000 generations. However, the first 1000 trees were excluded as burn-in,
this exclusion being considered to be conservative. Posterior probabilities (PP)
2.1. Samples were calculated from the remaining trees by means of a majority rule consensus
tree produced using PAUP.
Flies were obtained from a variety of locations, either trapped by the authors A further analysis was conducted using 13 individuals of Ch. rufifacies,
using liver-baited traps or kindly supplied by colleagues. They were identified using MEGA to perform neighbour-joining analysis with 500 bootstrap replica-
using traditional morphological characters. Specimens used in this study are tions. Hydrotaea rostrata was the assigned outgroup in all analyses, with the
listed in Table 1. exception of the Ch. rufifacies distance analysis.

2.2. DNA extraction


3. Results
DNA was extracted from the flight muscles of specimens using a DNEasy
Tissue Kit (Qiagen) according to manufacturer’s instructions, with an overnight A total of 47 individuals were sequenced and aligned over
incubation step. 1167 base pairs of the COI gene. A further 72 additional
70 M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76

Fig. 1. Bayesian inference tree constructed from 1167 bp of COI data. Posterior probabilities are indicated on nodes. Ch = Chrysomya; C = Calliphora;
Co = Cochliomyia; and L = Lucilia.

sequences were obtained from Genbank for comparative 3.1. Identification


purposes. The sequences correspond to positions 1776-2942
of the Drosophila yakuba mitochondrial genome (accession The muscid outgroup, Hydrotaea rostrata, was clearly
number NC_001322). No insertions or deletions were located separated from the calliphorids in the Bayesian Inference tree
over this region. Of the 388 variable positions identified, 318 of (Fig. 1). Furthermore, the calliphorid species were correctly
these were considered parsimony-informative. assigned to the sub-families Chrysomyinae, Luciliinae or
M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76 71

Calliphorinae, with each of the sub-families monophyletic. databases. A radial tree (Fig. 2) effectively illustrated the
Good posterior probability (PP) support for a sister-group subdivision within the cluster, with Malaysian and Taiwanese
relationship between the Calliphorinae and Luciliinae was individuals forming at least one group distinct from individuals
obtained, but support for a monophyletic Chrysomyinae was from Australia and the United States. Pairwise comparison
weak. indicated that when considering the species as two clades, one
The genus-level arrangement accurately reflected the containing the Malaysian and Taiwanese individuals, and the
affiliations of the species, and was supported by PP values remaining individuals as a group, 0.94% variation was
above 75%. observed.
At the species level, all of the species has PP support of over Pairwise calculation indicated a relatively high value of
94% except Ch. megacephala (paraphyletic, but with a large 3.94% variation across the two clades of L. cuprina. However,
internal clade with 83% support), L. sericata (81%), and Ch. variation within the two clades was much smaller (Table 2).
putoria (69%). Most specimens were accurately assigned to Unfortunately, insufficient sequences were available for a more
their respective species (Fig. 1). The exceptions were C. stygia detailed analysis.
and C. albifrontalis, which were intermingled; Ch. mega- Within each species, individuals from the same locality were
cephala, which formed a paraphyletic grade with respect to a intermixed with specimens from other sites (Fig. 1), and no
monophyletic (C. rufifacies + C. albiceps) and a monophyletic geographical patterns were obvious.
Ch. saffranea (Fig. 1) because of two Malaysian specimens
(AY909052 and AY909053); and Lucilia cuprina, which 3.3. Interspecific variation
formed two distinct clades that were collectively paraphyletic
with respect to L. sericata (Fig. 1). One clade of L. cuprina Levels of interspecific variation between calliphorid species
consisted of individuals from Australia, Senegal and Uganda, varied from 0.23 to 13.34% (Table 3). Species pairs such as Ch.
while the other represented Taiwan, Thailand and Hawaii. The rufifacies/Ch. albiceps and Ch. chloropyga/Ch. putoria were
latter group were more probably related to the L. sericata clade separated by 3.64% and 2%, respectively. Calliphora dubia and
than to conspecific individuals from the other clade. its sister species C. augur were clearly differentiated, yet
displayed only a 1.27% difference. Lucilia sericata differed
3.2. Intraspecific variation from L. cuprina by 2.8%, yet by only 0.93% from the Asian
clade. In general, species were separated by at least 3%, with
Where it could be calculated, all values for intraspecific the exception of some closely related pairings. Chrysomya
variation (Table 2) fell below 0.8%, with the exception of L. megacephala and Ch. saffranea differed by only 0.23%.
cuprina and Ch. rufifacies. Chrysomya rufifacies showed two
well-supported subgroups (Fig. 1). A separate neighbour- 4. Discussion
joining analysis of 722 bp of COI data for this species allowed
an analysis including more individuals from the public 4.1. Identification of taxa

