You are on page 1of 75

DR.

CENE FIŠER (Orcid ID : 0000-0003-1982-8724)

Article type : Invited Reviews and Syntheses


Accepted Article
Cryptic species as a window into the paradigm shift of the species concept

Cene Fišer1, Christopher T. Robinson2,3, Florian Malard4


1
SubBio Lab, Dept. of Biology, Biotechnical Faculty, University of Ljubljana; Jamnikarjeva 101,

Ljubljana, Slovenia.
2
Eawag, Dept. Aquatic Ecology, Überlandstrasse 133, 8600 Dübendorf, Switzerland
3
Institute of Integrative Biology, ETH Zürich, 8092 Zürich, Switzerland.
4
UMR5023 Ecologie des Hydrosystèmes Naturels et Anthropisés, Université Lyon 1, ENTPE, CNRS,

Université de Lyon, 6 Rue Raphaêl Dubois, 69622, Villeurbanne, France.

Key words: Amphipoda, biodiversity patterns, convergent evolution, integrative taxonomy,

phylogenetic niche conservatism, speciation

Author for correspondence:

Cene Fišer, cene.fiser@bf.uni-lj.si ; SubBio Lab, Dept. of Biology, Biotechnical Faculty, University of

Ljubljana; Jamnikarjeva 101, Ljubljana, Slovenia; phone: +386 1 3203 371 , fax: +386 1 2573 390

Running head: Cryptic species in biodiversity science

Abstract

The species concept is the cornerstone of biodiversity science and any paradigm shift in the

delimitation of species affects many research fields. Many biologists now are embracing a new

‘species’ paradigm as separately evolving populations using different delimitation criteria. Individual
This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process, which may
lead to differences between this version and the Version of Record. Please cite this article as
doi: 10.1111/mec.14486
This article is protected by copyright. All rights reserved.
criteria can emerge during different periods of speciation; some may never evolve. As such, a

paradigm shift in the species concept relates to this inherent heterogeneity in the speciation process
Accepted Article
and species category – which is fundamentally overlooked in biodiversity research. Cryptic species

fall within this paradigm shift: they are continuously being reported from diverse animal phyla but

are poorly considered in current tests of ecological and evolutionary theory. The aim of this review is

to integrate cryptic species in biodiversity science. In the first section, we address that the absence

of morphological diversification is an evolutionary phenomenon, a ‘process’ counterpart to the long-

studied mechanisms of morphological diversification. In the next section regarding taxonomy, we

show that molecular delimitation of cryptic species is heavily biased towards distance-based

methods. We also stress the importance of formally naming of cryptic species for better integration

into research fields that use species as units of analysis. Finally, we show that incorporating cryptic

species leads to novel insights regarding biodiversity patterns and processes, including large-scale

biodiversity assessments, geographic variation in species distribution, and species coexistence. It is

time for incorporating multi-criteria species approaches aiming to understand speciation across

space and taxa, thus allowing integration into biodiversity conservation while accommodating for

species uncertainty.

1. INTRODUCTION

Genes, species and ecosystems are three major components of biodiversity. Among them, species

has proven the most challenging to conceptualize. Decade-lasting disputes over the species concept

yielded no less than 24 species definitions (Mayden, 1997). Defining a species is central to

biodiversity science (de Queiroz, 1998, 2005b; Mayden, 1997) because any change in the species

concept directly affects many important research fields, including taxonomy, biogeography, ecology,

conservation and macroevolution (Agapow et al., 2004).

An important step forward was made with a distinction between conceptual and operational

issues, i.e. what a species is and how they should be delimited (Bock, 2004; Hey, Waples, Arnold,

This article is protected by copyright. All rights reserved.


Butlin, & Harrison, 2003). Most authors today agree conceptually (species concept, see Glossary), i.e.

that species are entities emerging from the speciation process as separately evolving segments of
Accepted Article
metapopulation lineages (Bock, 2004; de Queiroz, 1998; Hey, 2006; Hey et al., 2003; Mayden, 1997).

Establishing the correspondence between the species concept and species taxa, i.e. elementary units

used in biodiversity research (see Glossary), remains a challenge. From a temporal point of view, a

species is framed by a speciation event, at the origin of this process, and by event putting an end to

this species (further speciation or extinction) (Bock, 2004; Roux et al., 2016). Inference of speciation

depends on different properties that are expressed during separate evolution, such as morphological

distinguishability, reproductive isolation and monophyly. These properties remain the best lines of

evidence to identify lineage divergence and thus delimit species taxa (de Queiroz, 2005a,b, 2007).

However, the evolution of these properties is context dependent and can emerge at different times

and sequential order, if at all (de Queiroz, 2005a, 2007). Consequently, there is no particular

property that would universally delimit a species. Not surprisingly, difficulties inherent to species

delimitation prompted proliferation of species delimitation methods based on different premises

(Fontaneto, Flot, & Tang, 2015; Raxworthy, Ingram, Rabibisoa, & Pearson, 2007; Sites & Marshall,

2003, 2004), whose accuracy depended on the context of speciation (Dellicour & Flot, 2015).

Furthermore, species taxa delimited with these methods should be considered merely as scientific

hypotheses of separately evolving entities (de Queiroz, 2007; Hey et al., 2003), which are iteratively

tested using new methods and data in the procedure called “integrative taxonomy” (Padial, Miralles,

De la Riva, & Vences, 2010; Yeates et al., 2011). The outcome of this conceptual and methodological

development is recognition that species taxa represent a heterogeneous set of hypotheses whose

properties are contingent with the heterogeneous nature of speciation.

A central question raised by this paradigm shift is whether we can make use of the explicit

links between the speciation process and species hypotheses. If heterogeneity in the speciation

process leads to heterogeneity among species hypotheses, can we integrate heterogeneity in

speciation into evolutionary, biogeographic and ecological research? This issue has been partly

This article is protected by copyright. All rights reserved.


explored in the case of ecological speciation (Schluter, 2000a, 2016; Stroud & Losos, 2016).

Ecological speciation is a process where ancestral populations diverge due to resource partitioning
Accepted Article
and is associated with phenotypic differentiation (Nosil, 2012; Rundle & Nosil, 2005). It takes place

when ecologically accessible resources can be evolutionarily exploited (ecological opportunity)

(Schluter, 2000b; Stroud & Losos, 2016). Ecological speciation can contribute to speciation rates and

can underlie major evolutionary events such as adaptive radiation, but its relative importance

depends on ecological opportunities that vary in space and time (Schluter, 2016). Hence, ecological

speciation provides a mechanistic explanation between ecological opportunity and global

biodiversity patterns (Kozak & Wiens, 2016; Schluter, 2016). Further generalizations require better

understanding between environmental and speciation relationships, especially regarding speciation

heterogeneity.

Over the last two decades, increased evidence emerged for speciation governed by entirely

different mechanisms, leading to so-called sibling or cryptic species. The idea that species can evolve

similar morphologies is historical (Mayr, 1942), but the use of molecular delimitation methods has

now brought cryptic species to the forefront in many research arenas (Bickford et al., 2007;

Knowlton, 1993; Nygren, 2014). Indeed, we are now aware that cryptic species are quite common

(M. Adams, Raadik, Burridge, & Georges, 2014) and widespread across most animal phyla (Pérez-

Ponce de León & Poulin, 2016; Pfenninger & Schwenk, 2007). They even have been found in mostly

overlooked parasite (Pérez-Ponce de León & Nadler, 2010) and nematode species (Derycke et al.,

2008), but also among the largest and heavily studied mammalian species like giraffes (Fennessy et

al., 2016). In spite of being common, they are still poorly considered in testing ecological and

evolutionary theories (Beheregaray & Caccone, 2007). It is timely to intensify research beyond the

phenomenological level of exploration and incorporate what we can learn from cryptic species into

biodiversity science. The aim of this review is to propose a research agenda for fully integrating

cryptic species into biodiversity science, thereby fostering a better understanding of the

heterogeneous role of speciation in biodiversity pattern and process.

This article is protected by copyright. All rights reserved.


Towards this goal, we examine three inter-related facets of cryptic species in respect to

biodiversity science (Fig. 1). First, we reconsider and define the phenomenon of phenotypic
Accepted Article
similarity from an evolutionary perspective. We explore three non-mutually exclusive mechanisms

explaining the lack of morphological distinguishability among divergent lineages; i.e., recent

divergence, niche conservatism, and morphological convergence. Here, we evaluate the existing

empirical evidence and suggest novel tests to better elucidate each mechanism. Second, we review

the practices in delimiting cryptic species and reconsider them from a taxonomic perspective. We

explore other contingent properties of species currently used to identify species boundaries in the

absence of morphological distinguishability. Here, we look into the burgeoning fields of DNA

taxonomy and integrative taxonomy (Camargo & Sites, 2013; Fontaneto et al., 2015; Padial et al.,

2010), highlighting the recurrent need for naming species. We show how integrative taxonomy can

improve understanding of morphological cryptic evolution. Third, we discuss how cryptic species

have been integrated into biodiversity research and, more specifically, what can be gained by

incorporating different hypotheses (or properties) of speciation into biodiversity science. Lastly, we

use these facets to identify major gaps in biodiversity research and promising research directions

towards enhancing biodiversity science.

The ideas discussed in this review apply to all animal phyla. To provide data-driven

evaluation regarding the current research status on cryptic diversity, we focus (but not limit) our

literature search to the Amphipoda. Amphipods are an excellent invertebrate model for studying

morphologically cryptic species, underlying speciation and its relationship to biodiversity science.

They are species rich, ecologically diverse, thrive in most aquatic environments, play a crucial role in

ecosystem function, and are often used as biological indicators. Importantly, they comprise

numerous cryptic species within different ecological types, and thus are relatively well-studied in

eco-evolutionary studies (Box 1, Fig. 2).

This article is protected by copyright. All rights reserved.


2. COMPILATION OF LITERATURE DATA

Cryptic species were often detected as a by-product in phylogeographic studies of amphipods. For
Accepted Article
this reason, we searched for relevant articles in the ISI Web of Science using keywords

(Amphipoda/Amphipods * cryptic species), (Amphipoda/Amphipods * sibling species) and

(Amphipoda/Amphipods * phylogeography) and (Amphipoda/Amphipods * molecular taxonomy).

The search before the year 2000 yielded no results, hence we analyzed the period between 2000

and 2015 (search was made on September 2015). All titles were checked and duplicates removed,

yielding a total of 120 articles. From this search, our review covers 122 morphological species of

amphipods or morphospecies, i.e., species that were delimited on a basis of their morphological

distinctness. The list of morphospecies contained no subspecies. These morphospecies collectively

contained a total of 439 cryptic species. Of 122 nominal species, 13 were found to be synonyms, but

the remaining 109 morphospecies represent 304 (295-315, depending on the delimitation method of

the 439 total) cryptic species discovered for the first time. Individual articles were searched for

several categories of information that were relevant to the three topics below (see also Supporting

Information).

2.1. Mechanisms causing cryptic diversity

From these 120 articles, we searched for information regarding three related hypotheses explaining

the lack of morphological distinguishability, namely recent divergence, phylogenetic niche

conservatism and morphological convergence. The recent divergence hypothesis posits that cryptic

species have diverged recently and morphological differentiation presently is not evident (Egea et

al., 2016). Here, we analyzed the distribution of age of the cryptic species using reported estimations

of node age in phylogenetic trees. Descriptive estimations of age were discarded (e.g., early

Miocene), which reduced the number of relevant papers to 29.

This article is protected by copyright. All rights reserved.


The phylogenetic niche conservatism hypothesis (PNC) postulates that niche evolution and hence

morphological differentiation across descendant species is constrained by selection, a phenomenon


Accepted Article
known as morphological stasis (Bravo, Remsen, & Brumfield, 2014; Egea et al., 2016; Smith, Harmon,

Shoo, & Melville, 2011). To identify cases of cryptic species potentially matching this hypothesis, we

searched for reports of cryptic species among complexes of sister species that were ecologically

specialized at the micro-niche level. We assumed that a single origin of narrow specialization in an

ancestor parsimoniously explains specialization in sister descendants more than multiple and

independent origins.

The morphological convergence hypothesis also is based on selection and assumes that

morphological similarity may evolve independently among distantly-related species in response to

similar selection pressures (Bravo et al., 2014; Trontelj et al., 2009). Here, we explored whether

morphological similarity of cryptic species evolved independently through convergent evolution. For

this reason, we quantified the proportion of cryptic species that were not in sister relationships with

each other.

2.2 Taxonomic practices in delimiting cryptic species

Authors rarely a priori referred to a well-defined species concept while delimiting species taxa. For

this reason, the evaluation what they considered as a species was not possible. The following

analysis of past research is based on the premise that species are separately evolving entities, and

that species taxa can be delimited using the criteria of genetic isolation.

We reviewed the number and type of molecular markers as well as the methods used to delimit

cryptic species. Molecular markers included allozymes, mitochondrial and nuclear genes,

microsatellites and single nucleotide polymorphisms (SNPs). DNA taxonomy is a rapidly evolving field

such that molecular methods for species delimitation used a decade ago may not be considered

This article is protected by copyright. All rights reserved.


valid today (compare, for example, the classifications of Sites & Marshall, 2004, 2003 and Flot, 2015;

Fontaneto et al., 2015). All methods, albeit on different premises, search for cessation of gene flow
Accepted Article
between populations. Such genetic isolation is a primary pointer of ongoing speciation (Bock, 2004)

and emergence of independently evolving entities (de Queiroz, 2007), and serves for delimitation of

corresponding species taxa (see Glossary). The progress in methodology makes any classification

using different methods a difficult exercise. For this review, we used four categories for

classification: namely, distance-based, tree-based, allele-sharing and population genetics methods

(Table 1). Distance-based methods assume that speciation should result in smaller genetic distances

within than between species. Species are delimited either using a fixed threshold (Lefébure, Douady,

Gouy, & Gibert, 2006) or the best-fitting threshold calculated for the dataset at hand (Puillandre,

Lambert, Brouillet, & Achaz, 2012). Tree-based methods search for the most likely transition point

between species-level and coalescent processes using single locus (Monaghan et al., 2009; Pons et

al., 2006; Reid & Carstens, 2012; J. Zhang, Kapli, Pavlidis, & Stamatakis, 2013), multiple loci (Ence &

Carstens, 2011; Grummer, Bryson, & Reeder, 2014; Knowles & Carstens, 2007; O’Meara, 2010; Z.

