Sei sulla pagina 1di 20

Received: 4 October 2018 

|  Revised: 10 December 2018 


|
  Accepted: 10 December 2018

DOI: 10.1111/zsc.12338

ORIGINAL ARTICLE

On the importance of fine‐scale sampling in detecting alpha


taxonomic diversity among saproxylic invertebrates: A velvet
worm (Onychophora: Opisthopatus amaxhosa) template

Aaron Barnes  | Savel R. Daniels

Department of Botany and


Zoology, University of Stellenbosch,
Abstract
Matieland, South Africa The phylogeography and social structure of the narrow endemic velvet worm species
Opisthopatus amaxhosa were investigated by conducting fine‐scale sampling in its
Correspondence
Savel R. Daniels, Department of Botany distribution range in the Eastern Cape province of South Africa. In addition, and as
and Zoology, University of Stellenbosch, part of larger grant on forest biodiversity, Opisthopatus specimens sampled at locali-
Matieland, South Africa.
ties not included during a recent evaluation of the genus were included in a new
Email: srd@sun.ac.za
phylogeny. A total of 89 specimens from 18 sample localities were collected at three
Funding information forest patches for O. amaxhosa samples, while an additional six Opisthopatus sam-
National Research Foundation
ple localities were included. For O. amaxhosa, we sequenced the COI locus for all
specimens, while a subset of specimens was sequenced for two nuclear loci, 18S
rRNA and the fushi tarazu intron (FTz). Phylogenetic analyses using maximum like-
lihood and Bayesian inferences of the latter species revealed the presence of two
highly divergent clades, characterised by marked uncorrected sequence divergence
values. In addition, these two clades did not share any maternal haplotypes, were
characterised by high FST values and fixed nuclear difference for the 18S rRNA locus,
while the FTz intron was genetically invariant. Furthermore, the application of scan-
ning electron microscopy between the two genetically divergent clades also revealed
the presence of fixed ventral and dorsal scale numbers. Collectively, this provides
evidence for a novel species that is present at a fine scale. Divergence time estima-
tions suggest that the two clades diverged during the late and early Pleistocene with
climatic cycling potentially causal to the fragmentation. The social structure was
male‐biased, and samples from the same logs were not always genetically identical.
At the broader scale, the inclusion of new specimens within Opisthopatus revealed
no novel lineages. Fine‐scale sampling appears more important to detect alpha taxo-
nomic diversity compared to broadscale sampling.

KEYWORDS
conservation, fine‐scale, forest fragmentation, novel species, phylogeographic, population‐level

© 2019 Royal Swedish Academy of Sciences     1


Zoologica Scripta. 2019;1–20. wileyonlinelibrary.com/journal/zsc |
|
2       BARNES and DANIELS

1  |   IN TRO D U C T ION restricted to areas of high precipitation (Rutherford, Mucina,


& Powrie, 2006). Velvet worms are confined to saproxylic
Saproxylic environments are defined as decaying habitat environments such as among leaf litter and under decaying
such as logs of wood or leaf litter that are present in for- logs and are good indicators of habitat quality (Daniels et
ested regions. These habitats harbour a diverse array of in- al., 2009; Daniels & Ruhberg, 2010; Hamer, Samways, &
vertebrates that are critical in ecosystems function such as Ruhberg, 1997; McDonald & Daniels, 2012). Two velvet
nutrient recycling forming an important link in the terrestrial worm genera, Peripatopsis and Opisthopatus, are present in
food chain (Majka & Pollock, 2006). Generally, the saprox- South Africa (Hamer et al., 1997). Historically, the latter two
ylic invertebrate fauna are habitat specialists, require specific genera contained eight and three species, respectively, with
microclimatic regimes for survival and frequently exhibit four of the species being IUCN Red Listed (Hamer et al.,
low vagility (Daniels & Ruhberg, 2010; Myburgh & Daniels, 1997). However, renewed recent interest in the group has re-
2015). The result is high levels of localised endemism and sulted in an increased focus on intensive molecular analysis
marked genetic differentiation, rendering these species prone and scanning electron microscopy (SEM). This resulted in
to localised extinction (Anderson, 1994; Seibold et al., 2015; the discovery and description of nine and five new species
Williams & Pearson, 1997). Due to their life history char- within Peripatopsis and Opithopatus, respectively (Daniels
acteristics, saproxylic invertebrates are ideal non‐model or- et al., 2016, 2013; Ruhberg & Daniels, 2013). New IUCN
ganisms with which to test hypotheses related to dispersal Red Listing is planned for the recently discovered fauna con-
and colonisation in fragmented forest habitat. In Australia, sidering the narrow endemic distribution of several of the
Europe and North America, phylogeographic studies of sap- novel species (S. R. Daniels, personal communication). Since
roxylic invertebrates have revealed limited dispersal among velvet worm taxa are often cryptic, fine‐scale phylogenetic
conspecific populations, marked population differentiation study of velvet worms is necessary to understand the popula-
and cryptic speciation (Cicconardi, Nardi, Emerson, Frati, tion genetic structure and the importance of saproxylic hab-
& Fanciulli, 2010; Garrick et al., 2006; Gaublomme et al., itat conservation, especially, given the high risk of localised
2013; Jonsson, Kruys, & Ranius, 2005; Ranius, 2006). extinction.
Further, phylogeographic studies of saproxylic taxa have in- Within South Africa, population‐level molecular studies
vestigated the historical population dynamics and assessed have provided insight into the dispersal abilities of velvet
the impact of demographic drivers such as recolonisation, worms, generally revealing high levels of population differ-
extinction and population bottleneck events (Garrick et al., entiation linked to both ancient and contemporary change
2004). Collectively, these studies have argued for the con- (Daniels, 2011; Daniels et al., 2017; Myburgh & Daniels,
servation of both the taxa and the habitat of saproxylic in- 2015). Daniels (2011) demonstrated low haplotypic diver-
vertebrates. Nevertheless, in the Afrotropics, Neotropics and sity in the narrow endemic rose pink velvet worm species
Mesoamerica, saproxylic invertebrate fauna are poorly stud- Opisthophatus roseus. Similarly, in the velvet worm species
ied and their conservation remains limited partially due to a Peripatopsis overbergiensis, distributed in three allopatric
paucity of taxonomic studies. forest fragments in the Overberg district of the Western Cape
The isolation of saproxylic environments within for- province, Myburgh and Daniels (2015) demonstrated the
est patches results in similar microclimatic pressures, pro- absence of maternal and paternal dispersal, corroborating a
moting convergent evolution (Harmon, Kolbe, Cheverud, general trend of high genetic differentiation among conspe-
& Losos, 2005) resulting in cryptic speciation (Daniels, cific populations (Barclay, Rowell, & Ash, 2000; Daniels,
Dambire, Klaus, & Sharma, 2016; Daniels, Picker, Cowlin, 2011; Daniels et al., 2017). It is thought that velvet worm
& Hamer, 2009; Daniels & Ruhberg, 2010; Trewick, 1997). dispersal occurs at a natal stage and is male bias, offering a
Consequently, the alpha taxonomy of most saproxylic taxa possible theory for the apparent female bias in group com-
is poor and these groups exhibit high cryptic differentiation. position (Daniels, 2011; Daniels et al., 2017). Although dis-
This highlights the importance of molecular systematic re- persal plays a role in shaping group composition, it does not
search to document the biodiversity of saproxylic taxa. One exclusively govern population divergence or the lack thereof.
group of saproxylic taxa in South Africa that was recently Rather migratory dynamics are influenced by physiological
subjected to intense systematic scrutiny is the velvet worm limitations and population size variations which can be di-
fauna (Daniels et al., 2016, 2009; Daniels, Dreyer, & Sharma, rectly related to forest fragment size and habitat availability
2017; Daniels, McDonald, & Picker, 2013; McDonald & (Daniels, 2011; Daniels et al., 2017; Myburgh & Daniels,
Daniels, 2012; Myburgh & Daniels, 2015). Onychophora, or 2015). It is therefore hypothesised that a lack of vagility and
velvet worms, are soft‐bodied invertebrates, generally con- female bias will be observed among social groups within logs
fined to saproxylic habitats within forests. The South African and that marked genetic structure will be present where for-
forest biome is the smallest of the seven biomes covering ests are fragmented. Further, it is expected that genetic differ-
<0.5% of the countries land surface, is highly fragmented and entiation will increase with increased spatial sampling. The
BARNES and DANIELS   
|
   3