Table 2 This study showed that the COI successfully distinguished


Maximum intraspecific variation expressed as a percentage of the total of 1167
base pairs of COI data
all but four of 27 species, even when conspecific specimens
were drawn from well-separated parts of the geographical
Species Maximum %variation within species distributions. Genera were clearly separated and the subfamilial
C. albifrontalis 0.26 arrangement reflected morphological findings commonly
C. stygia 0.60 reported in most taxonomic literature, with the Luciliinae
C. ochracea 0.17
and Calliphorinae grouping together and Cochliomyia posi-
C. dubia 0.26
C. augur 0.69 tioned basally within the Chrysomyinae.
C. vicina 0.26 The sister species Ch. putoria and Ch. chloropyga are
L. illustris 0.77 Afrotropical blowflies associated with latrines and carrion,
L. cuprina 3.94 respectively. The two were long treated as synonymous
Asian clade 0.48
despite evidence of ecological and reproductive separation
Afro-Australian clade 0.30
L. sericata 0.26 [19,20]. Despite recent recognition of their status as distinct
Co. macellaria 0.34 species [20], the morphological characters used to distinguish
Ch. putoria 0.60 them are of little use with eggs and early instars [19]. Wells
Ch. chloropyga 0.51 et al. [19] sequenced the two species over 593 bp of the COI
Ch. marginalis 0.26
gene, and maximum parsimony analysis determined Ch.
Ch. rufifacies 1.37
Asian clade 0.64 chloropyga to be paraphyletic with respect to Ch. putoria. In
Americo-Australian clade 0.18 this study, with twice as many characters, both species were
Ch. albiceps 0.51 monophyletic. Apparently the increased size of the dataset
Ch. varipes 0.77 contributed valuable distinguishing nucleotide information.
Ch. megacephala 0.34
Our results confirm the identity of Ch. putoria in South
Ch. saffranea 0.18
America (Fig. 1).
72 M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76

Fig. 2. Radial neighbour-joining tree based on 1167 bp of COI data for Chrysomya rufifacies individuals from a variety of localities.

The sister species Ch. rufifacies and Ch. albiceps are Wales) and reported only 0.4% variation between them across a
particularly distinct, despite their morphological similarity. variety of regions, including 822 bp of the COI gene. It may be
This is in line with an estimate that the rufifacies lineage is at that the mtDNA regions sequenced to date, which include
least 4 million years old [12]. They are usually placed together 3008 bp from the COI, COII, ND4 and ND4L genes [12], are
as the sister clade to the rest of their genus (e.g. [12]), and their not useful in distinguishing these species, perhaps as a result of
position in Fig. 1 may be the cause of the low PP support for relatively recent speciation: they are estimated to have
Chrysomya in our analysis. originated within the last million years [12]. Since this is
Chrysomya megacephala and C. saffranea are generally probably insufficient time to complete lineage sorting, it is not a
regarded as morphologically and genetically similar [21,22] but surprise that C. megacephala is paraphyletic with respect to C.
ecologically distinct [23]. In this study, their maximum saffranea. Funk and Omland [24] have assessed that 23% of
interspecific sequence variation was only 0.33% even though metazoan species are not molecularly monophyletic. Fortu-
COI is a relatively fast-evolving gene. By comparison, variation nately, the very low variation in C. saffranea makes this species
within C. megacephala was 0.34% (Table 2). Explanations for sufficiently distinctive to be identifiable. As discussed already,
apparent paraphyly include misidentification, hybridisation and the position of the (Ch. rufifacies + Ch. albiceps) clade within
incomplete lineage sorting. Neither hybridisation nor mis- C. megacephala is probably an artefact, and does not affect the
identification would produce the very low interspecific genetic identifiability of any of these species.
distance (Table 2) and tree topology (Fig. 1) found in this study. The genus Calliphora is well-represented by a clade of
Wallman et al. [12] compared one individual each of Ch. endemic species in the Australasian region, but it is difficult to
saffranea (Queensland) and Ch. megacephala (New South separate some of these based on molecular data. Using four
Table 3
Calculated raw interspecific distances using a neighbour-joining approach with Tamura-Nei model of substitution
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29
1 Ch.megacephala
2 Ch.saffranea 0.23