Yang & Rannala, 2010, 2014) and genomic-scale data (Leache, Fujita, Minin, & Bouckaert, 2014).

Allele-sharing based methods (Flot et al., 2010) attempt to discern populations that are no longer

connected by gene flow and therefore evolved mutual allelic exclusivity. We recognize that

population genetic methods (e.g. nested clade analysis, indices of genetic differentiation) are not

intended to solely delimit species, but have been used often in the past in respect to species

delimitation. In fact, many authors considered genetically distinct populations as putative species.

In addition, we quantified cases when cryptic species were analyzed for differences in morphology,

physiology, ecology or behavior. Compositing different data types to delimit species is known as

integrative taxonomy (Padial et al., 2010; Wiens & Penkrot, 2002). It can either provide additional

support for a species hypothesis (Schlick-Steiner et al., 2010) or detect contradictions among data

(Padial et al., 2010); both help link taxonomy with evolutionary biology (Sites & Marshall, 2003,

2004). Finally, we examined whether cryptic species have a binomial name and, if so, it is based

This article is protected by copyright. All rights reserved.


solely on molecular criteria (Cook, Edwards, Crisp, & Hardy, 2010; Jörger & Schrödl, 2013; Renner,

2016). Although naming is not always done, the naming of new species integrates taxonomy with
Accepted Article
other fields of biology, (Pante, Schoelinck, & Puillandre, 2015).

2.3 Cryptic species in biodiversity research

Although cryptic species are being documented at an ever-increasing rate, they are integrated in

quite different ways across fields in biodiversity science. Here, we distinguish between two major

ways of incorporating cryptic species in biodiversity research. The first corresponded to studies that

explore whether cryptic species might have biased our understanding of biodiversity patterns and

thus misled conservation efforts. The second corresponded to studies that only retain DNA-based

criteria for delimiting species, thereby inherently using cryptic species as units of analysis.

3. MECHANISMS CAUSING CRYPTIC DIVERSITY

3.1 Recent divergence

The simplest evolutionary model predicts that morphological disparity increases due to stochastic

processes alone (see e.g., Adams et al. (2009) and Harmon et al. (2003) for changes in morphology

through time). The evolution of morphological distinctness depends only on time since divergence,

hence cryptic species are expected to be relatively young (Zúñiga-Reinoso & Benítez, 2015).

Diagnostic morphological traits evolve slowly if they are selectively neutral, effective population size

is large and ancestral polymorphism is high (Egea et al., 2016; Wiens & Servedio, 2000).

We obtained age estimates for 43 cryptic species of amphipods (Fig. 3). The age distribution was

typically skewed left, but only five out of the 43 cryptic species were younger than 1 Myr. This result

is in stark contrast to theoretical predictions mentioned above. One explanation may be that this

This article is protected by copyright. All rights reserved.


mechanistic explanation is unlikely. The evolutionary response to environmental selection is rapid

(Carroll, Hendry, Reznick, & Fox, 2007; Herrel et al., 2008), thus evolving species quickly attain their
Accepted Article
morphological distinctness. The alternative explanation invokes methodological biases. First, the

methods used for age estimation have (and are still ongoing) changed over time, and current

estimated ages are sensitive to a number of poorly understood parameters (dos Reis, Donoghue, &

Yang, 2015). Different authors estimated node ages using different methods, optimized for their

datasets. Some variation in age estimation among these studies is unavoidable and at times even

biased. Second, many evolutionarily young cryptic species might have been simply overlooked. Most

of the reviewed studies used coarse delimitation methods that accurately delineate only relatively

old species (see Section 3). Consequently, we conclude that recent divergence presently explains

morphological similarity of only a small fraction of known cryptic species.

To note, none of the examined studies explicitly tested for recent divergence. This hypothesis has

two central predictions: first, cryptic species are young and, second, morphological traits evolve

under the neutral model of evolution (Brownian motion, see Glossary) (Felsenstein 1985). Rigorous

testing of the predictions should consider three aspects as guidelines for future research. First,

testing requires accurate age estimations. This task is not easy and methodological details are

beyond the scope of this review; readers should consult more specialized papers (dos Reis et al.,

2015). Second, the distinction between “young” and “old” species depends on ancestral

polymorphism and population size, therefore it is clade specific. Appropriate evaluation of the

hypothesis needs to account for clade specificity. For instance, assessing whether cryptic species are

younger than non-cryptic ones requires analysis at the clade level rather than the species pair level.

Lastly, models of evolution can be estimated from extant morphology and phylogeny, but a changing

selection regime also can produce a pattern similar to neutral evolution and thus lead to erroneous

estimation of trait evolution. Fortunately, this issue can be solved using recently developed software

that allows for discrimination between models of neutral evolution and changing selective regimes

(Clavel, Escarguel, & Merceron, 2015).

This article is protected by copyright. All rights reserved.


3.2 Phylogenetic niche conservatism (PNC)

Recent divergence cannot explain relatively old cryptic species. An alternative hypothesis, so called
Accepted Article
“morphological stasis”, assumes that morphological differentiation of descendant species is

constrained by biological mechanisms (Bickford et al., 2007). Since a species’ morphology is partly

related to its ecological niche (e.g., Losos, 2009, Schluter, 2000a), this hypothesis is closely related to

the broader concept of phylogenetic niche conservatism (PNC). Initially, PNC was used to explain the

tendency for some lineages to retain their ecological niches over evolutionary time. More recently,

the concept was elaborated further (Pyron, Costa, Patten, & Burbrink, 2015), but here we refer to its

original intention. The biological mechanisms underlying PNC encompass stabilizing selection, gene

flow swamping local adaptation, pleiotropic effects constraining adaptation, and lack of genetic

variation (Losos, 2008; Wiens, 2004, 2008; Wiens et al., 2010; Wiens & Graham, 2005).

The PNC hypothesis was not explicitly tested in the reviewed studies here. Nevertheless, we found

several cases of cryptic species among complexes of sister amphipod species that were ecologically

highly specialized, and might support the PNC hypothesis. These include Cyamus species (whale lice)

living on distinct parts of a whale’s body (Kaliszewska et al., 2005), Niphargus species colonizing

fissures in rock (C. Fišer & Zagmajster, 2009; Meleg, Zakšek, Fišer, Kelemen, & Moldovan, 2013;

Švara, Delić, Rađa, & Fišer, 2015; Trontelj et al., 2009), some species of Gammarus inhabiting

mountain streams (Hou & Li, 2010), and sponge-dwelling species of Leucothoe (Richards, Stanhope,

& Shivji, 2012; White, Reimer, & Lorion, 2016). This list is likely longer, especially among deep sea

and subterranean amphipods. Regardless, attributing morphological crypsis to PNC for species with

unclear ecological specialization is problematic in the absence of explicit testing.

We now discuss three predictions of the PNC to stimulate future research regarding mechanisms

causing cryptic diversity (Pyron et al., 2015; Wiens et al., 2010). First, evolution of morphology is

constrained. Morphological disparity remains constant through time, as if being constrained by

stabilizing selection. The model of morphological evolution inferred from extant morphology and

This article is protected by copyright. All rights reserved.


phylogeny no longer corresponds to neutral Brownian motion but to models simulating non-neutral

evolution (but see Losos, 2008). These models include stabilizing selection (Orstein-Uhlenbeck
Accepted Article
model, see Glossary) or, arguably, constantly moving adaptive optima (white noise model, see

Glossary) (Pyron et al., 2015; Smith et al., 2011; Wiens et al., 2010). The preferred models are

selected based on likelihood values. Some authors proposed an alternative test using a special

model of Brownian motion that allows for different rates of trait evolution across clades. The rates

of trait evolution should be lower in niche-conserved than in niche non-conserved clades (O’Meara,

Ané, Sanderson, & Wainwright, 2006; Wiens et al., 2010). Second, sister cryptic species are expected

to share ecological niches. Equivalency among ecological niches can be tested either implicitly from

distributional data (Kozak & Wiens, 2006; Raxworthy et al., 2007; Warren, Glor, & Turelli, 2008) or

from experimental data (Kearney, Wintle, & Porter, 2010). Third, PNC promotes allopatric speciation

(Kozak & Wiens, 2006; Pyron et al., 2015; Wiens, 2004; Wiens & Graham, 2005). This prediction

requires analysis of the evolution of geographic ranges (Matzke, 2013; Ree & Smith, 2008). To note,

the third prediction may strengthen arguments supporting PNC, but it is not mandatory. Species can

evolve in sympatry along niche axes, e.g. salinity, pH or temperature (Delić, Švara, Coleman, Trontelj,

& Fišer, 2017) that are not related to morphology.

3.3 Morphological convergence

The third hypothesis infers that morphological similarity evolved among distantly-related species in

response to similar selective regimes (Ingram & Mahler, 2013). Morphological convergence based on

our reviewed data was common: 26% of the nominal amphipods examined contained cryptic species

that were not sister groups. This proportion is probably a low estimate because detection of

convergence depends on completeness of taxon sampling. Cases of morphological convergence

among cryptic species were documented in several amphipod genera, where species from the same

cryptic complex independently colonized freshwater-semiterrestrial (L. Yang, Hou, & Li, 2013) or

This article is protected by copyright. All rights reserved.


subterranean habitats (Villacorta, Jaume, Oromí, & Juan, 2008) or microhabitats within the

subterranean realm (Trontelj et al., 2009; Trontelj, Blejec, & Fišer, 2012).
Accepted Article
Evidence from the above studies is descriptive. A central prediction of this hypothesis states that

convergent evolution should result from shifts in selective regimes. Appropriate testing must show a

correlated evolution between morphology and the environment. The evolution of morphology

should correspond to the Ornstein-Uhlenbeck stabilizing selection model (see Glossary) (Butler &

King, 2004; Hansen, 1997). Further, Ingram & Mahler (2013) developed a method to map shifts in

continuous phenotypic characters towards convergent regimes on a phylogenetic tree and to test

whether these shifts are more numerous than expected under a given null model.

3.4 General summary of invisible species diversification

Patterns of cryptic species, i.e. the distribution of cryptic species across space and taxa as well as the

occurrence of closely-related / distantly-related cryptic species have been relatively well

documented in the amphipod literature and in the literature in general. Yet, there is comparatively

much less knowledge about the relative importance of different mechanisms generating cryptic

species. Since, as usual in science, patterns alone are not enough to infer mechanism (i.e. different

mechanisms can generate the same pattern), there is a need for a shift from pattern-oriented

studies of cryptic diversity to mechanism-oriented studies of cryptic diversity. The mechanisms

generating cryptic species are clearly diverse. They likely follow the same evolutionary and ecological

principles that generated the diversity of morphologically distinguishable species and have

fascinated naturalists for centuries (Collar, Schulte, & Losos, 2011; Harmon et al., 2010; Losos, 2011;

Nosil, 2012). The lack of morphological diversification is an evolutionary phenomenon, a counterpart

to the long-studied mechanisms of morphological diversification such as character displacement

(Schluter, 2000a; Stuart & Losos, 2013). Research on morphological similarity is hence closely related

This article is protected by copyright. All rights reserved.


to that on morphological diversification. It requires gathering information on the function and

heritability of traits as well as on how trait variation relates to fitness (Schluter, 2000a; Stuart &
Accepted Article
Losos, 2013). Contrary to morphological diversification, mechanisms causing morphological similarity

have received little attention. Most of the reviewed studies focused on revealing cryptic species but

did not test for causal mechanisms. Thus even if slightly morphologically different, any species pair

satisfying the predictions above might be treated as morphologically cryptic.

Many authors stated that perfect morphological similarity does not exist and that morphological

analyses based on molecular species delimitation often yield morphological distinctness. Hence,

some researchers consider the concept of cryptic species as overrated (Jugovic, Jalžić, Prevorčnik, &

Sket, 2012; Karanovic, Djurakic, & Eberhard, 2016; Knowlton, 1993; Zúñiga-Reinoso & Benítez, 2015).

Noteworthy, some authors have already coined a term “pseudo-cryptic” in order to distinguish

“completely identical species” from “morphologically diagnosable species” (Achurra, Rodriguez, &

Erséus, 2015; Knowlton, 1993; Sáez & Lozano, 2005). This classification may lead to questions such

as “When are species similar enough to be considered morphologically cryptic?” (Karanovic et al.,

2016; Lajus, Sukhikh, & Alekseev, 2015). This question is at risk to remain without epilogue.

Importantly, it could hinder exploration of a rather broadly widespread evolutionary phenomenon

and even impede progress in biodiversity science. As discussed further below, the quest for

mechanisms can lead taxonomic practices towards delivering species hypothesis that are

theoretically grounded in the heterogeneous properties of speciation (section 3), and can better

integrate the speciation process into ecology and evolution (section 5).

4. TAXONOMIC PRACTICES IN DELIMITING CRYPTIC SPECIES

4.1 Species delimitation methods

Up to now, most cryptic species of amphipods were delimited using either one or two genes (Fig.