social structure in velvet worms has rarely been considered as distance between sampled sites. Hence, the genetic diversity
a driver for dispersal. Sex ratio and life histories are two im- should increase with increased spatial sampling. The second
portant components and are determined largely by behaviour aim is to understand the sex ratio, size and composition of
which is more complex than previously conceptualised, as O. amaxhosa specimens within logs in relation to their ge-
observed by Reinhard and Rowell (2005). Demographic netic composition and to make inferences on the social struc-
variability has been observed between the sexes in the vel- ture of the species. We hypothesise that the sex ratio should
vet worm species Peripatoides novaezealandiae, where a fe- be female bias and that the sampled logs contain both closely
male bias sex ratio was observed (Tutt, Daughtery, & Gibbs, related maternal lineages as well as genetically highly differ-
2001). At birth velvet worms exhibit a 1:1 sex ratio, how- entiated specimens.
ever, observations of free‐living animals showed a female
bias ratio among sexually mature specimens, a pattern that
has been observed in several studies (Barclay et al., 2000;
2  |  M ATERIAL S AND M ETHO D S
Curach & Sunnucks, 1999; Daniels, 2011; Sunnucks et al.,
2.1  |  Taxon sampling
2000; Tutt et al., 2001; Weldon, Daniels, Clusella‐Trullas,
& Chown, 2013). Therefore, the sex ratio of a population is During April 2017, a total of 78 O. amaxhosa specimens
possibly dependent on the life stage and the time of sampling were collected from three forest fragments (Baziya, Jenca
(Tutt et al., 2001). and Nqadu) in the Eastern Cape province of South Africa
Opisthopatus amaxhosa, one of five recently de- (Figure 1). These samples were combined with 11 O. amax-
scribed velvet worm species, is confined to three allopatric hosa specimens used by Daniels et al. (2016) yielding a total
Afromontane forest fragments around Baziya, Langeni and of 89 specimens. Samples were hand collected from decay-
Nqadu in the Eastern Cape province of South Africa (Daniels ing logs, and coordinates were recorded using a hand GPS
et al., 2016). Notably, these forests have historically been (Garmin Trek Summit). The number of adults and juveniles
very poorly sampled, and the species was originally de- were counted along with the type of social grouping deter-
scribed based on 11 specimens. The habitat of O. amaxhosa mined as either “family” or “random” during sample collec-
is fragmented and bisected by lower lying dry corridors of tion (Table 1). A “family” was defined as a group comprised
grassland that potentially act as barriers to dispersals. In addi- of two adults and a large number of juveniles, while a “ran-
tion, recent plantations of Pinus and Eucalyptus have further dom” group is comprised of two or more adults without any
restricted the indigenous forest patches to deep valleys and juveniles occurring in the same log. Specimens were con-
gorges. Furthermore, anthropomorphic factors such as sub- sidered adult when they were >1 cm in length and sexually
sistence sheep and cattle farming by the local communities mature, while juveniles were <1 cm in length and sexually
have also contributed to the fragmentation and degradation immature. Samples were placed in plastic jars and killed by
of these forests, where dead wood is frequently collected preservation in absolute ethanol and stored at 4°C in a refrig-
for household usage (S. R. Daniels, personal communica- erator. A general collection permit was issued for sampling
tion). Considering the patchy nature of these indigenous (code CRO 60/17 CR). In addition, we included specimens
Afrotemperate forests, O. amaxhosa forms the ideal template of Opisthopatus from four new localities (Hogsback, Mboyti,
taxon with which to test the impact of forest fragmentation on Mazeppa Bay and Umslanga Nature Reserve) that was un-
the phylogeographic structure of the species. sampled by Daniels et al. (2016) to determine their phylo-
A fine‐scale phylogeographic analysis of O. amaxhosa genetic placement among the species described by the latter
has not been conducted to date, limiting our understanding author.
of the genetic structure and conservation of the saproxylic
environments. In addition, as part of a larger research grant
2.2  |  DNA extraction, PCR and sequencing
on the biodiversity of forests in the Eastern Cape province,
we included Opisthopatus sampled from areas that were not Tissue biopsies were taken from the lateral right‐hand side
sampled by Daniels et al. (2016) to explore the presence of specimens, without cutting across the mid‐dorsal line,
of additional novel species within the genus. The present preserving morphologically diagnostic characters. DNA was
study aims to examine the phylogeography and sex ratio of extracted using a Macherey‐Nagel DNA extraction kit follow-
O. amaxhosa at three spatial scales: within a decaying log, ing the manufacturer's protocol. Extracted DNA was stored
between decaying logs within a forest and between the three at 4°C in a refrigerator. DNA extractions were diluted with
major aforementioned forest patches where the species oc- deionised water making a 1 µl DNA in 19 µl of water solution.
curs. Considering the limited dispersal capability of the velvet Microsatellite development remains notoriously difficult
worms and its preference for specific microclimatic factors, in Onychophorans (Sands, Lancaster, Austin, & Sunnucks,
we hypothesise marked genetic differentiation, with the de- 2009). Further, failures in amplification, high‐frequency null
gree of genetic differentiation increasing with geographic alleles and failures in microsatellite cloning have led to the
|
4       BARNES and DANIELS

F I G U R E 1   Localities where Opisthopatus amaxhosa was collected in the Eastern Cape province of South Africa. The numbers on the map
correspond to the sample localities outlined in Table 1. The colours represent the altitudinal gradients of the region as described in the map legend
[Colour figure can be viewed at wileyonlinelibrary.com]

suggestion that the problem lies in the genomes of the group CAAA AAA TCA‐3′) were used to amplify the COI locus
rather than in the method (Sands et al., 2009). Thus, the use of (Folmer, Black, Hoeh, Lutz, & Vrijenhoek, 1994). The primer
microsatellite loci was not considered in this study. Instead, pairs 5F (5′‐GCG AAA GCA TTT GCC AAG AA‐3′) and 7F
three partial gene fragments were amplified and sequenced (5′‐GCATCA CAG ACC TGT TAT TGC‐3′) were used for
during the present study. These included the cytochrome c the 18S rRNA subunit (Giribet, Carranza, Baguna, Riutort,
oxidase one subunit (COI) and two nuclear DNA loci, the & Ribera, 1996). For the FTz intron, the forward (5’‐GTG
small ribosomal 18S subunit (18S rRNA) and the fushi tarazu AGG CTG CTGG AAAAATT‐3’C) and reverse (3’‐CAAA
(FTz) intron (Sands et al., 2009). The COI locus has been TCC GTTGTC TGT AAAT ATG‐5’) primer pairs were used
extensively used for population‐level analyses among vel- (Sands et al., 2009). The 18S rRNA subunit has been used as
vet worms (Daniels et al., 2013, 2009; Daniels & Ruhberg, complimentary to the COI subunit when determining velvet
2010). All samples were sequenced for the COI locus, while worm species boundaries (Daniels et al., 2016, 2017, 2013,
a subset of samples was sequenced for the two nuclear DNA 2009; McDonald & Daniels, 2012; Myburgh & Daniels,
(18S rRNA and the fushi tarazu ‐ FTz intron) loci. The primer 2015). The DNA sequences generated for the COI locus
pairs LCOI‐1490 (5′‐GGT CAA CAAATCA TAAA GAT for the O. amaxhosa specimens generated by Daniels et al.
ATTG‐3′) and HCOI‐2198 (5′‐TAAA CTT CAGGGT GAC (2016) were combined with sequences generated during the
BARNES and DANIELS   
   5
|
T A B L E 1   List of Opisthopatus
Locality no. Name N Social group Coordinates
amaxhosa specimens collected within the
Baziya forest complex in the Eastern Cape 1 BaziyaOLD* 5 N/A 31.31.250 S 28.24.738 E
province of South Africa 2 Nocu* 6 N/A 31.24.928 S 28.29.990 E
3 BaziyaJan 3 N/A 31.34.034 S 29.24.733 E
4 Baziya A 3 Random 31.34.034 S 29.24.733 E
5 Baziya B 5 Family 31.34.095 S 28.24.689 E
6 Baziya C 7 Family 31.33.892 S 28.24.601 E
7 Baziya D 3 Random 31.33.890 S 28.24.594 E
8 Baziya E 6 Family 31.33.914 S 28.24.580 E
9 Baziya F 4 Random 31.32.901 S 28.26.391 E
10 Baziya G 10 Random 31.34.671 S 28.24.671 E
11 Baziya H 4 Family 31.34.293 S 28.24.585 E
12 Baziya I 5 Family 31.34.312 S 28.24.712 E
13 Baziya J 3 Family 31.33.511 S 28.25.369 E
14 Baziya L 2 Random 31.34.885 S 28.23.696 E
15 Ndlukulu 2 Random 31.31.251 S 28.24.693 E
16 Langeni* 2 Random 31.31.252 S 28.24.692 E
17 Nqadu A 9 Family 31.42.839 S 28.73.539 E
18 Nqadu B 10 Family 31.42.840 S 28.73.539 E
Notes. N, the number of samples. N/A implies the designation for the social group was not determined.
Locality numbers correspond to those in Figure 1. The asterisk (*) denotes samples used by Daniels et al. (2016).

present study. Similarly, for the larger phylogenetic study, computed with the use of CLUSTAL X (Thompson, Gibson,
the newly sampled Opisthopatus specimens were included Plewniak, Jeanmougin, & Higgins, 1997). The phylogeny for
with those from Daniels et al. (2016). Opithsopatus and O. amaxhosa was constructed using maxi-
Polymerase chain reactions (PCRs) were conducted on a mum likelihood (ML) and Bayesian inference (BI). The ML
geneAmp PCR System Thermo cycler (Applied Biosystems, analyses were performed using the CIPRES Science Gateway
Foster City, CA, USA) under the following conditions for the version 3.3 (Miller, Pfeiffer, & Schwartz, 2010), using
COI locus: denatured for 94°C for 4 min, 94°C for 30 s, 42°C GARLI 2.01 (Zwickl, 2006) on XSEDE. A single replicate
for 40 s, 72°C for 45 s, for 36 cycles, with a final extension at search was conducted for the best tree. Branch support values
72°C for 10 min. For the 18S rRNA locus and FTz intron, the were estimated using a bootstrap analysis (Felsenstein, 1985)
annealing temperatures were 50°C and 52°C, respectively, with 1,000 pseudoreplicates and parameters set to default.
the number of cycles increased to 40. All PCRs were con- Bootstrap values >75% were deemed as sufficient statistical
ducted for a 25 µl reaction and held at 15°C. All three re- support for nodes. Analyses were performed using a heuristic
actions contained 14.9 μl of deionised water, 3 μl of 25 mM search algorithm starting at a random tree. Four rate catego-
MgCl2, 2.5 μl of 10 × Mg2+ free buffer, 0.5 μl of a 10 mM ries were included for gamma, and base frequencies were es-
dNTP solution and 0.5 μl of the primer sets at 10 mM, 0.1 timated. Bootstrap values were discerned via a 50% majority
unit of Taq polymerase and 1–3 μl of extracted DNA. PCR rule tree, generated using PAUP* 4.0 (Swofford, 2002).
products were gel purified in a 1% agarose gel over 180 min Bayesian analyses were conducted on the CIPRES Science
at 90 V. Gel bands were explicit under UV light, were re- Gateway version 3.3 (Miller et al., 2010) with MrBayes ver-
sected and purified using a BioFlux gel purification kit. sion 3.2.6 (Ronquist et al., 2012) on XSEDE. Each analysis
Sequence‐purified PCR products were then cycled under comprised of one run of four chains for 10 million genera-
standard protocols. An ABI 3,730 XL automated machine tions, sampling every 1,000 generations using default param-
conducted the sequencing. eters. A random tree was selected for the start of each chain.
Burn‐in was included in the command block, set to 20%. A
consensus tree was generated by MrBayes on CIPRES as part
2.3  |  Phylogenetic analyses
of the output. The percentage of time spent on node recovery
Forward and reverse strands were used to compute a consensus was used to estimate the posterior probabilities (pP) for each
sequence and to check for base ambiguities using Sequence node. Posterior probabilities (pP) <0.95 were regarded as
Navigator (Applied Biosystems). Sequence alignment was poorly supported. For both analyses, the Akaike information
|
6       BARNES and DANIELS