M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76


3 Ch.rufifacies 7.56 7.75
4 Ch.albiceps 6.9 7.07 3.64
5 Ch.chloropyga 6.12 6.04 8.79 8.23
6 Ch.putoria 5.21 5.13 8.05 7.42 2
7 Ch.marginalis 5.77 5.78 7.73 8.29 5.99 4.62
8 Ch.norrisi 5.75 5.76 8.04 10.44 7.49 6.27 6.02
9 Ch.pinguis 3.11 3.02 8.81 8.91 6.9 6.12 6.36 6.43
10 Ch.inclinata 5.4 5.32 9.28 8.59 4.2 3.73 6.02 6.95 6.13
11 Ch.bezziana 5.37 5.45 8.53 8.13 7.3 5.92 5.85 6.91 5.99 6.59
12 Ch.varipes 7.07 7.07 8.76 7.68 8.1 7.08 6.89 6.24 8.35 7.16 7.75
13 Co.hominivorax 10.25 10.23 12.81 11.71 11.29 10.95 11.93 10.96 11.93 10.91 11.16 11.97
14 Co.macellaria 8.74 8.77 12.04 11.4 10.18 8.94 9.13 8.56 8.73 8.99 9.4 10.11 8.65
15 C.stygia 12.06 12.05 13.29 13.19 12.54 11.79 12 10.95 12.78 12.02 12.51 12.33 12.17 11.56
16 C.albifrontalis 12.02 12.01 13.34 13.22 12.65 11.82 12.1 10.98 12.67 12.05 12.46 12.37 12.12 11.41 0.32
17 C.dubia 10.26 10.36 11.79 11.67 10.3 9.79 10.21 9.94 11.15 10.45 9.93 10.43 11.36 11.71 7.82 7.75
18 C.augur 10.43 10.59 11.76 11.92 10.59 10.27 10.61 10.41 11.52 10.72 10.45 10.72 11.72 12.25 8.23 8.31 1.3
19 C.vicina 10.52 10.51 11.66 11.3 10.08 9.61 10.05 10.19 10.98 9.8 9.98 11.38 10.73 12.19 9.2 9.13 7.69 8.37
20 C.vomitoria 10.85 10.63 11.8 11.75 9.88 9.34 10.12 9.53 12 9.97 10.3 11.99 11.28 11.91 8.31 8.22 6.82 7.47 4.49
21 C.varifrons 10.62 10.61 11.84 12.05 11.06 10.26 10.58 10.29 11.57 10.97 10.74 11.19 12.06 11.92 8.42 8.45 6.7 6.97 7.99 7.95
22 C.hilli 10.67 10.78 12.11 11.34 11.48 11.46 11.68 10.44 12.13 11.74 10.67 11.78 12.04 13.31 9.51 9.54 6.31 6.91 8.25 8.9 4.94
23 C.ochracea 11.84 11.83 12.14 11.24 11.73 10.94 11.46 10.68 12.45 12.19 11.7 11.99 12.43 12.97 8.59 8.56 8.03 8.49 8.82 8.74 8.25 8.06
24 L.sericata 8.79 8.82 11.72 10.97 10.44 9.71 8.69 9.05 10.18 9.99 10.19 10.29 11.18 9.64 9.85 9.81 9.06 9.61 8.67 8.46 9.53 10.59 9.98
25 L.cuprina 8.91 8.98 11.04 10.36 10.49 9.72 8.37 8.95 10.52 9.76 9.81 10.21 11.57 10.12 10.61 10.63 8.81 9.03 9.14 8.83 9.55 10.46 10.06 2.8
26 Asian clade 9.27 9.37 11.8 11 10.61 9.88 8.72 8.98 10.75 10.16 10.36 10.24 11.5 9.73 10.03 9.99 9.46 9.73 8.82 8.61 9.39 10.74 10.08 0.93
27 Afro-Australian clade 8.75 8.81 10.7 10.08 10.43 9.65 8.21 8.93 10.42 9.58 9.56 10.2 11.61 10.29 10.87 10.91 8.52 8.72 9.28 8.93 9.62 10.34 10.05 3.63 4.03
28 L.illustris 9.73 9.63 11.18 10.69 10.29 9.21 8.74 9.51 10.1 9.12 10.11 10.59 12.24 9.14 10.19 10.23 8.96 9.81 9.07 8.57 10.21 10.74 9.79 5.53 6.25 5.69 6.5
29 L.ampullacea 11.6 11.51 12.44 12.39 12.29 11.5 11.5 10.28 12.23 11.2 12.52 10.82 11.52 11.65 9.53 9.67 9.99 10.35 10.36 9.41 10.54 11.21 10.14 6.54 7.85 6.88 8.27 5.99
30 H.rostrata 16.97 17.12 15 14.42 17.65 16.89 16.53 15.16 17.66 17.94 17.23 16.46 17.24 17.03 15.89 15.97 15.7 15.6 16.64 16.22 16.09 16.27 16.12 15.74 15.97 16.03 15.94 16.69 16.06