4A). The mitochondrial cytochrome oxidase subunit I (COI) gene was used in 83% of the reviewed

This article is protected by copyright. All rights reserved.


studies, possibly reflecting the general use and applicability of this marker across most animal phyla

(Hebert, Cywinska, Ball, & DeWaard, 2003; Hebert, Ratnasingham, & Waard, 2003) and as a well-
Accepted Article
developed “barcoding” system for storage and identification of animal species (Ratnasingham &

Hebert, 2007). Other genetic markers included allozymes, additional mitochondrial markers (16S and

CytB) and nuclear markers (28S rDNA, ITS, H3, EF1α, ND II and 18S rDNA). Guerra-García et al. (2006)

used RAPDs to explore intra- and interspecific genetic differentiation in two species of caprellidean

amphipods. Microsatellites were developed for numerous amphipods (Baird, Miller, & Stark, 2012;

Rewicz, Wattier, Rigaud, Bacela-Spychalska, & Grabowski, 2015; Villacorta, Canovas, Oromi, & Juan,

2009; Weiss & Leese, 2016), but were never used for taxonomic purposes and are not considered

here.

Molecular delimitation of cryptic amphipod species most often relied on a single category of

methods; only 30% of the examined studies combined two or three methods (Fig. 4B). In most

studies, species were delimited using distance-based methods or using a combination of distance-

based and population genetics methods (Fig. 4B). There was a substantial lag in the application of

tree-based methods. The generalized mixed Yule-coalescent (Pons et al., 2006), multilocus

coalescent (Rannala & Yang, 2003; Z. Yang & Rannala, 2010) and Poisson tree process models (J.

Zhang et al., 2013) were used in only seven studies (Copilaş-Ciocianu & Petrusek, 2015; Esmaeili-

Rineh, Sari, Delić, Moškrič, & Fišer, 2015; Katouzian et al., 2016; King & Leys, 2011; Murphy, King, &

Delean, 2015; Weiss, Niklas, Anna, & Florian, 2014; L. Yang et al., 2013). This lag in the use of tree-

based methods occurred despite 78% of the studies inferring molecular phylogenies (supporting

information 1). Allele-sharing based methods, e.g., haplowebs (Flot, Couloux, & Tillier, 2010), were

used in only two studies (Brad, Fišer, Flot, & Sarbu, 2015; Flot et al., 2014).

Use of different molecular markers and species delineation methods revealed conflicting results

about the number of species. Alternative species delimitations suggested by different methods were

not in conflict with each other, but differed in the degree of splitting; distance methods typically

This article is protected by copyright. All rights reserved.


turned to be more conservative (Copilaş-Ciocianu & Petrusek, 2015; Delić, Švara, et al., 2017; Delić,

Trontelj, Rendoš, & Fišer, 2017; Eme et al., 2017; Katouzian et al., 2016; Weiss et al., 2014). In such
Accepted Article
cases, authors either provided a rough estimate of species number (Copilaş-Ciocianu & Petrusek,

2015; Weiss et al., 2014), opted for a particular species hypothesis by ranking the efficiency of

different methods (e.g., conservative ABGD was preferred over PTP and bGMYC; Katouzian et al.,

2016), reconsidered the results within spatial context (presence of syntopy (Delić, Trontelj, et al.,

2017)), or drew on other studies to justify the choice of a particular threshold value. Unfortunately,

cross-referencing is not always appropriate because the value of thresholds depend on how genetic

distances are calculated (e.g., uncorrected p, K2P, patristic distances) (Richards et al., 2012).

Moreover, the use of thresholds is case-specific because the rate of substitution differs among taxa

(but see Lefébure et al., 2006). Oddly, reasons for the discordant number of species number were

never explicitly addressed. Inconsistency among different species delimitation methods is well

known, arising due to violations of various assumptions required by the different delimitation

methods (Carstens, Pelletier, Reid, & Satler, 2013). Molecular species delimitation methods generally

yield congruent results when speciation rates are low, population sizes are small, and the molecular

markers used have a high mutation rate (Dellicour & Flot, 2015; Fontaneto et al., 2015). On the

contrary, evolutionarily young species with large effective population sizes can be a source of

dispute (as alluded to above). A highly promising solution for the issue seems to be genome-wide

multilocus coalescence (Fujita, Leaché, Burbrink, McGuire, & Moritz, 2012; Leache et al., 2014), and

also confrontation with non-molecular characters as discussed in the next section.

4.2 Confronting species criteria

Ideally, species should be delimited using an integrative approach including genetics, morphology,

ecology, behavior and geography (Edwards & Knowles, 2014; Guillot, Renaud, Ledevin, Michaux, &

Claude, 2012; Padial et al., 2010). Steps towards integrative taxonomy were made in 55 of the 122

This article is protected by copyright. All rights reserved.


(42%) nominal species examined. To note, integrative exploration of cryptic species complexes often

occurred in a series of consecutive publications by a research group. Molecular inferences were


Accepted Article
mostly supplemented with morphological and ecological data as well (Fig. 4C). Reproductive barriers

were explicitly inferred from mating behavior in five cryptic species complexes. Differences in

genome size were reported for a single species complex. No biological difference between cryptic

species, except genetic ones, was detected in only four of the 55 nominal species (Cothran,

Henderson, Schmidenberg, & Relyea, 2013; Meleg et al., 2013; Wellborn & Cothran, 2004).

4.3 Naming cryptic species

In most cases, cryptic species were un-named, even when supported by multiple lines of evidence.

The 304 (295-315 depending on the method, excluding nominal species) species of amphipods

discovered in the period 2000-2015 represented 17% of the 1822 amphipod species described

during the same time period (Fig. 5). Yet, only 21 of these cryptic species were described and named.

In other words, DNA-complemented taxonomy contributed only 1% to the total number of described

amphipod species. Only two studies relied solely upon DNA sequences to diagnose species (Delić,

Trontelj, et al., 2017; Murphy et al., 2015) and one study complemented morphological diagnoses

with diagnostic COI sequences (King & Leys, 2011).

4.4 Future challenges for the taxonomy of cryptic species

Taxonomic practices can contribute more to the mechanistic understanding of morphological crypsis

as well as to the integration of cryptic species into biodiversity research. Aside from technical issues

such as development of barcoding libraries (Morinière et al., 2017), we envision three important

issues.

This article is protected by copyright. All rights reserved.


First, testing whether cryptic species are evolutionarily young requires using species delimitation

methods that are sensitive enough to detect early divergence, e.g., before evolving species reach
Accepted Article
reciprocal monophyly. Lineage sorting can take a long time in species with large population sizes

(Dellicour & Flot, 2015; Knowles & Carstens, 2007). Haploweb and multilocus coalescent methods

are both designed to delimit non-monophyletic species. A major impediment to broaden

implementation of the two methods is the low number of nuclear markers exhibiting both suitable

variability and low homoplasy (Kornobis & Pálsson, 2011, 2013; Nahavandi, Ketmaier, & Tiedemann,

2012). Among amphipods, only two markers (ITS, EF1α) apparently meet these conditions (Flot,

Wörheide, & Dattagupta, 2010; Nahavandi et al., 2012). Fortunately, high-throughput sequencing

technologies are becoming readily accessible and affordable, thus multilocus species delimitation

may become standard within the next few years (Raupach, Amann, Wheeler, & Roos, 2016) and the

recent divergence hypothesis may increase in importance (see 3.1).

Second, taxonomy can strengthen the link between species hypotheses and the speciation process

or, more generally, between taxonomic practice and evolutionary theory (de Queiroz, 2007). Only

practices grounded in theory can fully appreciate the heterogeneous nature of speciation, and

appropriately confront multiple data (morphology, ecology, physiology, multiple DNA) to unveil the

full spectrum of divergent processes. This strengthening is a major step towards discovery of general

rules in the speciation process (Butlin et al., 2012) and linking speciation with other fields in

biodiversity science (section 5). Further, understanding why some lineages genetically diverge

without morphological difference, or conversely, why some genetically nearly-identical populations

ecologically differ (Herrel et al., 2008) can increase taxonomy rigor. Some taxonomists consider

species hypotheses supported by multiple lines of evidence as more robust than single line evidence

(Padial et al., 2010). Yet incongruences among data can yield species hypotheses as robust as those

supported by congruence as long as incongruences have sound evolutionary explanations.

This article is protected by copyright. All rights reserved.


Third, novel species, no matter if morphologically cryptic, can be described and named (Jörger &

Schrödl, 2013; Pante et al., 2015). This practice is one way to communicate cryptic species with
Accepted Article
other fields of biodiversity science that use Linnaean binomials in their research. It is paradoxical

that cryptic species remain nameless until described based on morphological criteria. Finding

morphological diagnostic traits among morphologically cryptic species is challenging (C. Fišer &

Zagmajster, 2009; Schlick-Steiner et al., 2010) and the practical value of such diagnoses may be

questionable (Jugovic et al., 2012). Indeed, The International Code of Zoological Nomenclature

(http://iczn.org/code, ICZN ) does not limit the diagnosis of species to the use of morphological

traits. The idea of diagnosing species using DNA sequences is not new (Goldstein & DeSalle, 2011;

Knowlton, 1993) but descriptions relying only on molecular data are generally still rare (Clouse &

Wheeler, 2014; Cook et al., 2010; Edgecombe & Giribet, 2008; Halt, Kupriyanova, Cooper, & Rouse,

2009; M. S. Harvey, Berry, Edward, & Humphreys, 2008; Jörger & Schrödl, 2013; Murphy et al., 2015;

Renner, 2016; Strand & Sundberg, 2011). All these studies broadly agree that holotypes should be

defined and deposited into collections. Technical aspects vary on how species are diagnosed

(Goldstein & DeSalle, 2011; Renner, 2016) and recent works recommend two technical details. First,

diagnosis should rely on diagnostic substitutions rather than genetic distances (i.e., nodes in

phylogeny). Second, detailed alignment construction, sequences and final alignments must be made

available to ensure repeatability and continuity in taxonomy (Jörger & Schrödl, 2013; Renner, 2016).

An important corollary of the molecular description of cryptic species is that species retaining the

original name need to be molecularly diagnosed as well. This step may require designation of

neotypes or co-types that can replace or supplement holotypes if the latter are lost or improperly

preserved for DNA analysis. Finally, formal naming sometimes is not considered desirable when

species hypotheses need to be strengthened with additional lines of evidence of lineage separation.

In such cases, Morard et al. (2016) proposed an interim nomenclature system to codify genetically

circumscribed entities, including potential cryptic species, as an effort to communicate data on

This article is protected by copyright. All rights reserved.


genetic diversity for multiple biodiversity stakeholders. Presently, we can only acknowledge that the

majority of cryptic species uncovered by DNA taxonomy remains silent to biodiversity research.
Accepted Article
5. CRYPTIC DIVERSITY IN BIODIVERSITY RESEARCH

5.1 Biodiversity patterns revisited and consequences for nature conservation

Following the plea by Bickford et al. (2007) to account for cryptic species in biodiversity science, we

now revisit a number of biodiversity-related issues. Notably, the number of examined studies

strongly depended on the scale of the analysis.

5.1.1 Global-scale analyses

Global-scale analyses are few. Pfenninger and Schwenk (2007) conducted a meta-analysis of cryptic

species to test for variation in the proportion of cryptic species across major metazoan phyla and

biogeographical regions. The authors found that cryptic species were homogeneously distributed

among taxa and regions, and concluded “that cryptic metazoan diversity can be treated as random

error in biodiversity assessments”. Their meta-analysis was later criticized for several methodological

problems (Trontelj & Fišer, 2009), and additional analyses almost a decade later suggested that

cryptic species actually may be heterogeneously distributed across animal phyla (Pérez-Ponce de

León & Poulin, 2016). The evidence for homogenous distribution of cryptic species across geographic

regions is conflicting (Eme et al., 2017; Gill et al., 2016; Voda, Dapporto, Dinca, & Vila, 2015);

perhaps an heterogeneous distribution can be detected only at global geographic scales.

Finding that cryptic species are homogeneously distributed across space does not necessarily imply

that the mechanisms causing cryptic species are themselves invariant. These mechanisms can

compensate each other across space to produce an even distribution of cryptic species. An untested

This article is protected by copyright. All rights reserved.


hypothesis is that selectively neutral mechanisms (e.g., the recent divergence hypothesis)

predominate in regions where recent climatic perturbations fragmented ranges of ancestral species
Accepted Article
(e.g., Pleistocene glaciations, Fišer et al., 2010), whereas selection-based mechanisms (PNC,

morphological convergence) operate more frequently in environments with strong directional

selection (Bickford et al., 2007; Trontelj & Fišer, 2009).

It also is unclear why cryptic species are overrepresented in some phyla. Knowlton (1993) suggested

that cryptic species may be more frequent in taxa that do not rely on visual information, but we are

not aware of a single study that explicitly addressed this question. Analogous to geographic aspects

of cryptic diversity, the relative importance of the three mentioned mechanisms causing

morphological crypsis in different taxonomic groups awaits to be quantified and explained. In

particular, PNC may well explain the morphological similarity in organisms, the taxonomy of which

has been based on functional somatic characters related to ecology (e.g., amphipods: Dahl, 1977;

Fišer et al., 2009). In contrast, it may be less important in species delimited essentially on sexual

characters, such as isopods or spiders.

5.1.2 Geographically and phylogenetically limited analyses

Most studies are limited either in geographic or phylogenetic scope. In the amphipod literature,

these include but are not limited to a species geographic range size (Trontelj et al., 2009), the extent

of ecological niche overlap and species coexistence (Ž. Fišer, Altermatt, Zakšek, Knapič, & Fišer,

2015), the strength of size-assortative pairing (Galipaud et al., 2015), the effectiveness of

conservation measures (Venarsky, Anderson, & Wilhelm, 2009), and the use of biological indicators

in ecosystem health assessment (Feckler et al., 2014; Feckler, Schulz, & Bundschuh, 2013; Feckler,

Thielsch, Schwenk, Schulz, & Bundschuh, 2012; Major, Soucek, Giordano, Wetzel, & Soto-Adames,

2013; Soucek, Dickinson, Major, & McEwen, 2013; Weston et al., 2013; Zettler et al., 2013). Notably,

This article is protected by copyright. All rights reserved.


all these studies consider cryptic species as a source of error that must be accounted for when

documenting and explaining biodiversity patterns.