criterion (AIC) (Akaike, 1973) ascertained the best‐fit substi- substitution model for the entire COI subset. Ten million
tution model using jModelTest2 (Posada & Crandall, 1998) generations were run for 10 independent MCMC chains,
on XSEDE through CIPRES (Miller et al., 2010). PAUP*10b with sampling performed every 1,000 generations. After
(Swofford, 2002) computed the uncorrected sequence diver- appropriate burn‐in, the convergence of the 10 combined
gence for the COI locus. chains was determined by effective sample size for each
parameter in tracer (Rambaut, Suchard, & Drummond,
2013). Trees were amalgamated using logcombiner which
2.4  |  Phylogeographic analyses sourced from
contained 10 chains, after which they were assessed
COI data for O. amaxhosa
using treeannotator. The program figtree version 1.4.3
The haplotype network was constructed using TCS 1.21 (Rambaut, 2009) was used to formulate a chronogram.
using 95% confidence interval (Clement, Posada, & Crandall,
2000). Population genetic structure was computed on arle-
2.6  |  Morphological examination using SEM
quin version 3.5.1.2 (Excoffier & Lischer, 2010) using the
for O. amaxhosa
COI data. A hierarchical analysis of molecular variance
(AMOVA) was conducted over all sample localities, and Live O. amaxhosa specimens were photographed using
then for each of the two COI clades evident from the prelimi- a Canon EOS 70D camera. The colour was noted, and the
nary analyses. FST values were computed across all sample number of oncopods was counted. Five specimens from each
localities. Standard diversity indices were further calculated of the two clades (observed during the preliminary phylo-
at each of the localities within fragments. These indices in- genetic analyses) were examined using a SEM. Specimens
cluded haplotype diversity (h), the number of polymorphic were dehydrated firstly in absolute ethanol, followed by par-
sites (Np), the number of haplotypes (Nh) and nucleotide di- tial dehydration in a 1:1 solution of hexamethyldisilazane
versity (π). Fu's Fs (Fu, 1997) was used to clarify deviations (HMDS) and absolute ethanol for 30 min. Specimens were
from neutralities in allele frequencies due to its sensitivity to finally dehydrated in 100% HMDS in a glass petri dish for
population equilibrium fluctuations. Fu's Fs is preferred over an additional 30 min, before being left to dry inside a drying
Tajima's D due to its higher sensitivity and significance level oven at 25°C overnight. Images were captured using a Zeiss
of p < 0.02 (Fu, 1997). MERLIN Field Emission SEM at the Electron Microbeam
Unit of Stellenbosch University's Central Analytical Facility
in the Department of Geology. Prior to imaging, the sam-
2.5  |  Divergence time estimation for
ples were mounted on aluminium stubs with double‐sided
O. amaxhosa
carbon tape. The samples were then gold coated to 10 nm,
The complete COI locus dataset was used for estimating di- using an Edwards S150A Gold Sputter Coater. Beam condi-
vergence times between haplotypes despite of known faults tions during surface analysis were 5 kV and approximately
within the locus (Daniels, 2011; McDonald & Daniels, 250 nÅ, with a working distance of 6 mm. During the SEM,
2012). The 18S rRNA and FTz intron loci were not con- we focussed on analysis of the dorsal and ventral integu-
sidered for the divergence time estimations due to a lack ment structure since previous studies successes fully used
of mutation rates for the latter two nuclear loci and be- these characters to delineating cryptic velvet worm species
cause only a subset of data was generated for these two (Daniels et al., 2013; McDonald, Ruhberg, & Daniels, 2012;
loci (Clouse et al., 2015; Daniels et al., 2016; Fernandez Oliveira, Read, & Mayer, 2012; Ruhberg & Daniels, 2013).
& Giribet, 2014). Divergence time estimations for the COI Scanning electron micrograph images were examined for dif-
locus were conducted using a Bayesian framework, which ferences in the number of scale rings for the primary and ac-
makes use of a probabilistic model to define the molecular cessory papillae on both the dorsal and ventral integument
sequence divergence of lineages which further makes use surface.
of the MCMC method to estimate clade ages. A relaxed
molecular clock was used (Drummond, Ho, Phillips, &
2.7  |  Population sex ratio
Rambaut, 2006), and the analyses were run using the pro-
gram beast version 2.4.8 (Drummond & Rambaut, 2007). The sex of O. amaxhosa specimens was determined using
The mutation rates of 1.5%–2.3% per million years were standard light microscopic examination (Hamer et al., 1997).
used for the COI locus (Boyer, Baker, & Giribet, 2007; In addition, the total length (TL), measured from the anterior
Brower, 1994; Farrell, 2001; Trewick & Wallis, 2001), to the posterior end, and the diameter of the body (DB) each
with a mean mutation rate of 1.9% per million years specimen in millimetre (mm) were recorded using a digital
being selected (McDonald & Daniels, 2012). A multiple vernier calliper for all ethanol fixed samples. The subsequent
coalescent model was used (Heled & Drummond, 2010). results were analysed for significance of ratio for each sex
MODELTEST was used to determine the parameters and within each locality using a t test (Quinn & Keough, 2002).
T A B L E 2   Haplotype (H) frequency distribution for the COI locus for Opisthopatus amaxhosa sampled from the Baziya forest complex in the Eastern Cape province of South Africa

Baziya Baziya Baziya Baziya Baziya Baziya Baziya Baziya Baziya Baziya Baziya Nqadu Nqadu
BaziyaOLD Nocu BaziyaJan A B C D E F G H I J L Ndlukulu Langeni A B
H1 1
H2 1
BARNES and DANIELS

H3 1
H4 2
H5 1
H6 1
H7 1
H8 1
H19 7 9
H10 2
H11 1
H12 1 3 1
H13 1 4 4 1 1
H14 1
H15 2
H16 1
H17 1
H18 1
H19 1
H20 1
H21 1
H22 1
H23 1
H24 1
H25 1
H26 1 5 5 3 5
H27 1 1 1 1
H28 1
H29 2 1
H30 2
H31 1
  

H32 2
|   7

Note. The haplotype numbers correspond to Figure 4.


|
8       BARNES and DANIELS
BARNES and DANIELS   
   9
|
F I G U R E 2   Bayesian inference tree topology for Opisthopatus amaxhosa derived from the mtDNA COI locus. The nodal values below
branches represent the bootstrap values (>75%) for ML. Posterior probability (pP) values for BI are shown above branches (>0.95pP). Values
below the prescribed threshold are indicated by an asterisk for ML (*) and a hashtag (#) for BI