73
74 M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76

genes, Wallman et al. [12] had difficulty separating members of Chrysomya rufifacies is primarily a tropical Australasian and
the C. hilli group (including C. varifrons), C. augur/C. dubia Oriental fly, but the species now also inhabits areas of the
and C. stygia/C. albifrontalis. In this study C. hilli and C. United States [25]. The behaviour of Ch. rufifacies is variable
varifrons were clearly distinguished with 4.94% interspecific across its distribution [5,25]. Such variable behaviour may be a
variation. It appears that the three additional genes used by result of ecological interaction with other carrion-colonising
Wallman et al. [12] obscured the signal from the COI gene, but species. It may also be that a species displaying a wide
our result may equally be due to the increase in size of the COI geographic distribution and variable behaviour might also
region we used. display correlated genetic variation.
Like the previous species, C. stygia and C. albifrontalis are Based on data from mitochondrial genes, Wallman et al. [12]
morphologically similar flies that speciated less than a million suggested that there are two sibling species within Ch.
years ago [12], and appear to be present together in certain rufifacies in Australia, one in the south-east and one in the
locations [4]. The COI data in this study did not successfully north-east. Our Western Australian individuals seem to be part
distinguish between the two species, which are mutually of the south-east taxon. However, combining the COI data from
polyphyletic. New regions of DNA such as the internal both studies (Fig. 2) showed that the Western Australian,
transcribed spacer (ITS) regions are more useful in distinguish- Tasmanian and United States individuals at most constitute
ing such closely related species (Harvey, unpublished data). only one slightly variable clade (Table 2). Our analyses (Figs. 1
The sister species Calliphora augur and C. dubia also and 2) provide evidence that the Malaysian and Taiwanese
formed monophyletic clades, although the latter received no PP individuals formed a separate cluster from the Australian/
support (Fig. 1). Wallman et al. [12] found a comparable pattern American clade. Clearly, the evidence for cryptic species within
of PP support, and estimated the pair to have diverged just over Australia must come from the other mitochondrial genes
a million years ago. These results imply that lineage sorting of sequenced by Wallman et al. [12]. Given the seminal nature of
COI in blowflies may take about two million years, and that a current molecular studies of the Calliphoridae, it is risky to
faster-evolving molecular marker is needed for younger form conclusions about species status from a single-gene
species. phylogeny and limited sampling, but the occurrence of an Asian
The sister species L. sericata and L. cuprina are clade and an Australian/American clade is also indicated by
morphologically similar, yet on the basis of this data are preliminary nuclear internal transcribed spacer (ITS) data
separated quite convincingly. High support was obtained for L. (Harvey, unpublished data).
sericata as distinct from L. cuprina. Lucilia cuprina is classified into two subspecies, L. c.
cuprina (Wiedemann) and L. c. dorsalis Robineau-Desvoidy
4.2. Geographical variation [26]. The former inhabits the New World, Asia, Indonesia, and
Oceania, while the latter is Afrotropical and Australasian
The molecular taxonomic facet of forensic entomology [27,28]. The two are found to readily interbreed in the
generally accumulates genetic data for a DNA region suitable laboratory, and hybrid populations are suggested to exist in
for identifying all species of forensic importance in a specific areas of Australia [26]. The observation of two paraphyletic
locality, or for global specimens of a focus taxon. Our results clades of L. cuprina in this study could be explained by genetic
show that geographical variation was not more pronounced in variation between subspecies. The clade from Australia and
species sampled from larger geographical distributions. For Africa would represent L. c. dorsalis, and the Pacific
example, Ch. albiceps and Ch. megacephala show levels of individuals L. c. cuprina.
intraspecific variation comparable with C. augur, C. stygia and Wallman et al. [12] suggested based on mitochondrial genes
C. chloropyga (Table 2). Furthermore, no geographical pattern that their Western Australian L. cuprina may be L. c. dorsalis,
of relationship was evident in most species (Fig. 1), except Ch. while their New South Wales and Queensland individuals
rufifacies and L. cuprina. Part of the explanation for this may be represented L. c. cuprina. They rightly indicated that, given the
that most carrion-breeding flies are synanthropic, and are history of hybrid populations on the east coast of Australia, this
spread by human activities. This is illustrated by the occurrence cannot be concluded with certainty, but suggest that cuprina
of European species like C. vicina in Tasmania and L. sericata and dorsalis lineages are separate species, an interpretation
in New Zealand, and the African species Ch. putoria in South supported by the paraphyly found in their analyses and ours. In
America (Fig. 1). This is reassuring for the use of COI our study, the Townsville, Queensland individual appears to be
sequences in identification of blowflies, because it implies that L. c. dorsalis, so that it is likely that both taxa are resident in
most of the variation within a species can be captured by Queensland and extensive sampling and study is required, given
relatively localised samples. the tendency of the variants to interbreed.