Accepted Article
The question of whether the range size of cryptic species is smaller than that of corresponding

morphological species was addressed explicitly in amphipods from Antarctic deep-sea (Baird et al.,

2012; Lörz, Maas, Linse, & Coleman, 2009), desert springs Australia and North America (Murphy,

Adams, Guzik, & Austin, 2013; Witt, Threloff, & Hebert, 2006) and groundwater amphipods from

Australia, Canary Islands and Europe (e.g., Bauzà-Ribot, Jaume, Fornós, Juan, & Pons, 2011; Bradford

et al., 2013; Bradford, Adams, Humphreys, Austin, & Cooper, 2010; Meleg et al., 2013; Trontelj et al.,

2009a; Villacorta et al., 2008). In general, large range sizes of morphological species were

“symmetrically” divided into a series of smaller ranges for cryptic species. Yet, some studies

reported “asymmetrical” fragmentation where the range of at least one cryptic species was as large

as that of the morphological species (Havermans, Nagy, Sonet, De Broyer, & Martin, 2011; Lefébure

et al., 2006). Whether symmetrical and asymmetrical fragmentations of morphospecies ranges are

homogeneously distributed across geographical regions has yet to be determined. One study

indicates, however, that consideration of cryptic species does not affect continental scale patterns of

range size, the so-called Rapoport effect (i.e., the pattern of increasing range size of morphological

species at higher latitudes in the Palearctic region; Stevens, 1989) (Eme et al., 2017).

Another important aspect is whether morphologically cryptic species have the same ecological

requirements (i.e., Grinnellian niche) and play similar roles in ecosystems (i.e., Eltonian niche) (Ž.

Fišer et al., 2015). Almost all detailed ecological studies showed that cryptic species differed in their

requirements along one or more dimensions of their Grinnellian niche (Colson-Proch, Renault,

Gravot, Douady, & Hervant, 2009; Eisenring, Altermatt, Westram, & Jokela, 2016; Krebes, Blank, &

Bastrop, 2011; Krebes, Blank, Jürss, Zettler, & Bastrop, 2010; Rock, Ironside, Potter, Whiteley, &

Lunt, 2007). Apparently, cryptic species may not be cryptic to other community members. Studies on

cryptic Hyalella species imply different vulnerabilities to fish predation (Cothran, Henderson, et al.,

This article is protected by copyright. All rights reserved.


2013; Wellborn & Cothran, 2004, 2007). Westram et al. (2011) showed different susceptibilities to

acanthocephalan parasites among cryptic species of the Gammarus fossarum species complex, and
Accepted Article
Colson-Proch et al. (2009) found substantial differences in behavioral, metabolic and biochemical

responses to cold among cryptic species of the subterranean amphipod Niphargus

rhenorhodanensis. Ecotoxicological tests further revealed large differences in sensitivity to a number

of pollutants across a range of cryptic taxa (Feckler et al., 2012; Feckler et al., 2014, 2013; Major et

al., 2013; Soucek et al., 2013; Weston et al., 2013). Interpretation of the magnitude of these

differences in an ecological context is highly challenging but studies on co-occurrence of cryptic

species indicate they are insufficient to mediate stable coexistence (Box 2).

Cryptic species having a roughly similar role in the ecosystem (Eltonian niche) but differing in their

environmental requirements and responses to various stressors may be functionally redundant. If

so, cryptic species may be critically important in stabilizing ecosystem functioning (Elmqvist et al.,

2003; E. Harvey, Gounand, Ward, & Altermatt, 2017; Scheffer et al., 2015). To our knowledge, a

single study using cryptic nematodes as a model system explored differences in the functional role of

cryptic species (De Meester, Gingold, Rigaux, Derycke, & Moens, 2016). A microcosm experiment by

these authors unveiled species-specific effects on organic matter decomposition, challenging the

hypothesis of functional redundancy.

Results from biogeographic and ecological studies are difficult for nature conservation. First, range

sizes of many species are much smaller than previously conceived. Species with small range sizes are

much more vulnerable to external disturbances than broadly distributed species and even local

perturbations can lead to biodiversity loss (Bland & Collen, 2016; Delić, Trontelj, et al., 2017).

Second, ignoring the vulnerability of cryptic species to multiple stressors can underestimate threats

to local populations and ultimately biodiversity. Niche models forecasting the future of aquatic

cryptic species, for example, indicate that global climate change will cause the loss of species along

with genetic diversity throughout Europe (Bálint et al., 2011). There is an urgent need for assessing

This article is protected by copyright. All rights reserved.


cryptic diversity to reduce the potential loss from environmental change. One step in this direction

may be predictive phylogeographies. These use environmental data and known phylogeographies of
Accepted Article
co-distributed taxa to predict the presence of cryptic species in other taxa in the same biome

(Espíndola et al., 2016). In conjunction with molecular operational units (discussed below), these

rapid assessments of cryptic diversity can assist towards improving policies and decision making in

conservation.

5.2 Molecular operational taxonomic units

Cryptic species are de facto taken into consideration by an increasing number of biodiversity studies

that use sequence clusters rather than nominal species as units for taxonomic diversity (Ryberg,

2015). Such clusters, referred to as molecular operational taxonomic units (OTUs, also referred as

MOTUs), are delimited using methods presented in Section 2. In the amphipod literature, OTUs were

used to explore determinants of diversification rates (Hou, Sket, Fišer, & Li, 2011; Mamos, Wattier,

Burzyński, & Grabowski, 2016; L. Yang et al., 2013), the geographic origin of European groundwater

fauna (McInerney et al., 2014), and species diversity-environment relationships of local assemblages

(Knox et al., 2012). Importantly, failing to account for cryptic diversity in these studies is equivalent

to omitting known species; thereby underestimating diversification rates, incorrectly reconstructing

ancestral ranges, and undervaluing local diversity (Morvan et al., 2013; Pyron & Burbrink, 2013).

Inferences about eco-evolutionary processes and patterns drawn from OTUs are free from biases

due to morphological crypsis, although still being dependent on the accuracy of molecular

delimitation methods.

The use of OTUs in diversity analyses can be viewed as a logical extension of de Queiroz’s (1999)

view that each biologist can select species properties that are most relevant to the question being

addressed. OTUs cannot be extended across all fields of biodiversity science with similar success,

This article is protected by copyright. All rights reserved.


although they may become mandatory in ecology in the near future (Gill et al., 2016). Neglecting

cryptic species can bias inferences derived from large-scale patterns of phylogenetic diversity and
Accepted Article
phylogenetic turnover, especially when the proportion of convergently evolved cryptic species is

high (Fritz & Rahbek, 2012; Safi et al., 2011). However, applying this approach to conservation

biology (Asmyhr, Linke, Hose, & Nipperess, 2014) and local community ecology (Elbrecht & Leese,

2015; Jones, Ghoorah, & Blaxter, 2011) can be risky. OTUs of unknown vulnerability to different

threats lack protection, and their ecological role in local assemblages cannot be inferred from

phylogenetic relatedness alone (Kembel, 2009; Best & Stachowicz, 2013, 2014; Best, Caulk, &

Stachowicz, 2013).

5.3 Making use of different species criteria

Using OTUs and treating cryptic species as a source of accountable bias are complementary

approaches for integrating cryptic diversity into biodiversity science. However, neither satisfactorily

integrates the various properties of speciation towards a better understanding of biodiversity

patterns. We now discuss how a broader use of species properties in testing eco-evolutionary

hypotheses can foster our understanding of large-scale biodiversity patterns.

The regional number of species is ultimately determined by three processes – speciation, extinction

and dispersal (Wiens, 2011). Further, the contribution of speciation to species richness and the

relative importance of different speciation properties can vary across geographical regions and in

time (Jablonski, Roy, & Valentine, 2006). Thus, geographic variation in the proportion of

morphologically cryptic (OTUs) and morphologically distinguishable species could pinpoint

(hypothesis driven) how environmental factors contribute to species richness patterns via variation

in species properties during speciation. An underlying hypothesis is that morphological and

ecological distinctness can be acquired much faster along a resource gradient with varying selection

This article is protected by copyright. All rights reserved.


pressures (ecological speciation) than from a physical barrier with similar selection pressures in each

environment (non-ecological speciation, perhaps driven by PNC; Wiens et al., 2010; Pyron et al.,
Accepted Article
2015). An important point here is that these hypothesis-driven predictions can be tested by

integrating different species criteria into ecological analyses (Fig. 6).

One step towards this testing consists of building occurrence databases in which thousands of

specimens collected across large geographic regions can be associated to multiple species

hypotheses based on morphology and multiple DNA-based criteria (Fig. 6). Note that species

hypotheses can be refined as more data and methods become available (e.g., new genes, ecological

data). Ultimately, one can assess the relative importance of different mechanisms explaining species

richness patterns for the different species hypotheses. In our example (Fig. 6), the relative

importance of spatial heterogeneity, present climate/productivity, and historical climate variability

for explaining biodiversity patterns is assessed using multi-model inferences and variance

partitioning (see for example Eme et al., 2015, 2017). For illustration, DNA-based criteria increase

the proportion of geographic variation in species richness attributed to spatial heterogeneity due to

the disproportionately higher number of allopatric cryptic species in highly fragmented regions. In

fact, this approach can contribute towards understanding how mechanisms causing cryptic diversity

vary in strength relative to each other across taxa and geographical regions.

Hey et al. (2003) used the term “species uncertainty” to denote the ambiguous correspondence

between a taxon delimited using any criteria and the evolving entity for which it is used in a

hypothesis. The authors argued that “this scientific uncertainty cannot be ‘solved’ or stamped out,

but neither need it be ignored or feared”. Recent progress made in conceptualizing and delimiting

species notably opens the door for broader consideration of this uncertainty in hypothesis testing of

eco-evolutionary theories in biodiversity science.

This article is protected by copyright. All rights reserved.


6. CONCLUSIONS

(1) Cryptic species refer to a heterogeneous and negatively-defined set of species lacking
Accepted Article
morphological distinguishability. Evolutionary mechanisms leading to morphological similarity are

heterogeneous, comprising recent divergence, niche conservatism and morphological convergence.

They are as important and fascinating as mechanisms leading organisms to diversify into a multitude

of phenotypes. We are only just beginning to understand the ‘invisible’ world of cryptic species. We

now need explicit testing of whether morphological traits evolve neutrally, are constrained by

stabilizing selection, or converge from distantly related lineages in response to similar directional

selection on morphology. At present, the relative importance of different explanations of cryptic

diversity remains largely unknown.

(2) Taxonomy can provide evidence for evolutionary explanations if research efforts are oriented

towards species hypotheses that are theoretically and empirically grounded in the speciation

process. DNA taxonomy provides an ever-increasing variety of methods for delimiting cryptic

species, and be used to explore species boundaries related to morphology, ecology, and physiology.

Comprehensive analyses of multiple types of data can unveil generalities in the speciation process,

improve our understanding of the heterogeneity (species properties) in speciation, and allow better

integration in biodiversity science.

(3) Cryptic species are often considered in biodiversity research as evidence of incomplete species

inventory or as potential sources of bias. Paradoxically, the majority of known cryptic species are

undescribed or un-named, and thus overlooked in research fields relying on species binomials.

Studies in evolution and ecology increasingly use molecular operational taxonomic units (OTUs),

thereby inherently incorporating cryptic species into their analyses, but no study to date has used

multiple criteria for species delineation. Extending multi-criteria approaches to different fields of

biodiversity science can better integrate cryptic species in ecological and evolutionary theories,

paving the way to new frontiers in biodiversity science. For example, are biodiversity patterns

This article is protected by copyright. All rights reserved.


inferred from morphologically and DNA delimited species discordant? If so, is discrepancy a

contingency to ecological and processes generating cryptic species? How do different mechanisms
Accepted Article
causing cryptic diversity vary in strength relative to each other across geographical space, and

environmental gradients? And lastly, do coexistence mechanisms at local scales change as lineage

divergence proceeds?

(4) The phenomena of cryptic and ecological species can be parallel. Importantly, cryptic species

clearly complement research on speciation. We do envision that with the development of new

methodologies, cryptic species pervasiveness within the broader context of biodiversity science will

likely fade over time and be viewed by the next generation of scientists as a constructive

epiphenomenon towards integrating speciation and its heterogeneity (species properties) into

biodiversity science. This integration necessarily must include integrative taxonomy and the use of

multiple species criteria, as already evident in basic biodiversity research. Importantly, shifts to

integrative taxonomy and multi-criteria have profound implications in terms of forecasting and

consolidating new skills and occupations in biodiversity assessment and conservation. In the future,

conservationists and policy-makers must be prepared to make informed decisions using a more

complex but novel body of knowledge on species diversity and distribution. For example, this

information may include continental maps of species richness and geographic range size using

different species delimitation methods that can pinpoint regions in method incongruences. These

incongruences then will reveal geographic variation in the speciation process. Thus, better

understanding of the causes of this variation in speciation can reinforce process-oriented

conservation strategies. Lastly, incongruences that cannot be readily explained can promote the

development of adaptive conservation principles and public policies that explicitly acknowledge the

scientific uncertainty of species entities.

This article is protected by copyright. All rights reserved.


AUTHOR CONTRIBUTION

C.F. gathered the literature data, all authors contributed original ideas and helped draft the paper.
Accepted Article
CONFLICT OF INTERESTS

Authors declare no conflict of interest.