3  |   R E S U LTS sequence divergence was 9.34%, while within clade 2, the


maximum uncorrected sequence divergence was 4.09%.
3.1  |  Phylogenetic analyses of the COI data Within clade 2, the basal specimen labelled Baziya A4,
for Opisthopatus amaxhosa identified as O. kwazululandi was 18.36% divergent from
the remainder of the samples in clade 2. The inclusion of
A 610‐base pair fragment of the COI locus was amplified
a genus‐wide analysis representing both the COI and 18S
and sequenced for the 78 O. amaxhosa specimens and com-
genes for Opisthopatus clarified that the sample Baziya A4
bined the 11 specimens from Daniels et al. (2016) yielding a
is in fact part of and not sister to the species complex of
total of 89 specimens. The novel sequences were deposited
O. kwazululandi, hence the highly differential uncorrected
in GenBank (Accession numbers MK236395–MK236470).
sequence divergence and the classification of this sample
Both the BI and ML analyses produced near identical tree
as a third sympatric species.
topologies, and hence, only the BI topology is shown. The
DNA substitution model selected using the AIC criteria
was TIM2 + I + G (−1nl = 3,106.46; AIC = 6,592.91)
3.2  |  Population genetics for the COI locus
(Akaike, 1973). The base frequencies were A = 33.87%,
C = 17.88%, G = 12.29% and T = 35.97%. Similar results A total of 32 haplotypes were retrieved for the 89 COI se-
of an A/T rich base frequency bias have been reported in quences using TCS analyses (Table 2; Figure 4). The hap-
other velvet worm studies (Daniels et al., 2009; Daniels lotype network corroborated the phylogenetic analyses and
& Ruhberg, 2010, McDonald & Daniels, 2012). The rate retrieved two haploclusters. Haplocluster 1 comprised of 20
matrix was R(a) [A‐C] = 5.44, R(b) [A‐G] = 22.30, R(c) haplotypes while clade 2 comprised the remaining 13 hap-
[A‐T] = 5.44, R(d) [C‐G] = R(f) [G‐T] = 1.00 and R(e) lotypes. Due to the classification of the sample Baziya A4
[C‐T] = 40.54, the proportion of invariable (I) sites was as a separate species, O. kwazululandi (haplotype 3), it was
0.5320 and the gamma (G) shaped distribution was 0.6990. excluded from the network representation of the popula-
The BI topology (Figure 2) retrieved two statistically well‐ tion‐level analyses for O. amaxhosa (Figure 3). The lack of
supported (>75% and 1.00 = pP) clades. Clade 1 consisted connectivity and shared haplotypes between clades 1 and 2
of all the O. amaxhosa specimens described by Daniels et is indicative of the absence of gene flow (Table 3). Within
al. (2016), plus sample localities from Nqadu A, Nqadu B, haplocluster 1, only populations in close geographic proxim-
Langeni, Ndlukulu, Baziya C, E and BaziyaJan while clade ity were connected indicating limited maternal dispersal at a
2, sister group to Opisthopatus kwazululandi, comprised small spatial scale. In contrast, the large number of unsam-
specimens from Baziya A, B, C, D, E, F, G, H, I, J and L. pled or missing mutational steps between haplotypes in hap-
At BaziyaJan, Baziya C and Baziya E, the two clades were locluster 2 suggests a similar lack of dispersal between more
sympatric and genetically diverged indicating maternal distantly related sample sites (Table 3).
isolation of the two clades. Divergence time estimations The AMOVA results over all O. amaxhosa samples re-
indicate that the two clades diverged 2.82 million years ago vealed that 80.04% variation occurred among populations
(95% highest posterior density [HPD]: 1.94–3.69 Mya) (to- (Va = 26.29, df = 17, s.s. = 2,287.00, p < 0.01) while
pology not shown). These results indicate that the diver- 19.96% variation occurred within populations (Vb = 6.56,
gence of O. amaxhosa occurred during the late Pliocene df = 71, s.s. = 465.67, p < 0.01). Within clade 1, 78.14% of
to early Pleistocene. The updated phylogenetic analyses of the variation occurred among populations (ΦST = 0.78135,
the newly collected Opithopatus specimens (Figure 3) did Va = 14.24, df = 8, s.s. = 464.57, p < 0.01) while 21.86%
not reveal any novel lineages, except for those from the variation occurred within populations (Vb = 4.01, df = 28,
Baziya forest complex (Figure 2). Further the tree topol- s.s. = 112.34, p < 0.01). Within clade 2, 59.88% of the varia-
ogy presented (Figure 3) placed the novel lineage within tion occurred among populations (ΦST = 0.59883, Va = 4.14,
the species complex of O. kwazululandi, this is most likely df = 11, s.s. = 223.48, p < 0.01) while 40.12% of the vari-
due to the inclusion of the conserved 18S rRNA locus. The ation occurred within populations (Vb = 2.77, df = 40, s.s.
result from the BI analysis including only the COI locus re- 110.85, p < 0.01). Pairwise FST values across all sample lo-
vealed the novel species as sister to the existing combined calities indicate marked genetic structure, suggesting limiting
clade of O. kwazululandi and O. amaxhosa (Figure 3). maternal dispersal (Table 3). The number of haplotypes (Nh),
The uncorrected sequence distance between clades 1 and number of polymorphic sites (Np), haplotype diversity (h),
2 was 19.50%. Within clade 1, the maximum uncorrected nucleotide diversity (π) and Fu's Fs for the 18 sample sites
|
10       BARNES and DANIELS

F I G U R E 3   Bayesian inference tree topology derived from the combined COI and 18S data for the Opisthopatus genus, containing one
sample from each study site. The nodal values below branches represent the bootstrap values (>75%) for ML. Posterior probability (pP) values for
BI are shown above branches (>0.95pP). Values below the prescribed threshold are indicated by an asterisk for ML (*) and a hashtag (#) for BI.
New sample localities are highlighted in blue and the single sample from the novel clade of the present study is highlighted in red. The abbreviation
“NR” depicts whether the locality is a nature reserve [Colour figure can be viewed at wileyonlinelibrary.com]
BARNES and DANIELS   
|
   11

F I G U R E 4   The haplotype networks derived from the 32 COI haplotypes demonstrating two haploclusters corresponding to the phylogenetic
tree in Figure 2 for Opisthopatus amaxhosa. The numbers within circles represent the haplotype number and correspond to Table 2. Solid black
circles represent missing or unsampled haplotypes [Colour figure can be viewed at wileyonlinelibrary.com]

are summarised in Table 4. The sample locality Baziya A GenBank accession numbers for the 10 FTz intron specimens
had the highest nucleotide diversity (Nh), compared to the sequenced were MK236476‐MK236485. For the 18S rRNA
remaining samples. This was followed in order of descent by locus, a 427‐bp fragment was amplified and sequenced for
BaziyaOLD, Baziya C and Baziya E. The highest number of two specimens from each of the clades 1 and 2, respectively
polymorphic sites (Np) corresponded to the sample localities (GenBank Accession numbers MK236486–MK236489).
where the two clades co‐occurred. Fu's Fs values contained Within clade 1, the observed uncorrected sequence distance
a mixture of negative and positive values with a bias towards was 0.23%, while within clade 2, the observed uncorrected
the latter. The positive values are representative of population sequence distance was 9.83%. Between clades 1 and 2, the
bottlenecks or overdominant selection, shown by a deficiency uncorrected distance was 12.18%.
of alleles. The negative values are characterised by an excess
in the number of alleles and indicate a recent population ex-
3.4  |  Morphological examination using SEM
pansion or genetic hitchhiking (Fu, 1997). The latter result
is present at three localities: Baziya G, Nqadu A and Nqadu Scanning electron microscopy revealed morphological di-
B. At the clade level, Fu's Fs values were both positive, in- agnostic differences between the two O. amaxhosa clades
dicative of a reduction in gene flow and a potential historical in the number of scale rings for the primary dermal papilla
bottleneck. on the dorsal and ventral integumental areas. Opisthopatus
amaxhosa (clade 1) exhibited six rings on the dorsal primary
papillae and four rings on the ventral primary papillae, while
3.3  |  nu DNA (FTz intron and 18S rRNA)
specimens in clade 2 exhibited eight rings on the dorsal pri-
For the FTz intron, a 250‐bp fragment was sequenced; how- mary papillae and six rings on the ventral primary papillae
ever, the intron was genetically invariant between the two (Figure 5). All specimens had 16 pairs of oncopods, and the
clades, and hence, this locus was not considered further. The dorsal integument was rose pink with white and burgundy
T A B L E 3   Pairwise FST values for the COI locus across all 18 of the different sample localities for Opisthopatus amaxhosa
|

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
12      

1
0.29097
2
0.38228 0.55472
3
0.80131 0.87931 0.16442
4
0.73653 0.87931 −0.05270 0.18328
5
0.83631 0.91338 0.32533 0.77143 0.22680
6
0.65688 0.74874 0.03454 0.35771 −0.08709 −0.04551
7
0.78792 0.89003 0.14231 0.70714 0.03846 0.00000 −0.15711
8
0.62762 0.73058 −0.00085 0.35666 −0.08485 −0.03448 −0.17321 −0.15385
9
0.80697 0.89447 0.15751 0.68326 0.38381 0.89813 0.15832 0.86047 0.16355
10
0.89504 0.94069 0.50181 0.12586 0.70575 0.96942 0.61871 0.96394 0.62650 0.94237
11
0.75787 0.85217 0.03795 0.31834 −0.18567 0.33594 −0.02331 0.18725 0.00423 0.28148 0.73734
12
0.72986 0.83883 −0.08420 0.13114 −0.33613 0.37824 −0.04797 0.20000 −0.03190 0.24638 0.67729 −0.36565
13
0.70533 0.83391 −0.26599 −0.21495 −0.44134 0.60396 −0.02848 0.41463 −0.03643 0.37173 0.58148 −0.27121 −0.56818
14
−0.00372 0.47071 0.32709 0.91976 0.87940 0.99385 0.71101 0.98978 0.67890 0.97654 0.98759 0.88486 0.86493 0.86500
15
0.46952 0.72579 0.44415 0.92371 0.88582 0.99603 0.71641 0.99339 0.68338 0.98039 0.98881 0.89055 0.87220 0.87440 0.94624
16
0.72152 0.84400 0.79886 0.96857 0.96549 0.99726 0.86141 0.99671 0.85494 0.99187 0.99309 0.95995 0.96089 0.97099 0.98192 0.98866
17
0.74113 0.85643 0.81667 0.97214 0.97004 0.99872 0.87085 0.99849 0.86546 0.99391 0.99433 0.96459 0.96587 0.97550 0.98803 0.99300 0.00454
18
BARNES and DANIELS

Note. Values in bold are statistically significant (p < 0.005). Localities 1–18 correspond to those in Table 1.
BARNES and DANIELS   
   13
|
T A B L E 4   Summary list of the population genetic parameters measured using the COI locus in Opisthopatus amaxhosa