4.3. Cryptic species 5. Conclusions

The geographical structuring of the C. rufifacies and L. The molecule-based identification of calliphorids relies on
cuprina clades is linked with unusually high intraspecific the location of unique stretches of DNA sequence that are
variation that may be suggestive of the presence of currently common to (at least sub-sets of) all members of the chosen
unrecognised taxa. taxon, yet distinct from all other taxa. This necessitates
M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76 75

extensive sequencing of conspecific individuals to verify the of incomplete lineage sorting. This may necessitate the
robust nature of markers chosen as species identifiers. This adoption of a secondary assay for distinction of such species,
study indicates that most of the variation within a species may perhaps the ITS regions.
be adequately captured by samples of as few as 10 Sequencing of individuals from a wide variety of localities,
geographically distant conspecific individuals. This may be and thus the pooling of data from various researchers, will
partly attributed to the high mobility of blowflies. greatly contribute to the taxonomy of calliphorids on a global
Some species show more intraspecific variation than scale. Molecular forensic entomology, while in its infancy, has
others, and the discovery of an outlying sample must address increased the ability to identify unknown immature calliphor-
the question of whether it is an extreme example, or if it ids. It is the extensive sampling of populations and more species
represents an otherwise unsampled species. It is tempting to that will ultimately contribute to the increased relevance of the
seek an empirical heuristic threshold [29,30], but the field to the courtroom.
differentiation of species on predetermined thresholds of
genetic divergence relies on the degree of overlap between Acknowledgements
intraspecific variation and interspecific divergence. While
thresholds of intraspecific variation and interspecific diver- We thank numerous people for their generous donation of
gence have been used in other studies to delimit species flies for this study: Geoff Allen, Dallas Bishop, Amoret Brandt,
boundaries and infer novel species, this is a fraught practice M. Lee Goff, Ana Carolina Junqueira, Gary Levot, Nicola Lunt,
where taxa are undersampled [29], as they are in the Mervyn Mansell, Nolwazi Mkize, Carol Simon, Tony Postle,
Calliphoridae. Moritz and Ciccero [30] state that threshold Cameron Richards, Kabew Sukontason and Ruxton Villet.
overlap is greater where a large proportion of closely related Jennifer Roy provided technical assistance. Some field work
species are included, such as the Australian Calliphora was funded by Rhodes University grants to MHV. MLH was
species. In a review of the use of percent DNA sequence supported in part by Rotary International, and the American
difference across the COI in insects, Cognato [31] revealed Australian Association.
that such thresholds may fail up to 45% of the time when
used to diagnose species due to overlap in inter and References
intraspecific sequence divergences. Harvey et al. [6] and
Wells and Sperling [32] reported levels of 0.8% maximum [1] F.A.H. Sperling, G.S. Anderson, D.A. Hickey, A DNA-based approach to
intraspecific variation and 3% minimum interspecific the identification of insect species used for postmortem interval estima-
divergence in the Calliphoridae, but in this study these tion, J. Forensic Sci. 39 (1994) 418–427.
[2] Y. Malgorn, R. Coquoz, DNA typing for identification of some species of
thresholds overlap. Calliphoridae: an interest in forensic entomology, Forensic Sci. Int. 102