ACKNOWLEDGEMENTS

This study was funded by the PHC Proteus programme (no. 31199UM), the Slovenian Research

Agency (ARRS, Programme P0-0184) and the Swiss National Science Foundation (grant

IZK0Z3_169642). Anne-Nina Lörz analyzed data from Amphipoda Newsletters. We thank F.

Altermatt, T. Datry, C. Douady, D. Eme, and J.-F. Flot for comments on early versions of the

manuscript. This article is based upon work from COST Action DNAqua-Net (CA15219), supported by

the COST (European Cooperation in Science and Technology) program.

REFERENCES

Achurra, A., Rodriguez, P., & Erséus, C. (2015). Pseudo-cryptic speciation in the subterranean

medium: A new species of Stylodrilus Claparède, 1862, with a revision of the status of Bichaeta

Bretscher, 1900 (Annelida, Clitellata, Lumbriculidae). Zoologischer Anzeiger - A Journal of

Comparative Zoology, 257, 71–86. https://doi.org/10.1016/j.jcz.2015.05.003

Adams, D. C., Berns, C. M., Kozak, K. H., & Wiens, J. J. (2009). Are rates of species diversification

correlated with rates of morphological evolution? Proceedings of the Royal Society B: Biological

Sciences, 276(1668), 2729–2738. https://doi.org/10.1098/rspb.2009.0543

This article is protected by copyright. All rights reserved.


Adams, M., Raadik, T. A., Burridge, C. P., & Georges, A. (2014). Global biodiversity assessment and

hyper-cryptic species complexes: More than one species of elephant in the room? Systematic
Accepted Article
Biology, 63(4), 518–533. https://doi.org/10.1093/sysbio/syu017

Agapow, P. M., Emonds, O. R. P. B. E., Crandall, K. A., Gittleman, J. C., Mace, G. M., Marshall, J. C., &

Purvis, A. (2004). The Impact of Species Concept on Biodiversity Studies. The Quarterly Review

of Biology, 79(2), 161–179.

Alexander, M. E., Dick, J. T. A., & O’Connor, N. E. (2013). Born to kill: Predatory functional responses

of the littoral amphipod Echinogammarus marinus Leach throughout its life history. Journal of

Experimental Marine Biology and Ecology, 439, 92–99.

https://doi.org/10.1016/j.jembe.2012.10.006

Ashrafi, S., Beck, A., Rutishauser, M., Arlettaz, R., & Bontadina, F. (2011). Trophic niche partitioning

of cryptic species of long-eared bats in Switzerland: Implications for conservation. European

Journal of Wildlife Research, 57(4), 843–849. https://doi.org/10.1007/s10344-011-0496-z

Asmyhr, M. G., Linke, S., Hose, G., & Nipperess, D. A. (2014). Systematic Conservation Planning for

Groundwater Ecosystems Using Phylogenetic Diversity. PLoS ONE One, 9, e115132.

https://doi.org/10.1371/journal.pone.0115132

Baird, H. P., Miller, K. J., & Stark, J. S. (2012). Genetic Population Structure in the Antarctic Benthos :

Insights from the Widespread Amphipod , Orchomenella franklini. PloS One, 7(3), e34363.

https://doi.org/10.1371/journal.pone.0034363

Bálint, M., Domisch, S., Engelhardt, C. H. M., Haase, P., Lehrian, S., Sauer, J., … Nowak, C. (2011).

Cryptic biodiversity loss linked to global climate change. Nature Climate Change, 1, 313–318.

https://doi.org/10.1038/nclimate1191

Bauzà-Ribot, M. M., Jaume, D., Fornós, J. J., Juan, C., & Pons, J. (2011). Islands beneath islands:

This article is protected by copyright. All rights reserved.


phylogeography of a groundwater amphipod crustacean in the Balearic archipelago. BMC

Evolutionary Biology, 11(1), 221. https://doi.org/10.1186/1471-2148-11-221


Accepted Article
Beheregaray, L. B., & Caccone, A. (2007). Cryptic biodiversity in a changing world. Journal of Biology,

6(4), 9. https://doi.org/10.1186/jbiol60

Best, R. J., Caulk, N. C., & Stachowicz, J. J. (2013). Trait vs. phylogenetic diversity as predictors of

competition and community composition in herbivorous marine amphipods. Ecology Letters,

16(1), 72–80. https://doi.org/10.1111/ele.12016

Best, R. J., & Stachowicz, J. J. (2013). Phylogeny as a Proxy for Ecology in Seagrass Amphipods: Which

Traits Are Most Conserved? PLoS ONE, 8(3), 1–14.

https://doi.org/10.1371/journal.pone.0057550

Best, R. J., & Stachowicz, J. J. (2014). Phenotypic and phylogenetic evidence for the role of food and

habitat in the assembly of communities of marine amphipods. Ecology, 95(3), 775–786.

https://doi.org/10.1890/13-0163.1

Bickford, D., Lohman, D. J., Sodhi, N. S., Ng, P. K. L., Meier, R., Winker, K., … Das, I. (2007). Cryptic

species as a window on diversity and conservation. Trends in Ecology & Evolution, 22(3), 148–

155. https://doi.org/10.1016/j.tree.2006.11.004

Bland, L., & Collen, B. (2016). Species loss: lack of data leaves a gap. Nature, 537, 488.

https://doi.org/10.1049/ep.1974.0517

Bock, W. J. (2004). Species: The concept, category and taxon. Journal of Zoological Systematics and

Evolutionary Research, 42(3), 178–190. https://doi.org/10.1111/j.1439-0469.2004.00276.x

Brad, T., Fišer, C., Flot, J.-F., & Sarbu, S. M. (2015). Niphargus dancaui sp . nov . (Amphipoda ,

Niphargidae) – a new species thriving in sulfidic groundwaters in southeastern Romania.

European Journal of Taxonomy, 164, 1–28.

This article is protected by copyright. All rights reserved.


https://doi.org/http://dx.doi.org/10.5852/ejt.2015.164 ISSN

Bradford, T. M., Adams, M., Guzik, M. T., Humphreys, W. F., Austin, A. D., & Cooper, S. J. B. (2013).
Accepted Article
Patterns of population genetic variation in sympatric chiltoniid amphipods within a calcrete

aquifer reveal a dynamic subterranean environment. Heredity, 111(1), 77–85.

https://doi.org/10.1038/hdy.2013.22

Bradford, T. M., Adams, M., Humphreys, W. F., Austin, A. D., & Cooper, S. J. B. (2010). DNA barcoding

of stygofauna uncovers cryptic amphipod diversity in a calcrete aquifer in Western Australia’s

arid zone. Molecular Ecology Resources, 10(1), 41–50. https://doi.org/10.1111/j.1755-

0998.2009.02706.x

Bravo, G. A., Remsen, J. V, & Brumfield, R. T. (2014). Adaptive processes drive ecomorphological

convergent evolution in antwrens (Thamnophilidae). Evolution, 68(10), 2757–2774.

https://doi.org/10.1111/evo.12506

Butler, M., & King, A. (2004). Phylogenetic comparative analysis: a modeling approach for adaptive

evolution. The American Naturalist, 164(6), 683–695. Retrieved from

http://www.jstor.org/stable/10.1086/426002

Butlin, R., Debelle, A., Kerth, C., Snook, R. R., Beukeboom, L. W., Castillo Cajas, R. F., … Schilthuizen,

M. (2012). What do we need to know about speciation? Trends in Ecology and Evolution, 27(1),

27–39. https://doi.org/10.1016/j.tree.2011.09.002

Camargo, A., & Sites, J. J. (2013). Species Delimitation: A Decade After the Renaissance. In I. Y.

Pavlinov (Ed.), The species problem - ongoing issues. (pp. 225–247). InTech.

https://doi.org/10.5772/52664

Carroll, S. P., Hendry, A. P., Reznick, D. N., & Fox, C. W. (2007). Evolution on Ecological Time-Scales.

Functional Ecology, 21(3), 387–393.

This article is protected by copyright. All rights reserved.


Carstens, B. C., Pelletier, T. A., Reid, N. M., & Satler, J. D. (2013). How to fail at species delimitation.

Molecular Ecology, 22(17), 4369–4383. https://doi.org/10.1111/mec.12413


Accepted Article
Chesson, P. (2000). Mechanisms of Maintenance of Species Diversity. Annual Review of Ecology and

Systematics, 31, 343–366. https://doi.org/10.1146/annurev.ecolsys.31.1.343

Clavel, J., Escarguel, G., & Merceron, G. (2015). mvMORPH: an R Package for Fitting Multivariate

Evolutionary Models To Morphometric Data. Methods in Ecology and Evolution, 6, 1311–1319.

https://doi.org/10.1111/2041-210X.12420

Clouse, R. M., & Wheeler, W. C. (2014). Descriptions of two new, cryptic species of Metasiro

(Arachnida: Opiliones: Cyphophthalmi: Neogoveidae) from South Carolina, USA, including a

discussion of mitochondrial mutation rates. Zootaxa, 3814(2), 177–201.

https://doi.org/10.11646/zootaxa.3814.2.2

Collar, D. C., Schulte, J. A. I., & Losos, J. B. (2011). Evolution of Extreme Body Size Disparity in

Monitor Lizards (Varanus). Evolution, 65(9), 2664–2680. https://doi.org/10.1111/j.1558-

5646.2011.01335.x

Colson-Proch, C., Renault, D., Gravot, A., Douady, C. J., & Hervant, F. (2009). Do current

environmental conditions explain physiological and metabolic responses of subterranean

crustaceans to cold? Journal of Experimental Biology, 212(12), 1859–1868.

https://doi.org/10.1242/jeb.027987

Cook, L. G., Edwards, R. D., Crisp, M. D., & Hardy, N. B. (2010). Need morphology always be required

for new species descriptions? Invertebrate Systematics, 24(3), 322–326.

https://doi.org/10.1071/IS10011

Copilaş-Ciocianu, D., & Petrusek, A. (2015). The southwestern Carpathians as an ancient centre of

diversity of freshwater gammarid amphipods: insights from the Gammarus fossarum species

This article is protected by copyright. All rights reserved.


complex. Molecular Ecology, 24(15), 3980–3992. https://doi.org/10.1111/mec.13286

Corrigan, L. J., Horton, T., Fotherby, H., White, T. A., & Hoelzel, A. R. (2013). Adaptive Evolution of
Accepted Article
Deep-Sea Amphipods from the Superfamily Lysiassanoidea in the North Atlantic. Evolutionary

Biology, 154–165. https://doi.org/10.1007/s11692-013-9255-2

Cothran, R. D., Henderson, K. A., Schmidenberg, D., & Relyea, R. A. (2013). Phenotypically similar but

ecologically distinct: differences in competitive ability and predation risk among amphipods.

Oikos, 122(January), 1429–1440. https://doi.org/10.1111/j.1600-0706.2013.00294.x

Cothran, R. D., Noyes, P., & Relyea, R. A. (2015). An empirical test of stable species coexistence in an

amphipod species complex. Oecologia, 178(3), 819–831. https://doi.org/10.1007/s00442-015-

3262-1

Cothran, R. D., Stiff, A. R., Chapman, K., Wellborn, G. A., & Relyea, R. A. (2013). Reproductive

interference via interspecific pairing in an amphipod species complex. Behavioral Ecology and

Sociobiology, 67(8), 1357–1367. https://doi.org/Doi 10.1007/S00265-013-1564-Z

Cristescu, M. E. A., Witt, J. D. S., Grigorovich, I. A., Hebert, P. D. N., & MacIsaac, H. J. (2004). Dispersal

of the Ponto-Caspian amphipod Echinogammarus ischnus: invasion waves from the Pleistocene

to the present. Heredity, 92(3), 197–203. https://doi.org/10.1038/sj.hdy.6800395

Dahl, E. (1977). The amphipod functional model and its bearing upon systematics and phylogeny.

Zoologica Scripta, 6, 221–228. https://doi.org/10.1111/j.1463-6409.1978.tb00773.x

De Meester, N., Derycke, S., Bonte, D., & Moens, T. (2011). Salinity effects on the coexistence of

cryptic species: a case study on marine nematodes. Marine Biology, 158(12), 2717–2726.

https://doi.org/10.1007/s00227-011-1769-5

De Meester, N., Gingold, R., Rigaux, A., Derycke, S., & Moens, T. (2016). Cryptic diversity and

ecosystem functioning: a complex tale of differential effects on decomposition. Oecologia.

This article is protected by copyright. All rights reserved.


https://doi.org/10.1007/s00442-016-3677-3

de Queiroz, K. (1998). The General Lineage Concept of Species , Species Criteria , and the Process of
Accepted Article
Speciation and Terminological Recommendations. In D. J. Howard & S. H. Berlocher (Eds.),

Endless Forms: Species and Speciation (pp. 57–75). Oxford: Oxford University Press.

de Queiroz, K. (2005a). A unified species concept and its consequences for the future of taxonomy.

Proceedings of the California Academy of Sciences, 56(June), 196–215.

https://doi.org/10.1073/pnas.0502030102

de Queiroz, K. (2005b). Different species problems and their resolution. BioEssays, 27(12), 1263–

1269. https://doi.org/10.1002/bies.20325

de Queiroz, K. (2007). Species concepts and species delimitation. Systematic Biology, 56(6), 879–886.

https://doi.org/10.1080/10635150701701083

Delić, T., Švara, V., Coleman, C. O., Trontelj, P., & Fišer, C. (2017). The giant cryptic amphipod species

of the subterranean genus Niphargus (Crustacea, Amphipoda). Zoologica Scripta, 46(6), 740–

752. https://doi.org/10.1111/zsc.12252

Delić, T., Trontelj, P., Rendoš, M., & Fišer, C. (2017). The importance of naming cryptic species and

the conservation of endemic subterranean amphipods. Scientific Reports, 7, 3391.

https://doi.org/10.1038/s41598-017-02938-z

Dellicour, S., & Flot, J.-F. (2015). Delimiting Species-Poor Data Sets using Single Molecular Markers: A

Study of Barcode Gaps, Haplowebs and GMYC. Systematic Biology, 64(6), 900–908.

https://doi.org/10.1093/sysbio/syu130

Derycke, S., Remerie, T., Backeljau, T., Vierstraete, A., Vanfleteren, J., Vincx, M., & Moens, T. (2008).