No. of haplotypes No. of polymorphic Haplotype


Locality N (Nh) sites (Np) diversity (h) Nucleotide diversity (π) Fu's Fs
BaziyaOLD* 5 5 56 1.0000 ± 0.1265 0.054098 ± 0.033428 1.08892
Nocu* 6 4 28 0.8667 ± 0.1291 0.026995 ± 0.016234 3.71423
BaziyaJan 3 3 109 1.0000 ± 0.2722 0.121858 ± 0.091558 3.20102
Baziya A 3 3 25 1.0000 ± 0.2722 0.027322 ± 0.021085 1.67558
Baziya B 5 1 0 0.0000 ± 0.0000 0.000000 ± 0.000000 N/A
Baziya C 7 3 106 0.5238 ± 0.2086 0.051444 ± 0.029388 10.36994
Baziya D 3 1 0 0.0000 ± 0.0000 0.000000 ± 0.000000 N/A
Baziya E 6 2 107 0.3333 ± 0.2152 0.058470 ± 0.034414 13.48405
Baziya F 4 2 3 0.6667 ± 0.2041 0.003279 ± 0.002750 2.19722
Baziya G 10 4 3 0.7778 ± 0.0907 0.001785 ± 0.001443 −0.69898
Baziya H 4 3 26 0.8333 ± 0.2224 0.023224 ± 0.015839 3.21925
Baziya I 5 2 24 0.4000 ± 0.2373 0.015738 ± 0.010184 6.77643
Baziya J 3 3 27 1.0000 ± 0.2722 0.030601 ± 0.023530 1.79304
Baziya L 2 2 24 1.0000 ± 0.5000 0.039344 ± 0.040156 3.17805
Langeni* 2 2 3 1.0000 ± 0.5000 0.004918 ± 0.005679 1.09861
Ndlukulu 2 2 2 1.0000 ± 0.5000 0.003279 ± 0.004016 0.69315
Nqadu A 9 3 2 0.4167 ± 0.1907 0.000729 ± 0.000806 −1.08110
Nqadu B 10 2 1 0.2000 ± 0.1541 0.000328 ± 0.000499 −0.33931
Note. The number of samples (N), haplotype number (Nh), number of polymorphic sites (Np), haplotype diversity (h), nucleotide diversity (π) and Fu's Fs of the separate
populations. Localities marked with an asterisk were used by Daniels et al. (2016).

dermal papillae (Figure 6) except for a single specimen from uncorrected COI distances where sympatric, clearly indi-
Baziya A4, which was slate black. cating genetic isolation. In addition, the two clades can fur-
ther be differentiated based on the nuclear 18S rRNA locus
and exhibit fixed morphological differences, based on the
3.5  |  Population sex ratio
number of scale rings for the primary papillae using SEM.
A male bias was observed across all sample localities with Collectively, these results suggest the presence of an addi-
females being found only present at five sample localities tional novel, as yet undescribed species within O. amaxhosa
(p < 0.05). Juveniles were only present at four sample lo- sensu lato. All the specimens originally used by Daniels et
calities (Baziya B, C, E and I). Social groupings classified al. (2016) to describe O. amaxhosa are assigned to clade 1,
as families did not comprise of a pair of single‐sex adults suggesting the specimens in clade 2 represent a novel op-
and juveniles, suggesting that social aggregations are ran- erational taxonomic unit. These results clearly indicate the
dom. Female specimens were longer and wider compared importance of fine‐scale sampling in an attempt to uncover
to male specimens; however, this trend was non‐statistically the alpha diversity among saproxylic taxa, a fact that is often
significant (p > 0.05). Morphological data are summarised in disregarded when taxonomic studies are designed despite its
Table 5. critical importance.
Our first hypothesis of observing marked levels of genetic
differentiation, increasing with the degree of spatial separa-
4  |   D IS C U SSION tion across three spatial scales, that is, within a decaying log;
between decaying logs within a forest; and between the three
The present study clearly indicates that at a large spa- major forest patches, was partially supported within the two
tial scale we did not detect any novel lineages (Figure 3). observed clades. Within each of the two clades, the general
However, at a fine spatial scale, our phylogeographic analy- trend exhibited was limited maternal dispersal between con-
ses provide evidence that two statistically well‐supported specific sample localities. We can reject our second hypoth-
clades are present within O. amaxhosa (Figure 2). The two esis of a female bias in specimens, since we observed a male
clades within O. amaxhosa are characterised by marked bias in our sampling.
|
14       BARNES and DANIELS

(a) (c)

(b) (d)

F I G U R E 5   Scanning electron micrographs of primary and accessory dermal papillae for the two Opisthopatus amaxhosa clades. Each white
dot represents a scale ring. (a) Dorsal papilla of O. amaxhosa. (b) Dorsal papilla of clade 2. (c) Ventral papilla of O. amaxhosa. (d) Ventral papilla
of clade 2. The abbreviation “br” is for the sensory bristle

the presence of two distinct evolutionary lineages present


4.1  |  Evidence for a new species
within what has hitherto been considered a single species,
We observed a clear paraphyly with clade 2 being sister to with clade 2 representing the novel lineage and clade 1 rep-
O. kwazululandi, while clade 1 is monophyletic. Species resenting the nominal O. amaxhosa species. Considering
paraphyly can be attributed to several factors including that these two clades are sympatric at sites Baziya C, E and
poor taxonomy, introgression (interspecific hybridisation) BaziyaJan, we can confidently define them as distinct lin-
and incomplete lineage sorting (Funk & Omland, 2003). eages. Within clade 2, a single slate black specimen from
Differentiation between the possible causes of species para- site Baziya A had an uncorrected sequence divergence of
phyly can be difficult as introgression and incomplete lineage 18.36% for the mtDNA COI locus compared to the remain-
sorting may result in similar gene tree topologies and for a der of clade 2. Under further investigation at a broad scale,
brief period in evolutionary time, diverging lineages remain this specimen was observed to fall within the O. kwazulu-
incompletely sorted making it near impossible to differenti- landi clade, forming a sister group to the species lineage
ate from poor taxonomy when there are insufficient synapo- from Weza Forest which also contains black specimens
morphies (Funk & Omland, 2003; McKay & Zink, 2010). (Daniels et al., 2016) (Figures 2 and 3). This was supported
The paraphyly and the proximity of genetic relatedness are by both the COI and the 18S markers, grouping the spec-
further exhibited in Figure 3 whereby the novel lineage from imen as a sister species to clade 2 of this study (Figures 2
this study is grouped with O. kwazululandi. This result is and 3). The marked uncorrected sequence distance values
most likely due to the inclusion of the conserved 18S rRNA between the two large clades as well as the distance be-
locus within the analyses, and a lack of differentiation be- tween the slate black specimen falls within the region of
tween the two species at a nuclear level. what has been used by Daniels et al. (2016) to differentiate
Multiple lines of independent evidence (both genetics Opisthopatus species. This level of divergence of the COI
and morphology) are congruent and provide evidence for locus is high and similar studies investigating Opisthopatus
BARNES and DANIELS   
|
   15

Kotze, & Lawes, 2017; Hughes, Möller, Bellstedt, Edwards,


& Villiers, 2005). Aridification of the area suggest that once
continuous forest were forced into refugia among the shel-
tered valleys of the adjacent escarpment, thus creating islands
of suitable habitat for onychophoran taxa. These refugia are
characterised and limited by their conformity to favourable
Afromontane forest conditions such as increased precipita-
tion and cooler temperatures when compared to the surround-
ing grasslands, which enhance refugia limitation through fire.
The grasslands create a fundamental‐realised niche effect
whereby the forest is driven further into refugia than what
climate permits, augmenting isolation. Thus, refugia would
have been local centres from which the habitat would expand
once climate during the Holocene produced wetter conditions
(Eeley, Lawes, & Piper, 1999; Lawes, 1990).
The extent of forest contractions and expansions for the
greater area must also be considered due to the inclusion of
the sample from Baziya A4 and its identification as part of
the O. kwazululandi species complex. Due to the observed
limited dispersal of the animals, it is highly unlikely for
movement to take place at such a large scale without suit-
able habitat connectivity. Therefore, an assumption can be
F I G U R E 6   Live photographic images of Opisthopatus made that the forests between the sample sites including the
amaxhosa collected at the Baziya forest in the Eastern Cape province specimens from those such as Weza and Ixopo containing
of South Africa. Scale bar = 1 cm [Colour figure can be viewed at O. kwazululandi were once more continuous and connected
wileyonlinelibrary.com] to those in Baziya. The Eastern Cape is hypothesised to be
a place of Afrotemperate forest refuge during glacial maxi-
sister species with a maximum sequence divergence of mums due to the combination of climate and the surrounding
19.27% (Daniels et al., 2016). These two clades could fur- geography (Hughes et al., 2005; Lawes, 1990; Lawes, Eeley,
ther be differentiated based on the nuclear 18S rRNA data, Findlay, & Forbes, 2007). Hughes et al. (2005, Lawes (1990,
further validating the presence of two distinct species. We and Lawes et al. (2007 noted a trend of the recolonisation of
caution against assigning species recognition solely on se- Afrotemperate patches and associated animal and plant gen-
quence divergent values and on the basis of a single molec- era radiating north‐east into KwaZulu‐Natal, sourcing from
ular marker. However, our result is also corroborated by the the southern refugia found in the Eastern Cape. Species with
18S rRNA locus. Morphologically, the dorsal integument low vagility such as O. kwazululandi must have had their
colour and the leg pair numbers have been shown to be lim- movement facilitated by the extended distributions of forest
ited as diagnostic characters between cryptic Opisthopatus patches, enhancing connectivity. Assessing the topology of
species (Daniels et al., 2016). However, SEM examination the area, it is most likely forests followed a route along the
of the primary papillae, on the dorsal and ventral integ- escarpment of the Natal Drakensburg mountains, remaining
ument, corroborated the genetic differentiation observed in areas of suitable climate refuge when conditions were ap-
with the DNA sequence data (Figure 5). propriate (Lawes, 1990), facilitating migration events and
the co‐distribution of O. kwazululandi.
A trend of a reduction in species richness was also ob-
4.2  |  Population genetic structure
served among various taxa as the distance increased from
The divergence time analysis indicates that the cladogenesis the refugial sites in the Eastern Cape (Lawes et al., 2007).
between the two O. amaxhosa clades occurred during the late Thus, refugial expansion and a reduction in species richness
Pliocene to early Pleistocene, 2.82 Mya. This time period is with a northward trend may provide further evidence for an
characterised by intense aridification and sea‐level reduc- expansion of O. kwazululandi into KwaZulu‐Natal from the
tions induced largely by temperature variations and conse- Eastern Cape refugia, especially under conditions of higher
quent rainfall regime alterations during glaciation cycles connectivity. The impact of the refuge as a sink for biodi-
(Van Andel, 1989; Lawes, 1990). Forest fragmentation was versity during climate ameliorations may alter how we con-
an end result, with the climate effects being assisted by the ceive the origin of species and may explain the observed high
evolution of C4 grasses and subsequent fire regimes (Adie, level of diversity at Baziya. Thus, it is possible that the sites
|
16      