It is suggested that forensic entomologists heed the (1999) 111–119.
published cautions against the use of percent DNA sequence [3] J. Stevens, R. Wall, Genetic relationships between blowflies (Calliphor-
divergence in species delimitation. A concerted effort in the idae) of forensic importance, Forensic Sci. Int. 120 (2001) 116–123.
[4] J.F. Wallman, S.C. Donnellan, The utility of mitochondrial DNA
field to gather the necessary data to sufficiently sample the
sequences for the identification of forensically important blowflies (Dip-
relevant taxa, identify the limitations of the COI and locate tera: Calliphoridae) in southeastern Australia, Forensic Sci. Int. 120
alternate regions for identification in recently diverged taxa will (2001) 60–67.
allow the relationships revealed in molecular phylogenies to [5] M.L. Harvey, I.R. Dadour, S. Gaudieri, Mitochondrial DNA cytochrome
distinguish individuals. The designation of novel species is not oxidase I gene: potential for distinction between immature stages of some
likely based solely on molecular bases. Paraphyly and forensically important fly species (Diptera) in western Australia, Forensic
Sci. Int. 131 (2003) 134–139.
polyphyly may provide an indication of some species process, [6] M.L. Harvey, M.W. Mansell, M.H. Villet, I.R. Dadour, Molecular identi-
however, they may likewise represent an inappropriate gene fication of some forensically important blowflies of southern Africa and
choice for the specific focus taxa, and further study will be Australia, Med. Vet. Entomol. 17 (2003) 363–369.
necessary. Unpublished data (Harvey) has indicated that the [7] S.D. Ratcliffe, D.W. Webb, R.A. Weinzierl, H.M., Robertson. PCR-RFLP
identification of Diptera (Calliphoridae, Muscidae and Sarcophagidae)—a
COI para/polyphyly observed in some calliphorid taxa may be
generally applicable method, J. Forensic Sci. 48(4) (2003) 1–3.
resolved through an alternate gene choice, and it is such study [8] J. Stevens, R. Wall, Genetic variation in populations of the blowflies
that indicates the importance of thorough global study of Lucilia cuprina and Lucilia sericata (Diptera: Calliphoridae). Random
calliphorids of interest if principles of monophyly are to be amplified polymorphic DNA analysis and mitochondrial DNA sequences,
enforced in identification. Biochem. Systematics Ecol. 25 (1997) 81–97.
This study has again illustrated the enormous potential for [9] S. Vincent, J. Vian, M.P. Carlotti, Partial sequencing of the cytochrome
oxidase b subunit gene I: a tool for the identification of European species
the use of the COI gene for distinguishing between forensically of blow flies for postmortem interval estimation, J. Forensic Sci. 45 (2000)
significant calliphorid species, considering these species from a 820–823.
global perspective and linking allopatric populations of species [10] D.H. Lunt, D.-X. Zhang, J.M. Szymura, G.M. Hewitt, The insect cyto-
often considered in isolation. This study has supported the chrome oxidase I gene: evolutionary patterns and conserved primers for
existence of sibling species within L. cuprina, and illustrated phylogenetic studies, Insect Mol. Biol. 5 (1996) 153–165.
[11] C. Garros, R.E. Harbach, S. Manguin, Systematics and biogeographical
the potential unsuitability of the COI gene for distinction implications of the phylogenetic relationships between members of the
between young species like Ch. saffranea and Ch. mega- funestus and minimus groups of Anopheles (Diptera: Culicidae), J. Med.
cephala, and C. stygia and C. albifrontalis, probably as a result Entomol. 42 (2005) 7–18.
76 M.L. Harvey et al. / Forensic Science International 177 (2008) 66–76