Phylogeography of the Rhabditis ( Pellioditis ) marina species complex: evidence for long-

distance dispersal, and for range expansions and restricted gene flow in the northeast Atlantic.

This article is protected by copyright. All rights reserved.


Molecular Ecology, 17(14), 3306–3322. https://doi.org/10.1111/j.1365-294X.2008.03846.x

Dick, J. T., & Platvoet, D. (2000). Invading predatory crustacean Dikerogammarus villosus eliminates
Accepted Article
both native and exotic species. Proceedings of the Royal Society of London, 267(267), 977–983.

https://doi.org/10.1098/rspb.2000.1099

Dionne, K., Vergilino, R., Dufresne, F., Charles, F., & Nozais, C. (2011). No Evidence for Temporal

Variation in a Cryptic Species Community of Freshwater Amphipods of the Hyalella azteca

Species Complex. Diversity, 3(4), 390–404. https://doi.org/10.3390/d3030390

Dodd, J. A., Dick, J. T. A., Alexander, M. E., MacNeil, C., Dunn, A. M., & Aldridge, D. C. (2014).

Predicting the ecological impacts of a new freshwater invader: functional responses and prey

selectivity of the “killer shrimp”, Dikerogammarus villosus , compared to the native Gammarus

pulex. Freshwater Biology, 59(2), 337–352. https://doi.org/10.1111/fwb.12268

dos Reis, M., Donoghue, P. C. J., & Yang, Z. (2015). Bayesian molecular clock dating of species

divergences in the genomics era. Nature Reviews Genetics, 17(2), 1–10.

https://doi.org/10.1038/nrg.2015.8

Edgecombe, G. D., & Giribet, G. (2008). A New Zealand species of the trans-Tasman centipede order

Craterostigmomorpha (Arthropoda:Chilopoda) corroborated by molecular evidence.

Invertebrate Systematics, 22(1), 1–15. https://doi.org/10.1071/IS07036

Edwards, D. L., & Knowles, L. L. (2014). Species detection and individual assignment in species

delimitation: can integrative data increase efficacy? Proceedings of the Royal Society B:

Biological Sciences, 281(January), 20132765. https://doi.org/10.1098/rspb.2013.2765

Egea, E., David, B., Choné, T., Laurin, B., Féral, J. P., & Chenuil, A. (2016). Morphological and genetic

analyses reveal a cryptic species complex in the echinoid Echinocardium cordatum and rule out

a stabilizing selection explanation. Molecular Phylogenetics and Evolution, 94, 207–220.

This article is protected by copyright. All rights reserved.


https://doi.org/10.1016/j.ympev.2015.07.023

Eisenring, M., Altermatt, F., Westram, A. M., & Jokela, J. (2016). Habitat requirements and ecological
Accepted Article
niche of two cryptic amphipod species at landscape and local scales. Ecoshphere, 7(5), 1–13

(e01319).

Elbrecht, V., & Leese, F. (2015). Can DNA-Based Ecosystem Assessments Quantify Species

Abundance? Testing Primer Bias and Biomass—Sequence Relationships with an Innovative

Metabarcoding Protocol. Plos One, 10(7), e0130324.

https://doi.org/10.1371/journal.pone.0130324

Elmqvist, T., Folke, C., Nyström, M., Peterson, G., Bengtsson, J., Walker, B., & Norberg, J. (2003).

Response diversity, ecosystem change, and resilience. Frontiers in Ecology and the

Environment, 1(9), 488–494. https://doi.org/10.1890/1540-

9295(2003)001[0488:RDECAR]2.0.CO;2

Eme, D., Zagmajster, M., Delić, T., Fišer, C., Flot, J. F., Konecny-Dupre, L., … Malard, F. (2017). Do

cryptic species matter in macroecology? Sequencing European groundwater crustaceans yields

smaller ranges but does not challenge biodiversity determinants. Ecography, (February), 1–13.

https://doi.org/10.1111/ecog.02683

Eme, D., Zagmajster, M., Fišer, C., Galassi, D., Marmonier, P., Stoch, F., … Malard, F. (2015). Multi-

causality and spatial non-stationarity in the determinants of groundwater crustacean diversity

in Europe. Ecography, 38(5), 531–540. https://doi.org/10.1111/ecog.01092

Ence, D. D., & Carstens, B. C. (2011). SpedeSTEM: a rapid and accurate method for species

delimitation. Molecular Ecology Resources, 11(3), 473–480. https://doi.org/10.1111/j.1755-

0998.2010.02947.x

Esmaeili-Rineh, S., Sari, A., Delić, T., Moškrič, A., & Fišer, C. (2015). Molecular phylogeny of the

This article is protected by copyright. All rights reserved.


subterranean genus Niphargus ( Crustacea : Amphipoda ) in the Middle East : a comparison

with European Niphargids. Zoological Journal of the Linnean Society, 174, 812–826.
Accepted Article
https://doi.org/10.1111/zoj.12296

Espíndola, A., Ruffley, M., Smith, M. L., Carstens, B. C., Tank, D. C., & Sullivan, J. (2016). Identifying

cryptic diversity with predictive phylogeography. Proceedings of the Royal Society B: Biological

Sciences, 283(1841), 20161529. https://doi.org/10.1098/rspb.2016.1529

Feckler, A., Schulz, R., & Bundschuh, M. (2013). Cryptic lineages-same but different? Integrated

Environmental Assessment and Management, 9(1), 172–173.

https://doi.org/10.1002/ieam.1370

Feckler, A., Thielsch, A., Schwenk, K., Schulz, R., & Bundschuh, M. (2012). Differences in the

sensitivity among cryptic lineages of the Gammarus fossarum complex. Science of the Total

Environment, 439, 158–164. https://doi.org/10.1016/j.scitotenv.2012.09.003

Feckler, A., Zubrod, J. P., Thielsch, A., Schwenk, K., Schulz, R., & Bundschuh, M. (2014). Cryptic

species diversity: an overlooked factor in environmental management? Journal of Applied

Ecology, 51(4), 958–967. https://doi.org/10.1111/1365-2664.12246

Felsenstein, J. (1985). Phylogenies and the Comparative Method. The American Naturalist, 125(1), 1–

15.

Fennessy, J., Bidon, T., Reuss, F., Vamberger, M., Fritz, U., Janke Correspondence, A., … Janke, A.

(2016). Multi-locus Analyses Reveal Four Giraffe Species Instead of One. Current Biology,

26(18), 1–7. https://doi.org/10.1016/j.cub.2016.07.036

Fišer, C., Coleman, C. O., Zagmajster, M., Zwittnig, B., Gerecke, R., & Sket, B. (2010). Old museum

samples and recent taxonomy: A taxonomic, biogeographic and conservation perspective of

the Niphargus tatrensis species complex (Crustacea: Amphipoda). Organisms Diversity &

This article is protected by copyright. All rights reserved.


Evolution, 10, 5–22. https://doi.org/10.1007/s13127-010-0006-2

Fišer, C., Sket, B., & Trontelj, P. (2008). A phylogenetic perspective on 160 years of troubled
Accepted Article
taxonomy of Niphargus (Crustacea: Amphipoda). Zoologica Scripta, 37, 665–680.

https://doi.org/10.1111/j.1463-6409.2008.00347.x

Fišer, C., Trontelj, P., Luštrik, R., & Sket, B. (2009). Toward a unified taxonomy of Niphargus

(Crustacea: Amphipoda): a review of morphological variability. Zootaxa, 2061, 1–22.

Fišer, C., & Zagmajster, M. (2009). Cryptic species from cryptic space: the case of Niphargus fongi sp.

n. (Amphipoda, Niphargidae). Crustaceana, 82, 593–614.

https://doi.org/10.1163/156854009X407704

Fišer, Ž., Altermatt, F., Zakšek, V., Knapič, T., & Fišer, C. (2015). Morphologically cryptic Amphipod

species sre “ecological clones” at regional but not at local scale: a case study of four Niphargus

species. PLoS ONE, 10, e0134384. https://doi.org/10.1371/journal.pone.0134384

Flot, J.-F. (2015). Species delimitation’s coming of age. Systematic Biology, 64(6), 897–899.

Flot, J.-F., Bauermeister, J., Brad, T., Hillebrand-Voiculescu, A., Sarbu, S. M., & Dattagupta, S. (2014).

Niphargus - Thiothrix associations may be widespread in sulphidic groundwater ecosystems:

evidence from southeastern Romania. Molecular Ecology, 23, 1405–1417.

https://doi.org/10.1111/mec.12461

Flot, J.-F., Couloux, A., & Tillier, S. (2010). Haplowebs as a graphical tool for delimiting species: a

revival of Doyle’s “field for recombination” approach and its application to the coral genus

Pocillopora in Clipperton. BMC Evolutionary Biology, 10(1), 372. https://doi.org/10.1186/1471-

2148-10-372

Flot, J.-F., Wörheide, G., & Dattagupta, S. (2010). Unsuspected diversity of Niphargus amphipods in

the chemoautotrophic cave ecosystem of Frasassi, central Italy. BMC Evolutionary Biology, 10,

This article is protected by copyright. All rights reserved.


1–18. https://doi.org/10.1186/1471-2148-10-171

Fontaneto, D., Flot, J.-F., & Tang, C. Q. (2015). Guidelines for DNA taxonomy, with a focus on the
Accepted Article
meiofauna. Marine Biodiversity, 1–19. https://doi.org/10.1007/s12526-015-0319-7

Fritz, S. A., & Rahbek, C. (2012). Global patterns of amphibian phylogenetic diversity. Journal of

Biogeography, 39(8), 1373–1382. https://doi.org/10.1111/j.1365-2699.2012.02757.x

Fujita, M. K., Leaché, A. D., Burbrink, F. T., McGuire, J. A., & Moritz, C. (2012). Coalescent-based

species delimitation in an integrative taxonomy. Trends in Ecology & Evolution, 27(9), 480–488.

https://doi.org/10.1016/j.tree.2012.04.012

Galipaud, M., Bollache, L., Wattier, R., Dubreuil, C., Dechaume-Moncharmont, F.-X., & Lagrue, C.

(2015). Overestimation of the strength of size-assortative pairing in taxa with cryptic diversity:

a case of Simpson’s paradox. Animal Behaviour, 102, 217–221.

https://doi.org/10.1016/j.anbehav.2015.01.032

Gill, B. A., Kondratieff, B. C., Casner, K. L., Encalada, A. C., Flecker, A. S., Gannon, D. G., … Funk, W. C.

(2016). Cryptic species diversity reveals biogeographic support for the “mountain passes are

higher in the tropics” hypothesis. Proceedings of The Royal Society B, 283(October), 7–12.

https://doi.org/10.1098/rspb.2016.0553

Goldstein, P. Z., & DeSalle, R. (2011). Integrating DNA barcode data and taxonomic practice:

Determination, discovery, and description. BioEssays, 33(2), 135–147.

https://doi.org/10.1002/bies.201000036

Grabowski, M., Bacela, K., & Konopacka, A. (2007). How to be an invasive gammarid (Amphipoda:

Gammaroidea)–comparison of life history traits. Hydrobiologia, 590(1), 75–84.

https://doi.org/10.1007/s10750-007-0759-6

Grummer, J. A., Bryson, R. W., & Reeder, T. W. (2014). Species delimitation using bayes factors:

This article is protected by copyright. All rights reserved.


Simulations and application to the sceloporus scalaris species group (Squamata:

Phrynosomatidae). Systematic Biology, 63(2), 119–133. https://doi.org/10.1093/sysbio/syt069


Accepted Article
Guerra-García, J. M., Redondo-Gómez, S., Espina, Á. G., Castillo, J. M., Luque, T., García-Gómez, J. C.,

& Enrique Figueroa, M. (2006). Caprella penantis Leach, 1814 and Caprella dilatata Kroyer,

1843 (Crustacea: Amphipoda) from the Strait of Gibraltar: a molecular approach to explore

intra- and interspecific variation. Marine Biology Research, 2(2), 100–108.

https://doi.org/10.1080/17451000600672511

Guillot, G., Renaud, S., Ledevin, R., Michaux, J., & Claude, J. (2012). A unifying model for the analysis

of phenotypic, genetic, and geographic data. Systematic Biology, 61(6), 897–911.

https://doi.org/10.1093/sysbio/sys038

Halt, M. N., Kupriyanova, E. K., Cooper, S. J., & Rouse, G. W. (2009). Namig species with no

morphological indicators: species status of Galeolaria caespitosa (Annelida: Serpulidae)

inferred from nuclear and mitochondrial gene sequences and morphology. Invertebrate

Systematics, 23, 205–222. Retrieved from http://www.publish.csiro.au/?paper=IS09003

Hansen, T. F. (1997). Stabilizing Selection and the Comparative Analysis of Adaptation. Evolution,

51(5), 1341–1351.