T A B L E 5   Summary of the morphological characters of specimens examined from the 14 Opisthopatus amaxhosa sample localities of the current study

Mean Mean Mean


female juvenile juvenile
Locality Male (N) Mean male length Mean male width Female (N) length Mean female width Juvenile (N) length width Social grouping
Baziya A 3 19.16 2.92 Random
Baziya B 4 17.25 2.87 1 13.42 1.90 Family
Baziya C 2 21.42 3.33 1 26.02 3.27 3 13.70 1.65 Family
Baziya D 3 21.99 2.66 Random
Baziya E 5 19.43 2.87 2 13.29 2.52 Family
Baziya F 5 22.47 3.13 Random
Baziya G 8 31.36 2.69 3 28.95 2.16 Random
Baziya H 3 16.99 2.23 1 13.12 2.12 Family
Baziya I 3 18.98 3.05 1 14.27 2.67 1 10.63 0.58 Family
Baziya J 3 20.53 3.63 Family
Baziya L 2 17.85 2.36 1 21.38 2.89 Random
Ndlukulu 2 14.20 2.26 Random
Langeni 2 11.72 1.97 Random
Total 45 21.13 2.80 7 23.6 2.56 7 13.1 1.78
Note. All measurements are in millimetres (mm). Number of samples = N.
BARNES and DANIELS
BARNES and DANIELS   
|
   17

in Baziya experienced the enforced sympatry of O. kwazu- Where there is gene flow among populations, it is only
lulandi and O. amaxhosa due to its historical role as a ref- among adjacent populations thus representing limited but ex-
uge. The subsequent fine‐scale cryptic speciation between istent migration in clade 1 and a near similar lack of disper-
O. amaxhosa and the novel lineage likely occurred during the sal in clade 2 (Figure 4). This may be due to the proximity
early Pleistocene/late Pliocene climate oscillations as a re- of sample sites, for example, the clustering of only adjacent
sult of increased aridification. Thus, a phenomenon of forest sampling localities within the haplotypes 28 and 30. Due to
patch isolation at two levels is hypothesised with broadscale limited dispersal, the p distances within the clades are mod-
forest contractions taking place between the two provinces, erate. When compared, clade 1 showcased uncorrected diver-
isolating organisms from O. kwazululandi within Baziya, fol- gence of 9.34% within the clade and clade 2 showcased an
lowed by the fine‐scale contractions within the Baziya forest uncorrected divergence of 4.09%. This value is vastly higher
complex itself, which then facilitated the origin of O. amax- than values found for another Opisthopatus species (O. ro-
hosa and the novel lineage showcased as clade 2 in this study. seus) analysed at relatively fine scale in the study by Daniels
The two species from clades 1 and 2 follow a north to south (2011) however is skewed by the presence of the second spe-
trend with O. amaxhosa (clade 1) being found frequently in the cies within the clade, O. kwazululandi. Clade 1 exhibits a
northern sample localities and clade 2 being more frequently lack of expansive haploclusters coupled with a high number
found in the southern localities (Figures 1 and 2). This trend of private alleles which is supportive of a lack of maternal
may hint at how the region followed the pattern of forest con- dispersal at relatively small ranges. This is presumably due
traction–expansion and to which fragment each of the species to unsuitable environmental conditions induced by fragmen-
was assigned. The lack of current dispersal, as indicated by the tation with migration routes being compromised via anthro-
significant FST values between populations within clades of the pogenic influence.
main forest patches, is suggestive of short distance isolation,
inducing high genetic structure (Avise, 2000). The AMOVA
4.3  |  Population sex structure
results showed that more variation occurred among popula-
tions rather than within them further indicating the high levels The sex ratio did not exhibit the typical female bias ob-
of isolation between them. This isolation has been induced by served in the literature (Barclay et al., 2000; Curach
recent anthropogenic reductions in forest size and the frag- & Sunnucks, 1999; Sunnucks et al., 2000; Tutt et al.,
mentation of larger patches (Cawe & McKenzie, 1989; Eeley 2001). Sunnucks et al. (2000) observed that logs inhab-
et al., 1999; Hughes et al., 2005), limiting the suitability of ited by few velvet worms typically had a male bias due
areas between microhabitats, inhibiting migration (Barclay et to recent migrations, in contrast logs with more individu-
al., 2000; Woodman, Cooper, & Haritos, 2007). This historical als showed a female bias. This pattern did not hold with
climatic reconnection followed by anthropogenic detachment our results; however, it may explain the lack of females
is further supported by the high, positive Fu's Fs values, in- found since the sample sizes averaged at four individuals.
dicative of a population bottlenecks after becoming disjointed, Further, male bias is often attributed to newly colonised
suggesting that even at localities with close proximity, there is logs (Curach & Sunnucks, 1999); however, the positive
no contemporary maternal gene flow between them. Recent Fu's Fs values for the majority of the sample sites suggest
change due to fragmentation is supported by the samples from population bottlenecks. Therefore, the lack of females is
Nqadu which represent the highest sequence divergence in probably a result of stochasticity or the time of sampling.
clade 1, suggesting that the fragment has recently separated. The timing of sampling plays a role in the observed sex
Nqadu (including both localities) being the most isolated for- ratios and the presence of juvenile numbers since velvet
est patch displayed the lowest haplotype and nucleotide diver- worms typically exhibit complex life histories and behav-
sity (Table 4) indicative of population instability. These were iour which are largely dependent on environmental condi-
the only sites apart from Baziya G that had a negative Fu's tions and cues. The reproductive cycle in velvet worms
Fs value, indicative of a population expansion (Fu, 1997). is largely associated with their microhabitat, which can
Nqadu therefore contains two break‐off populations which be different within each of the sample localities; thus, the
harbour four separate haplotypes with high amounts of disper- number of observed juveniles is dependent on these condi-
sal within the patch but are however unstable. The pairwise tions (Monge‐Nájera 2015). Therefore, there is very little
FST results for the sample localities at Nqadu were significant to be extrapolated from the number of juveniles observed
when compared to every other locality from the present study. other than a recent lack of suitable birthing conditions,
These two sites are thus the only two of those sampled from since females have the ability to store sperm (Sunnucks et
O. amaxhosa that showcase little limitation on their dispersal al., 2000). Although females have been observed to reject
capabilities within the forest patch and seem to have migrated migratory males (Reinhard & Rowell, 2005), the aggrega-
recently but with enough elapsed time to be captured as a pri- tion composition within a log is evident to be largely influ-
vate haplocluster. enced by the abiotic environment while migratory patterns
|
18       BARNES and DANIELS