[12] J.F. Wallman, R. Leys, K. Hogendoorn, Molecular systematics of Aus- [23] H. Kurahashi, Dispersal of filth flies through natural and human agencies:
tralian carrion-breeding blowflies (Diptera: Calliphoridae) based on mito- origin and immigration of a synanthropic form of Chrysomya megace-
chondrial DNA, Invertebrate Systematics 19 (2005) 1–15. phala, in: M. Laird (Ed.), Commerce and the spread of pests and disease
[13] P.D.N. Hebert, M.Y. Stoeckle, T.S. Zemlak, C.M. Francis, Identification of vectors, Praeger, New York, 1984.
birds through DNA barcodes, PLoS Biol. 2 (2004) e312. [24] D.J. Funk, K.E. Omland, Species-level paraphyly and polyphyly: fre-
[14] C. Simon, F. Frati, A. Beckenbach, B. Crespi, H. Liu, P. Flook, Evolution, quency, causes, and consequences, with insights from animal mitochon-
weighting, and phylogenetic utility of mitochondrial gene-sequences and drial DNA, Annu. Rev. Ecol., Evol. Systematics 34 (2003) 397–423.
a compilation of conserved polymerase chain-reaction primers, Ann. [25] D.L. Baumgartner, Review of Chrysomya rufifacies (Diptera: Calliphor-
Entomol. Soc. Am. 87 (1994) 651–701. idae), J. Med. Entomol. 30 (1993) 338–352.
[15] E.H. Harley, Dapsa (DNA and Protein Sequence Alignment), Department [26] K.R. Norris, Evidence for the multiple exotic origin of Australian popula-
of Chemical Pathology, University of Cape Town, Cape Town, South tions of the sheep Blowfly, Lucilia cuprina (Wiedemann) (Diptera:
Africa, 1996. Calliphoridae), Aust. J. Zool. 38 (1990) 635–648.
[16] S. Kumar, K. Tamura, I.B. Jakobsen, M. Nei, MEGA2: Molecular [27] B. Montgomery, Elucidation of the evolutionary status of Queensland
Evolutionary Genetics Analysis software, Bioinformatics 17 (2001) coastal Lucilia cuprina (Wiedemann). Master of Science Thesis. Uni-
1244–1245. versity of Queensland (1990).
[17] D.L. Swofford, PAUP* Phylogenetic Analysis Using Parsimony (*and [28] J. Stevens, R. Wall, Species, sub-species and hybrid populations of the
other methods). Version 4, Sinauer Associates, Sunderland, Massachu- blowflies Lucilia cuprina and Lucilia sericata (Diptera: Calliphoridae),
setts, 2002. Proc. Roy. Soc. Lond. B 263 (1996) 1335–1341.
[18] J.P. Huelsenbeck, F. Ronquist, MRBAYES. Bayesian inference of phy- [29] C.P. Meyer, G. Paulay, DNA barcoding: error rates based on comprehen-
logeny, Bioinformatics 17 (2001) 754–755. sive sampling, PLoS Biol. 3 (2005) 2229–2238.
[19] J.D. Wells, N. Lunt, M.H. Villet, Recent African derivation of Chrysomya [30] C. Moritz, C. Ciccero, DNA barcoding: promise and pitfalls, PLoS Biol. 2
putoria from C. chloropyga and mitochondrial DNA paraphyly of cyto- (2004) 1529–1531.
chrome oxidase subunit one in blowflies of forensic importance, Med. Vet. [31] A.I. Cognato, Standard percent DNA sequence difference for insects
Entomol. 18 (2004) 445–448. does not predict species boundaries, J. Economic Entomol. 99 (2006)
[20] K. Rognes, H.E.H. Paterson, Chrysomya chloropyga (Wiedemann, 1818) 1037–1045.
and Chrysomya putoria (Wiedemann, 1830) (Diptera, Calliphoridae) are [32] J.D. Wells, F.A.H. Sperling, DNA-based identification of forensically
two different species, African Entomol. 13 (2005) 49–70. important Chrysomyinae (Diptera: Calliphoridae), Forensic Sci. Int. 120
[21] N. Evenhuis (Ed.), Catalog of the Diptera of the Australasian (2001) 110–115.
and Oceanian Regions, Bishop Museum Special Publication, 1989 , [33] W.-Y. Chen, T.-H. Hung, S.-F. Shiao, Molecular identification of foren-
p. 86. sically important blow fly species (Diptera: Calliphoridae) in Taiwan, J.
[22] F.H. Ullerich, M. Schöttke, Karyotypes, constitutive heterochromatin, and Medical Entomol. 41 (2004) 47–57.
genomic DNA values in the blowfly genera Chrysomya, Lucilia, and [34] J.R. Stevens, The evolution of myiasis in blowflies (Calliphoridae), Int. J.
Protophormia (Diptera: Calliphoridae), Genome 49 (6) (2006) 584–597. Parasitol. 33 (2003) 1105–1113.

Você também pode gostar