Harmon, L. J., Losos, J. B., Jonathan Davies, T., Gillespie, R. G., Gittleman, J. L., Bryan Jennings, W., …

Mooers, A. Ø. (2010). Early Bursts of Body Size and Shape Evolution Are Rare in Comparative

Data. Evolution, 64, 2385–2369. https://doi.org/10.1111/j.1558-5646.2010.01025.x

Harmon, L. J., Schulte, J. A., Larson, A., & Losos, J. B. (2003). Tempo and mode of evolutionary

radiation in iguanian lizards. Science, 301(5635), 961–964.

https://doi.org/10.1126/science.1084786

Harvey, E., Gounand, I., Ward, C., & Altermatt, F. (2017). Bridging ecology and conservation: from

This article is protected by copyright. All rights reserved.


ecological networks to ecosystem function. Journal of Applied Ecology, 54(2), 371–379.

https://doi.org/10.1111/1365-2664.12769
Accepted Article
Harvey, M. S., Berry, O., Edward, K. L., & Humphreys, G. (2008). Molecular and morphological

systematics of hypogean schizomids (Schizomida:Hubbardiidae) in semiarid Australia.

Invertebrate Systematics, 22(2), 167–194. https://doi.org/10.1071/IS07026

Havermans, C. (2016). Have we so far only seen the tip of the iceberg? Exploring species diversity

and distribution of the giant amphipod Eurythenes. Biodiversity, 8386(April), 1–14.

https://doi.org/10.1080/14888386.2016.1172257

Havermans, C., Nagy, Z. T., Sonet, G., De Broyer, C., & Martin, P. (2011). DNA barcoding reveals new

insights into the diversity of Antarctic species of Orchomene sensu lato (Crustacea: Amphipoda:

Lysianassoidea). Deep Sea Research Part II: Topical Studies in Oceanography, 58(1–2), 230–241.

https://doi.org/10.1016/j.dsr2.2010.09.028

Hebert, P. D. N., Cywinska, A., Ball, S. L., & DeWaard, J. R. (2003). Biological identifications through

DNA barcodes. Proceedings of The Royal Society B, 270(1512), 313–21.

https://doi.org/10.1098/rspb.2002.2218

Hebert, P. D. N., Ratnasingham, S., & Waard, J. (2003). Barcoding animal life : cytochrome c oxidase

subunit 1 divergences among closely related species Barcoding animal life : cytochrome c

oxidase subunit 1 divergences among closely related species. Proceedings of Royal Society of

Biology (Supplement), 270, S96–S99. https://doi.org/10.1098/rsbl.2003.0025

Herrel, A., Huyghe, K., Vanhooydonck, B., Backeljau, T., Breugelmans, K., Grbac, I., … Irschick, D. J.

(2008). Rapid large-scale evolutionary divergence in morphology and performance associated

with exploitation of a different dietary resource. Proceedings of the National Academy of

Sciences of the United States of America, 105(12), 4792–5.

https://doi.org/10.1073/pnas.0711998105

This article is protected by copyright. All rights reserved.


Hey, J. (2006). On the failure of modern species concepts. Trends in Ecology and Evolution, 21(8),

447–450. https://doi.org/10.1016/j.tree.2006.05.011
Accepted Article
Hey, J., Waples, R. S., Arnold, M. L., Butlin, R. K., & Harrison, R. G. (2003). Understanding and

confronting species uncertainty in biology and conservation. Trends in Ecology and Evolution,

18(11), 597–603. https://doi.org/10.1016/j.tree.2003.08.014

Horton, T., Lowry, J. K., De Broyer, C., Bellan-Santini, D., Coleman, C. O., Daneliya, M. E., … Zeidler,

W. (2016). World Amphipoda Database. Retrieved April 19, 2016, from

http://www.marinespecies.org/amphipoda

Hou, Z., & Li, S. (2010). Intraspecific or interspecific variation: delimitation of species boundaries

within the genus Gammarus (Crustacea, Amphipoda, Gammaridae), with description of four

new species. Zoological Journal of the Linnean Society, 160(2), 215–253.

https://doi.org/10.1111/j.1096-3642.2009.00603.x

Hou, Z., Sket, B., Fišer, C., & Li, S. (2011). Eocene habitat shift from saline to freshwater promoted

Tethyan amphipod diversification. Proceedings of the National Academy of Sciences of the

United States of America, 108(35), 14533–14538. https://doi.org/10.1073/pnas.1104636108

Ingram, T., & Mahler, D. L. (2013). SURFACE: Detecting convergent evolution from comparative data

by fitting Ornstein-Uhlenbeck models with stepwise Akaike Information Criterion. Methods in

Ecology and Evolution, 4(5), 416–425. https://doi.org/10.1111/2041-210X.12034

Jablonski, D., Roy, K., & Valentine, J. W. (2006). Out of the Tropics: Evolutionary Dynamics of the

Latitudinal Diversity Gradient. Science, 314(5796), 102–106.

https://doi.org/10.1126/science.1130880

Jamieson, A. J., Fujii, T., Mayor, D. J., Solan, M., & Priede, I. G. (2010). Hadal trenches: the ecology of

the deepest places on Earth. Trends in Ecology and Evolution, 25(3), 190–197.

This article is protected by copyright. All rights reserved.


https://doi.org/10.1016/j.tree.2009.09.009

Jazdzewski, K., Konopacka, A., & Grabowski, M. (2004). Recent drastic changes in the gammarid
Accepted Article
fauna (Crustacea, Amphipoda) of the Vistula River deltaic system in Poland caused by alien

invaders. Diversity and Distributions, 10(2), 81–87. https://doi.org/10.1111/j.1366-

9516.2004.00062.x

Jones, M., Ghoorah, A., & Blaxter, M. (2011). JMOTU and taxonerator: Turning DNA barcode

sequences into annotated operational taxonomic units. PLoS ONE, 6(4).

https://doi.org/10.1371/journal.pone.0019259

Jörger, K. M., & Schrödl, M. (2013). How to describe a cryptic species? Practical challenges of

molecular taxonomy. Frontiers in Zoology, 10(1), 59. https://doi.org/10.1186/1742-9994-10-59

Jugovic, J., Jalžić, B., Prevorčnik, S., & Sket, B. (2012). Cave shrimps Troglocaris s. str. (Dormitzer,

1853), taxonomic revision and description of new taxa after phylogenetic and morphometric

studies. Zootaxa, 3421, 1–31.

Kaliszewska, Z. A., Seger, J., Rowntree, V. J., Barco, S. G., Benegas, R., Best, P. B., … Yamada, T. K.

(2005). Population histories of right whales (Cetacea:Eubalaena) inferred from mitochondrial

sequence diversities and divergences of thier whale lice (Amphipoda: Cyamus). Molecular

Ecology, 14, 3439–3456. https://doi.org/10.1111/j.1365-294X.2005.02664.x

Karanovic, T., Djurakic, M., & Eberhard, S. M. (2016). Cryptic Species or Inadequate Taxonomy?

Implementation of 2D Geometric Morphometrics Based on Integumental Organs as Landmarks

for Delimitation and Description of Copepod Taxa. Systematic Biology, 65(2), 304–327.

https://doi.org/10.1093/sysbio/syv088

Katouzian, A.-R., Sari, A., Macher, J. N., Weiss, M., Saboori, A., Leese, F., & Weigand, A. M. (2016).

Drastic underestimation of amphipod biodiversity in the endangered Irano-Anatolian and

This article is protected by copyright. All rights reserved.


Caucasus biodiversity hotspots. Scientific Reports, 6, 22507. https://doi.org/10.1038/srep22507

Kearney, M. R., Wintle, B. A., & Porter, W. P. (2010). Correlative and mechanistic models of species
Accepted Article
distribution provide congruent forecasts under climate change. Conservation Letters, 3(3), 203–

213. https://doi.org/10.1111/j.1755-263X.2010.00097.x

Kembel, S. W. (2009). Disentangling niche and neutral influences on community assembly: assessing

the performance of community phylogenetic structure tests. Ecology Letters, 12(9), 949–960.

https://doi.org/10.1111/j.1461-0248.2009.01354.x

Ketmaier, V., & Pavesi, L. (2013). Patterns of genetic structuring and levels of differentiation in

supralittoral talitrid amphipods: an overview. Crustaceana, 86(7–8), 890–907.

https://doi.org/10.1163/15685403-00003212

King, R. A., & Leys, R. (2011). The Australian freshwater amphipods Austrochiltonia australis and

Austrochiltonia subtenuis (Amphipoda : Talitroidea : Chiltoniidae) confirmed and two new

cryptic Tasmanian species revealed using a combined molecular and morphological approach.

Invertebrate Systematics, 25(3), 171–196. https://doi.org/10.1071/is10035

Knowles, L. L., & Carstens, B. C. (2007). Delimiting species without monophyletic gene trees.

Systematic Biology, 56(6), 887–95. https://doi.org/10.1080/10635150701701091

Knowlton, N. (1993). Sibling Species in the Sea. Annual Review of Ecology and Systematics, 24(1993),

189–216.

Knox, M. A., Hogg, I. D., Pilditch, C. A., Lörz, A.-N., Hebert, P. D. N., & Steinke, D. (2012).

Mitochondrial DNA (COI) analyses reveal that amphipod diversity is associated with

environmental heterogeneity in deep-sea habitats. Molecular Ecology, 21(19), 4885–4897.

https://doi.org/10.1111/j.1365-294X.2012.05729.x

Kornobis, E., & Pálsson, S. (2011). Discordance in Variation of the ITS Region and the Mitochondrial

This article is protected by copyright. All rights reserved.


COI Gene in the Subterranean Amphipod Crangonyx islandicus. Journal of Molecular Evolution,

73(1–2), 34–44. https://doi.org/10.1007/s00239-011-9455-2


Accepted Article
Kornobis, E., & Pálsson, S. (2013). The ITS region of groundwater amphipods: length, secondary

structure and phylogenetic information content in Crangonyctoids and Niphargids. Journal of

Zoological Systematics and Evolutionary Research, 51(1), 19–28.

https://doi.org/10.1111/jzs.12006

Kozak, K. H., & Wiens, J. J. (2006). Does niche conservatism promote speciation? A case study in

North American salamanders. Evolution, 60(12), 2604–2621. https://doi.org/10.1111/j.0014-

3820.2006.tb01893.x

Kozak, K. H., & Wiens, J. J. (2016). Testing the Relationships between Diversification, Species

Richness, and Trait Evolution. Systematic Biology, 65(6), 975–988.

https://doi.org/10.1093/sysbio/syw029

Krebes, L., Blank, M., & Bastrop, R. (2011). Phylogeography , historical demography and postglacial

colonization routes of two amphi-Atlantic distributed amphipods Phylogeography , historical

demography and postglacial colonization routes of two amphi-Atlantic distribute. Systematic

and Biodiversity, 9(3), 259–273. https://doi.org/10.1080/14772000.2011.604359

Krebes, L., Blank, M., Jürss, K., Zettler, M. L., & Bastrop, R. (2010). Molecular Phylogenetics and

Evolution Glacial-driven vicariance in the amphipod Gammarus duebeni. Molecular

Phylogenetics and Evolution, 54(2), 372–385. https://doi.org/10.1016/j.ympev.2009.07.034

Lajus, D., Sukhikh, N., & Alekseev, V. (2015). Cryptic or pseudocryptic: can morphological methods

inform copepod taxonomy? An analysis of publications and a case study of the Eurytemora

affinis species complex. Ecology and Evolution, 5(12), 2374–2385.

https://doi.org/10.1002/ece3.1521

This article is protected by copyright. All rights reserved.


Leache, A. D., Fujita, M. K., Minin, V. N., & Bouckaert, R. R. (2014). Species delimitation using

genome-wide SNP Data. Systematic Biology, 63(4), 534–542.


Accepted Article
https://doi.org/10.1093/sysbio/syu018

Lefébure, T., Douady, C. J., Gouy, M., & Gibert, J. (2006). Relationship between morphological

taxonomy and molecular divergence within Crustacea: Proposal of a molecular threshold to

help species delimitation. Molecular Phylogenetics and Evolution, 40, 435–447.

https://doi.org/10.1016/j.ympev.2006.03.014

Lortie, C. J., & Callaway, R. M. (2006). Re-analysis of meta-analysis: Support for the stress-gradient

hypothesis. Journal of Ecology, 94(1), 7–16. https://doi.org/10.1111/j.1365-2745.2005.01066.x

Lörz, A.-N., & Held, C. (2004). A preliminary molecular and morphological phylogeny of the Antarctic

Epimeriidae and Iphimediidae ( Crustacea , Amphipoda ). Molecular Phylogenetics and

Evolution, 31, 4–15. https://doi.org/10.1016/j.ympev.2003.07.019

Lörz, A.-N., Maas, E. W., Linse, K., & Coleman, C. O. (2009). Do circum-Antarctic species exist in

peracarid Amphipoda ? A case study in the genus Epimeria Costa , 1851 ( Crustacea , Peracarida

, Epimeriidae). Zookeys, 18, 91–128. https://doi.org/10.3897/zookeys.18.103

Losos, J. B. (2008). Phylogenetic niche conservatism, phylogenetic signal and the relationship

between phylogenetic relatedness and ecological similarity among species. Ecology Letters,

11(10), 995–1003. https://doi.org/10.1111/j.1461-0248.2008.01229.x

Losos, J. B. (2009). Lizards in an Evolutionary Tree: Ecology and Adaptive Radiation of Anoles.

Berkeley: University of California Press.

Losos, J. B. (2011). Convergence, Adaptation, and Constraint. Evolution, 65(7), 1827–1840.

https://doi.org/10.1111/j.1558-5646.2011.01289.x

Macneil, C., Dick, J. T. A., & Elwood, R. W. (1997). The trophic ecology of freshwater Gammarus spp.

This article is protected by copyright. All rights reserved.