appear to be determined by climate rather than social in- assistance with tree development and design as well as for the
centive and acceptance. live photography of the animals. Ilse Boonzaaier is thanked for
help with GIS and map design. The environmental affairs de-
partment of the Cacadu district is thanked for issuing collection
4.4  |  Conservation implications
permits.
The onychophora as a habitat specialist group is negatively
impacted by the removal of suitable forest habitat, and the
ORCID
decaying logs that collect on the forest floor. To date, stud-
ies involving the conservation of velvet worms in South Savel R. Daniels  https://orcid.org/0000-0003-2956-3256
Africa have analysed the genetic structure and impacts on
genetic diversity of velvet worms at a broad scale, assess-
ing separate forest patches and species (Daniels & Ruhberg, R E F E R E NC E S
2010; McDonald & Daniels, 2012; Myburgh & Daniels, Adie, H., Kotze, D. J., & Lawes, M. J. (2017). Small fire refugia in
2015). Only a single study has focused on O. roseus in the the grassy matrix and the persistence of Afrotemperate forest in
Ngele forest and detected shallow genetic structure for the the Drakensberg mountains. Scientific Reports, 7, 6549. https://doi.
species (Daniels, 2011). During the present study, the use of org/10.1038/s41598-017-06747-2
fine‐scale sampling detected the presence of a second line- Akaike, H. (1973). Information theory and an extension of the maximum
likelihood principle. In B. N. Petrov, & F. Csaki (Eds.), Proceedings
age where historically only a single species was known to
of the 2nd International Symposium on Information Theory (pp.
occur. Within each lineage a large number of distinct hap-
267–281). Budapest, Hungary: Akademiai Kiado.
lotypes were present suggesting limited maternal gene flow Anderson, S. (1994). Area and endemism. The Quarterly Review of
among conspecific localities of both species. This result sug- Biology, 69, 451–471. https://doi.org/10.1086/418743
gests that at a smaller spatial scale, decaying wood logs in Avise, J. C. (2000). Phylogeography: The history and formation of spe-
forests are important units for conservation. The main result cies. Cambridge, MA, London, UK: Harvard University Press.
evident from the present study is the need for intensive sam- Barclay, S., Rowell, D., & Ash, J. (2000). Pheromonally mediated
pling in forested areas since the biodiversity of saproxylic colonization patterns in the velvet worm Euperipatoides rowelli
(Onychophora). Journal of Zoology, 250, 437–446. https://doi.
fauna remains poorly studied. The presence of three sympa-
org/10.1111/j.1469-7998.2000.tb00787.x
tric species within Baziya reinforces the need for intensive Boyer, S. L., Baker, J. M., & Giribet, G. (2007). Deep genetic divergences in
sampling within forest complexes for appropriate diversity Aoraki denticulata (Arachnida, Opiliones, Cyphophthalmi): A wide-
detection. In terms of velvet worms and saproxylic taxa with spread ‘mite harvestman’ defies DNA taxonomy. Molecular Ecology,
low vagility, it would be logical to follow a fine‐scale ap- 16, 4999–5016. https://doi.org/10.1111/j.1365-294X.2007.03555.x
proach when making efforts towards identification for con- Brower, A. V. (1994). Rapid morphological radiation and convergence
servation but also for studies assessing the social and genetic among races of the butterfly Heliconius erato inferred from pat-
terns of mitochondrial DNA evolution. Proceedings of the National
interactions between these taxa. Hence, we suggest fine‐scale
Academy of Sciences, 91(14), 6491–6495. https://doi.org/10.1073/
sampling in areas where velvet worms have previously been
pnas.91.14.6491.
poorly collected to determine the biodiversity of the group. Cawe, S. G., & McKenzie, B. (1989). The afromontane forests of
Furthermore, we argue for the limited removal of decaying Transkei, southern Africa. II: A floristic classification. South
wood logs in forested areas and the assistance of migration African Journal of Botany, 55, 31–39. https://doi.org/10.1016/
via suitable habitat condition. The results observed in the pre- S0254-6299(16)31230-3
sent study is a pattern that is also likely present in other low Cicconardi, F., Nardi, F., Emerson, B. C., Frati, F., & Fanciulli, P. P.
sedentary inverterbates with specific microclimatic regimes (2010). Deep phylogeographic divisions and long‐term persistence
of forest invertebrates (Hexapoda: Collembola) in the North‐Western
such as free‐living flatworms (Turbellaria), as well as in the
Mediterranean basin. Molecular Ecology, 19, 386–400. https://doi.
Annelida. Indeed, several fine‐scale studies of the latter ani-
org/10.1111/j.1365-294X.2009.04457.x
mal groups have revealed similar levels of high endemism, Clement, M., Posada, D., & Crandall, K. (2000). TCS : A computer
corroborating the importance of scale in sample design when program to estimate gene genealogies. Molecular Ecology, 9,
planning for conservation and taxonomic surveys. 1657–1659.
Clouse, R. M., Sharma, P. P., Stuart, J. C., Davis, L. R., Giribet, G.,
Boyer, S. L., & Wheeler, W. C. (2015). Phylogeography of the har-
ACKNOWLEDGEMENTS vestmen Metasiro (Arthropoda, Arachnida, Opiliones) reveals a po-
tential solution to the Pangean paradox. Organisms Diversity and
The Foundational Biodiversity Initiative Programme, through
Evolution, 16, 167–184.
the South African National Biodiversity Initiative under Curach, N., & Sunnucks, P. (1999). Molecular anatomy of an on-
the NRF, is thanked for funding the Forest Research grant. ychophoran: Compartmentalized sperm storage and heteroge-
Madelaine Frazenburg is thanked for assistance during and co- neous paternity. Molecular Ecology, 8, 1375–1385. https://doi.
ordination of SEM procedures. Theo Busschau is thanked for org/10.1046/j.1365-294x.1999.00698.x
BARNES and DANIELS   
   19
|
Daniels, S. R. (2011). Genetic variation in the critically endangered animal mitochondrial DNA. Annual Review of Ecology, Evolution
velvet worm Opisthopatus roseus (Onychophora : Peripatopsidae). and Systematics, 34, 397–423. https://doi.org/10.1146/annurev.
African Zoology, 46, 419–424. ecolsys.34.011802.132421
Daniels, S. R., Dambire, C., Klaus, S., & Sharma, P. P. (2016). Garrick, R. C., Sands, C. J., Rowell, D. M., Tait, N. N., Greenslade,
Unmasking alpha diversity, cladogenesis and biogeographical P., & Sunnucks, P. (2004). Phylogeography recapitulates topog-
patterning in an ancient panarthropod lineage (Onychophora: raphy: Very fine‐scale local endemism of a saproxylic ‘giant’
Peripatopsidae: Opisthopatus cinctipes) with the description of five springtail at Tallaganda in the great dividing range of South‐
novel species. Cladistics, 32, 506–537. east Australia. Molecular Ecology, 13, 3315–3330. https://doi.
Daniels, S. R., Dreyer, M., & Sharma, P. P. (2017). Contrasting the pop- org/10.1111/j.1365-294X.2004.02340.x
ulation genetic structure of two velvet worm taxa (Onychophora: Garrick, R. C., Sands, C. J., & Sunnucks, P. (2006). The use and ap-
Peripatopsidae: Peripatopsis) in forest fragments along the south‐ plication of phylogeography for invertebrate conservation research
eastern Cape, South Africa. Invertebrate Systematics, 31, 781–796. and planning. In: S. J. Grove, & J. L. Hanula (Eds.), Insect biodi-
https://doi.org/10.1071/IS16085 versity and dead wood: Proceedings of a symposium for the 22nd
Daniels, S. R., McDonald, D. E., & Picker, M. D. (2013). Evolutionary International Congress of Entomology (pp. 15–22). U.S. Department
insight into the Peripatopsis balfouri sensu lato species complex of Agriculture Forest Service Southern Research Station General
(Onychophora: Peripatopsidae) reveals novel lineages and zoogeo- Technical Report SRS–93, Asheville NC, USA.
graphic patterning. Zoologica Scripta, 42, 656–674. Gaublomme, E., Maebe, K., Van Doninck, K., Dhuyvetter, H., Li, X.,
Daniels, S. R., Picker, M. D., Cowlin, R. M., & Hamer, M. L. (2009). Desender, K., & Hendrickx, F. (2013). Loss of genetic diversity
Unravelling evolutionary lineages among South African velvet worms and increased genetic structuring in response to forest area reduc-
(Onychophora : Peripatopsis) provides evidence for widespread cryp- tion in a ground dwelling insect: A case study of the flightless ca-
tic speciation. Biological Journal of the Linnean Society, 97, 200–216. rabid beetle Carabus problematicus (Coleoptera, Carabidae). Insect
Daniels, S. R., & Ruhberg, H. (2010). Molecular and morphological Conservation and Diversity, 6, 473–482.
variation in a South African velvet worm Peripatopsis moseleyi Giribet, G., Carranza, S., Baguna, J., Riutort, M., & Ribera, C. (1996).
(Onychophora, Peripatopsidae): Evidence for cryptic speciation. First molecular evidence for the existence of a Tardigrada +
Journal of Zoology, 282, 171–179. Arthropoda clade. Molecular Biology and Evolution, 13, 76–84.
Drummond, A. J., Ho, S. Y., Phillips, M. J., & Rambaut, A. (2006). https://doi.org/10.1093/oxfordjournals.molbev.a025573
Relaxed phylogenetics and dating with confidence. PLoS Biology, Hamer, M. L., Samways, M. J., & Ruhberg, H. (1997). A review of the
4, e88. https://doi.org/10.1371/journal.pbio.0040088 Onycophora of South Africa, with discussion of their conservation.
Drummond, A. J., & Rambaut, A. (2007). BEAST: Bayesian evolu- Annals of the Natal Museum, 38, 283–312.
tionary analysis by sampling trees. BMC Evolution Biology, 7, 214. Harmon, L. J., Kolbe, J. J., Cheverud, J. M., & Losos, J. B. (2005).
https://doi.org/10.1186/1471-2148-7-214 Convergence and the multidimensional niche. Evolution, 59, 409–
Eeley, H. A., Lawes, M. J., & Piper, S. E. (1999). The influence of cli- 421. https://doi.org/10.1111/j.0014-3820.2005.tb00999.x
mate change on the distribution of indigenous forest in KwaZulu‐ Heled, J., & Drummond, A. J. (2010). Bayesian inference of species
Natal, South Africa. Journal of Biogeography, 26, 595–617. https:// trees from multilocus data. Molecular Biology and Evolution, 27,
doi.org/10.1046/j.1365-2699.1999.00307.x 570–580. https://doi.org/10.1093/molbev/msp274
Excoffier, L., & Lischer, H. (2010). Arlequin suite ver 3.5: A new se- Hughes, M., Möller, M., Bellstedt, D. U., Edwards, T. J., & De Villiers, M.
ries of programs to perform population genetics analyses under (2005). Refugia, dispersal and divergence in a forest archipelago: A
Linux and Windows. Molecular Ecology, 10, 564–567. https://doi. study of Streptocarpus in eastern South Africa. Molecular Ecology,
org/10.1111/j.1755-0998.2010.02847.x 14, 4415–4426. https://doi.org/10.1111/j.1365-294X.2005.02756.x
Farrell, B. D. (2001). Evolutionary assembly of the milkweed fauna: Jonsson, B. G., Kruys, N., & Ranius, T. (2005). Ecology of species
Cytochrome oxidase I and the age of Tetraopes beetles. Molecular living on dead wood – Lessons for dead wood management. Silva
Phylogenetics and Evolution, 18, 467–478. https://doi.org/10.1006/ Fennica, 39, 289–309. https://doi.org/10.14214/sf.390
mpev.2000.0888 Lawes, M. J. (1990). The distribution of the samango monkey (Cercopithecus
Felsenstein, J. (1985). Confidence limits on phylogenies, an ap- mitis erythrarchus Peters, 1852 and Cercopithecus mitis labiatus
proach using the bootstrap. Evolution, 39, 783–791. https://doi. I. Geoffroy, 1843) and forest history in southern Africa. Journal of
org/10.1111/j.1558-5646.1985.tb00420.x Biogeography, 17, 669–680. https://doi.org/10.2307/2845148
Fernandez, R., & Giribet, G. (2014). Phylogeography and species delimita- Lawes, M. J., Eeley, H. A., Findlay, N. J., & Forbes, D. (2007). Resilient
tion in the New Zealand endemic, genetically hypervariable harvestman forest faunal communities in South Africa: A legacy of palaeocli-
species Aoraki denticulata (Arachnida, Opiliones, Cyphophthalmi). matic change and extinction filtering? Journal of Biogeography, 34,
Invertebrate Systematics, 28, 401–414. https://doi.org/10.1071/IS14009 1246–1264.
Folmer, O., Black, M., Hoeh, W., Lutz, R., & Vrijenhoek, R. (1994). Majka, C. G., & Pollock, D. A. (2006). Understanding saproxylic bee-
DNA primers for amplification of mitochondrial cytochrome c tles: New records of Tetratomidae, Melandryidae, Synchroidae, and
oxidase subunit I from diverse metazoan invertebrates. Molecular Scraptiidae from the Maritime provinces of Canada (Coleoptera:
Ecology, 3, 294–299. Tenebrionoidea). Zootaxa, 1248, 45–68.
Fu, Y. (1997). Statistical tests of neutrality of mutations against popula- McDonald, D. E., & Daniels, S. R. (2012). Phylogeography of the Cape
tion growth, hitchhiking and background selection. Genetics Society velvet worm (Onychophora: Peripatopsis capensis) reveals the impact
of America, 147, 915–925. of Pliocene ⁄ Pleistocene climatic oscillations on Afromontane forest
Funk, D. J., & Omland, K. E. (2003). Species‐level paraphyly and poly- in the Western Cape, South Africa. Journal of Evolutionary Biology,
phyly: Frequency, causes, and consequences, with insights from 25, 824–835. https://doi.org/10.1111/j.1420-9101.2012.02482.x
|
20       BARNES and DANIELS

McDonald, D. E., Ruhberg, H., & Daniels, S. R. (2012). Two new Conservation Biology, 29, 382–390. https://doi.org/10.1111/
Peripatopsis species (Onychophora: Peripatopsidae) from the cobi.12427
Western Cape province, South Africa. Zootaxa, 3380, 55–68. Sunnucks, P., Curach, N. C., Young, A., French, J., Cameron, R.,
McKay, B. D., & Zink, R. M. (2010). The causes of mitochondrial Briscoe, D. A., & Tait, N. N. (2000). Reproductive biology of the on-
DNA gene tree paraphyly in birds. Molecular Phylogenetics ychophoran Euperipatoides rowelli. Journal of Zoology (London),
and Evolution, 54, 647–650. https://doi.org/10.1016/j.ympev. 250, 447–460. https://doi.org/10.1111/j.1469-7998.2000.
2009.08.024 tb00788.x
Miller, M. A., Pfeiffer, W., & Schwartz, T. (2010). Creating the Swofford, D. (2002). PAUP* Phylogenetic analysis using persimony
CIPRES Science Gateway for inference of large phylogenetic trees. (and other methods) Version 4.10. Champaigne, IL: Illinois Natural
Proceedings of the Gateway Computing Environments Workshop History Survey.
(GCE), 14 November 2010, New Orleans, LA, 1–8. Thompson, J. D., Gibson, T. J., Plewniak, F., Jeanmougin, F., & Higgins,
Myburgh, A. M., & Daniels, S. R. (2015). Exploring the impact of D. G. (1997). The Clustal X windows interface: Flexible strategies
habitat size on phylogeographic patterning in the Overberg velvet for multiple sequence alignment aided by quality analysis tools.
worm Peripatopsis overbergiensis (Onychophora: Peripatopsidae). Nucleic Acid Research, 24, 4876–4882.
Journal of Heredity, 106, 296–305. https://doi.org/10.1093/jhered/ Trewick, S. A. (1997). Sympatric cryptic species in New Zealand
esv014 Onychophora. Biological Journal of the Linnean Society, 63,
Oliveira, I. D. S., Read, V. M. S. J., & Mayer, G. (2012). A world check- 307–329.
list of Onychophora (velvet worms), with notes on nomenclature Trewick, S. A., & Wallis, G. P. (2001). Bridging the “beech‐gap”: New
and status of names. ZooKeys, 211, 1–70. https://doi.org/10.3897/ Zealand invertebrate phylogeography implicates Pleistocene glacia-
zookeys.211.3463 tion and Pliocene isolation. Evolution, 55, 2170–2180.
Posada, D., & Crandall, K. A. (1998). MODELTEST: Testing the model Tutt, K., Daughtery, C., & Gibbs, W. (2001). Differential life‐history
of DNA substitution. Bioinformatics, 14, 817–818. https://doi. characteristics of male and female Peripatoides novaezealandiae
org/10.1093/bioinformatics/14.9.817 (Onychophora: Peripatopsidae). Journal of Zoology (London), 258,
Quinn, G. P., & Keough, M. J. (2002). Experimental design and data 257–267. https://doi.org/10.1017/S095283690200136X
analysis for biologists. Cambridge, UK: Cambridge University Van Andel, T. H. (1989). Late Quaternary sea‐level changes and
Press. archaeology. Antiquity, 63, 733–745. https://doi.org/10.1017/
Rambaut, A.. (2009). FigTree. Retrieved from http://tree.bio.ed.ac.uk/ S0003598X00076869
software/figtree/ Weldon, C. W., Daniels, S. R., Clusella‐Trullas, S., & Chown, S. L.
Rambaut, A., Suchard, M., & Drummond, A. (2013). Tracer. Retrieved (2013). Metabolic and water loss rates of two cryptic species in the
fromhttp://tree.bio.ed.ac.uk/software/tracer African velvet worm genus Opisthopatus (Onychophora). Journal of
Ranius, T. (2006). Measuring the dispersal of saproxylic insects: A key Comparitive Physiology B, 183, 323–332. https://doi.org/10.1007/
characteristic for their conservation. Population Ecology, 48, 177– s00360-012-0715-2
188. https://doi.org/10.1007/s10144-006-0262-3 Williams, S., & Pearson, R. (1997). Historical rainforest contractions,
Reinhard, J., & Rowell, D. M. (2005). Social behaviour in an Australian localized extinctions and patterns of vertebrate endemism in the
velvet worm, Euperipatoides rowelli (Onychophora: Perpatopsidae). rainforests of Australia’s wet tropics. The Royal Society, 264, 709–
Journal of Zoology (London), 267, 1–7. 716. https://doi.org/10.1098/rspb.1997.0101
Ronquist, F., Teslenko, M., Van Der Mark, P., Ayres, D. L., Darling, Woodman, J. D., Cooper, P. D., & Haritos, V. S. (2007). Effects
A., Höhna, S., & Huelsenbeck, J. P. (2012). MrBayes 3.2: Efficient of temperature and oxygen availability on water loss and car-
Bayesian phylogenetic inference and model choice across a bon dioxide release in two sympatric saproxylic invertebrates.
large model space. Systematic Biology, 61, 539–542. https://doi. Comparative Biochemistry and Physiology Part A. Molecular &
org/10.1093/sysbio/sys029 Integrative Physiology, 147, 514–520. https://doi.org/10.1016/j.
Ruhberg, H., & Daniels, S. R. (2013). Morphological assessment cbpa.2007.01.024
supports the recognition of four novel species in a widely‐distrib- Zwickl, D. J. (2006). Genetic algorithm approaches for the phylogenetic
uted velvet worm Peripatopsis moseleyi sense lato (Onychophora: analysis of large biological sequence datasets under the maximum
Peripatopsidae). Invertebrate Systematics, 27, 131–145. likelihood criterion. PhD Dissertation, The University of Texas at
Rutherford, M. C., Mucina, L., & Powrie, L. W. (2006). Biomes and Austin.
bioregions of southern Africa. The vegetation of South Africa,
Lesotho and Swaziland. Pretoria: South African Biodiversity
Institute. How to cite this article: Barnes A, Daniels SR. On
Sands, C. J., Lancaster, M. L., Austin, J. J., & Sunnucks, P. (2009). Single the importance of fine‐scale sampling in detecting
copy nuclear DNA markers for the onychophoran Phallocephale tal- alpha taxonomic diversity among saproxylic
lagandensis. Conservation Genetic Resources, 1, 17–19. https://doi.
invertebrates: A velvet worm (Onychophora:
org/10.1007/s12686-009-9004-0
Seibold, S., Brandle, R., Buse, J., Hothorn, T., Schmidel, J., Thorn,
Opisthopatus amaxhosa) template. Zool Scr.
S., & Müller, J. (2015). Association of extinction risk of sap- 2019;00:1–20. https://doi.org/10.1111/zsc.12338
roxylic beetles with ecological degradation of forests in Europe.

Potrebbero piacerti anche