(Crustacea: Amphipoda): Problems and perspectives concerning the functional feeding group

concept. Biological Reviews, 72(3), 349–364. https://doi.org/10.1111/j.1469-


Accepted Article
185X.1997.tb00017.x

Major, K., Soucek, D. J., Giordano, R., Wetzel, M. J., & Soto-Adames, F. (2013). The common

ecotoxicology laboratory strain of Hyalella azteca is genetically distinct from most wild strains

sampled in eastern North America. Environmental Toxicology and Chemistry, 32(11), 2637–

2647. https://doi.org/10.1002/etc.2355

Mamos, T., Wattier, R., Burzyński, A., & Grabowski, M. (2016). The legacy of a vanished sea: a high

level of diversification within a European freshwater amphipod species complex driven by 15

My of Paratethys regression. Molecular Ecology, 25, 795–810.

https://doi.org/10.1111/mec.13499

Marmonier, P., Maazouzi, C., Foulquier, A., Navel, S., François, C., Hervant, F., … Piscart, C. (2013).

The use of crustaceans as sentinel organisms to evaluate groundwater ecological quality.

Ecological Engineering, 57, 118–132. https://doi.org/10.1016/j.ecoleng.2013.04.009

Matzke, N. J. (2013). Probabilistic historical biogeography: new models for founder-event speciation,

imperfect detection, and fossils allow improved accuracy and model-testing. Frontiers in

Biogeography, 5(4), 242–248. https://doi.org/10.5811/westjem.2011.5.6700

Mayden, R. L. (1997). A hierarchy of species concepts: the denouement in the saga of the species

problem. In M. F. Claridge, H. A. Dawah, & M. R. Wilson (Eds.), Species. The units of biodiversity.

(pp. 381–423). Chapman and Hall.

Mayr, E. (1942). Systematics and the origin of species from the viewpoint of zoologist. Cambridge,

Massachusetts, London: Harvard University Press.

McInerney, C. E., Maurice, L., Robertson, A. L., Knight, L. R. F. D., Arnscheidt, J., Venditti, C., …

This article is protected by copyright. All rights reserved.


Hänfling, B. (2014). The ancient Britons: groundwater fauna survived extreme climate change

over tens of millions of years across NW Europe. Molecular Ecology, 23, 1153–1166.
Accepted Article
https://doi.org/10.1111/mec.12664

Meleg, I. N., Zakšek, V., Fišer, C., Kelemen, B. S., & Moldovan, O. T. (2013). Can Environment Predict

Cryptic Diversity? The Case of Niphargus Inhabiting Western Carpathian Groundwater. PLoS

ONE, 8(10), e76760. https://doi.org/10.1371/journal.pone.0076760

Monaghan, M. T., Wild, R., Elliot, M., Fujisawa, T., Balke, M., Inward, D. J. G., … Vogler, A. P. (2009).

Accelerated species Inventory on Madagascar using coalescent-based models of species

Delineation. Systematic Biology, 58(3), 298–311. https://doi.org/10.1093/sysbio/syp027

Montero-Pau, J., Ramos-Rodríguez, E., Serra, M., & Gómez, A. (2011). Long-term coexistence of

rotifer cryptic species. PloS One, 6(6), e21530. https://doi.org/10.1371/journal.pone.0021530

Montero-Pau, J., & Serra, M. (2011). Life-cycle switching and coexistence of species with no niche

differentiation. PloS One, 6(5), e20314. https://doi.org/10.1371/journal.pone.0020314

Morard, R., Escarguel, G., Weiner, A. K. ., André, A., Douady, C. J., Wade, C. M., … Kucera, M. (2016).

Nomenclature for the Nameless: a Proposal for an Integrative Molecular Taxonomy of Cryptic

Diversity Exemplified by Planktonic Foraminifera. Systematic Biology, 65(5), 925–940.

Morinière, J., Hendrich, L., Balke, M., Beermann, A. J., König, T., Hess, M., … Haszprunar, G. (2017). A

DNA barcode library for Germany′s mayflies, stoneflies and caddisflies (Ephemeroptera,

Plecoptera and Trichoptera). Molecular Ecology Resources, 17(6), 1293–1307.

https://doi.org/10.1111/1755-0998.12683

Morvan, C., Malard, F., Paradis, E., Lefébure, T., Konecny-Dupré, L., & Douady, C. J. (2013). Timetree

of Aselloidea reveals species diversification dynamics in groundwater. Systematic Biology,

62(4), 512–522. https://doi.org/10.1093/sysbio/syt015

This article is protected by copyright. All rights reserved.


Murphy, N. P., Adams, M., Guzik, M. T., & Austin, A. D. (2013). Molecular phylogenetics and

evolution extraordinary micro-endemism in Australian desert spring amphipods. Molecular


Accepted Article
Phylogenetics and Evolution, 66(3), 645–653. https://doi.org/10.1016/j.ympev.2012.10.013

Murphy, N. P., King, R. A., & Delean, S. (2015). Species, ESUs or populations? Delimiting and

describing morphologically cryptic diversity in Australian desert spring amphipods. Invertebrate

Systematics, 29(5), 457–467. https://doi.org/10.1071/IS14036

Myers, A. A., & Lowry, K. (2009). The biogeography of Indo-West Pacific tropical amphipods with

particular reference to Australia. Zootaxa, 127(2260), 109–127.

Nahavandi, N., Ketmaier, V., & Tiedemann, R. (2012). Intron structure of the Elongation Factor 1-

Alpha gene in the Ponto-Caspian amphipod Pontogammarus maeoticus (Sowinsky, 1894) and

its phylogeographic utility. Journal of Crustacean Biology, 32(3), 425–433.

https://doi.org/10.1163/193724012X626584

Nicholls, B., & Racey, P. A. (2006). Habitat Selection as a Mechanism of Resource Partitioning in Two

Cryptic Bat Species Pipistrellus pipistrellus and Pipistrellus pygmaeus. Ecography, 29(5), 697–

708.

Nosil, P. (2012). Ecological Speciation and its alternatives. Ecological Speciation, 1–21.

Nygren, A. (2014). Cryptic polychaete diversity: a review. Zoologica Scripta, 43(2), 172–183.

https://doi.org/10.1111/zsc.12044

O’Meara, B. C. (2010). New Heuristic Methods for Joint Species Delimitation and Species Tree

Inference. Systematic Biology, 59(1), 59–73. https://doi.org/10.1093/sysbio/syp077

O’Meara, B. C., Ané, C., Sanderson, M. J., & Wainwright, P. C. (2006). Testing for different rates of

continuous trait evolution using likelihood. Evolution; International Journal of Organic

Evolution, 60(5), 922–933. https://doi.org/10.1111/j.0014-3820.2006.tb01171.x

This article is protected by copyright. All rights reserved.


Ortells, R., Gomez, A., & Serra, M. (2003). Coexistence of cryptic rotifer species: ecological and

genetic characteristics of Brachionus plicatilis. Freshwater Biology, 48, 2194–2202.


Accepted Article
https://doi.org/10.1046/j.1365-2427.2003.01159.x

Padial, J. M., Miralles, A., De la Riva, I., & Vences, M. (2010). The integrative future of taxonomy.

Frontiers in Zoology, 7(1), 16. https://doi.org/10.1186/1742-9994-7-16

Pante, E., Schoelinck, C., & Puillandre, N. (2015). From integrative taxonomy to species description:

one step beyond. Systematic Biology, 64(1), 152–160. https://doi.org/10.1093/sysbio/syu083

Pérez-Ponce de León, G., & Nadler, S. a. (2010). What We Don’t Recognize Can Hurt Us: A Plea for

Awareness About Cryptic Species. Journal of Parasitology, 96(2), 453–464.

https://doi.org/10.1645/GE-2260.1

Pérez-Ponce de León, G., & Poulin, R. (2016). Taxonomic distribution of cryptic diversity among

metazoans: not so homogeneous after all. Biology Letters, 12(8), 20160371.

https://doi.org/10.1098/rsbl.2016.0371

Pfenninger, M., & Nowak, C. (2008). Reproductive Isolation and Ecological Niche Partition among

Larvae of the Morphologically Cryptic Sister Species Chironomus riparius and C. piger. PLoS

ONE, 3(5), e2157. https://doi.org/10.1371/journal.pone.0002157

Pfenninger, M., & Schwenk, K. (2007). Cryptic animal species are homogeneously distributed among

taxa and biogeographical regions. BMC Evolutionary Biology, 7, 121.

https://doi.org/10.1186/1471-2148-7-121

Pilgrim, E. M., Blum, M. J., Reusser, D. A., Lee, H. I., & Darling, J. A. (2013). Geographic range and

structure of cryptic genetic diversity among Pacific North American populations of the non-

native amphipod Grandidierella japonica. Biological Invasions, 15, 2415–2428.

https://doi.org/10.1007/s10530-013-0462-7

This article is protected by copyright. All rights reserved.


Pilgrim, E. M., & Darling, J. A. (2010). Genetic diversity in two introduced biofouling amphipods

(Ampithoe valida & Jassa marmorata) along the Pacific North American coast: investigation into
Accepted Article
molecular identification and cryptic diversity. Diversity and Distributions, 16, 827–839.

https://doi.org/10.1111/j.1472-4642.2010.00681.x

Pinchuk, A. I., Coyle, K. O., Farley, E. V., & Renner, H. M. (2013). Emergence of the Arctic Themisto

libellula (Amphipoda: Hyperiidae) on the southeastern Bering Sea shelf as a result of the recent

cooling, and its potential impact on the pelagic food web. ICES Journal of Marine Science, 70(6),

1244–1254. https://doi.org/10.1093/icesjms/fsv214

Poisot, T., Canard, E., Mouillot, D., Mouquet, N., & Gravel, D. (2012). The dissimilarity of species

interaction networks. Ecology Letters, 15(12), 1353–1361. https://doi.org/10.1111/ele.12002

Pons, J., Barraclough, T., Gomez-Zurita, J., Cardoso, A., Duran, D., Hazell, S., … Vogler, A. (2006).

Sequence-Based Species Delimitation for the DNA Taxonomy of Undescribed Insects.

Systematic Biology, 55(4), 595–609. https://doi.org/10.1080/10635150600852011

Puillandre, N., Lambert, A., Brouillet, S., & Achaz, G. (2012). ABGD, Automatic Barcode Gap Discovery

for primary species delimitation. Molecular Ecology, 21, 1864–1877.

https://doi.org/10.1111/j.1365-294X.2011.05239.x

Pyron, R. A., & Burbrink, F. T. (2013). Phylogenetic estimates of speciation and extinction rates for

testing ecological and evolutionary hypotheses. Trends in Ecology & Evolution, 28(12), 729–36.

https://doi.org/10.1016/j.tree.2013.09.007

Pyron, R. A., Costa, G. C., Patten, M. A., & Burbrink, F. T. (2015). Phylogenetic niche conservatism and

the evolutionary basis of ecological speciation. Biological Reviews, 90(4), 1248–1262.

https://doi.org/10.1111/brv.12154

Rannala, B., & Yang, Z. (2003). Bayes Estimation of Species Divergence Times and Ancestral

This article is protected by copyright. All rights reserved.


Population Sizes Using DNA Sequences From Multiple Loci. Genetics, 164(4), 1645–1656.

Retrieved from http://www.genetics.org/content/164/4/1645.abstract


Accepted Article
Ratnasingham, S., & Hebert, P. D. N. (2007). BARCODING, BOLD : The Barcode of Life Data System

(www.barcodinglife.org). Molecular Ecology Notes, 7(April 2016), 355–364.

https://doi.org/10.1111/j.1471-8286.2006.01678.x

Raupach, M. J., Amann, R., Wheeler, Q. D., & Roos, C. (2016). The application of “-omics”

technologies for the classification and identification of animals. Organisms Diversity &

Evolution, 16(1), 1–12. https://doi.org/10.1007/s13127-015-0234-6

Raxworthy, C., Ingram, C., Rabibisoa, N., & Pearson, R. (2007). Applications of Ecological Niche

Modeling for Species Delimitation: A Review and Empirical Evaluation Using Day Geckos

(Phelsuma) from Madagascar. Systematic Biology, 56(6), 907–923.

https://doi.org/10.1080/10635150701775111

Ree, R. H., & Smith, S. A. (2008). Maximum Likelihood Inference of Geographic Range Evolution by

Dispersal, Local Extinction, and Cladogenesis. Systematic Biology, 57(1), 4–14.

https://doi.org/10.1080/10635150701883881

Reid, N. M., & Carstens, B. C. (2012). Phylogenetic estimation error can decrease the accuracy of

species delimitation: a Bayesian implementation of the general mixed Yule-coalescent model.

BMC Evolutionary Biology, 12(1), 1–11. https://doi.org/10.1186/1471-2148-12-196

Renner, S. S. (2016). A return to Linnaeus’s focus on diagnosis, not description: The use of DNA

characters in the formal naming of species. Systematic Biology, 65(6), 1086–1095.

https://doi.org/10.1093/sysbio/syw032

Rewicz, T., Wattier, R. a., Rigaud, T., Bacela-Spychalska, K., & Grabowski, M. (2015). Isolation and

characterization of 8 microsatellite loci for the “killer shrimp’’, an invasive Ponto-Caspian

This article is protected by copyright. All rights reserved.


amphipod Dikerogammarus villosus (Crustacea: Amphipoda). Molecular Biology Reports, 42(1),

13–17. https://doi.org/10.1007/s11033-014-3742-0
Accepted Article
Richards, V. P., Stanhope, M. J., & Shivji, M. S. (2012). Island endemism, morphological stasis, and

possible cryptic speciation in two coral reef, commensal Leucothoid amphipod species

throughout Florida and the Caribbean. Biodiversity and Conservation, 21, 343–361.

https://doi.org/10.1007/s10531-011-0186-x

Rock, J., Ironside, J., Potter, T., Whiteley, N. M., & Lunt, D. H. (2007). Phylogeography and

environmental diversification of a highly adaptable marine amphipod , Gammarus duebeni.

Heredity, 99, 102–111. https://doi.org/10.1038/sj.hdy.6800971

Roux, C., Fraïsse, C., Romiguier, J., Anciaux, Y., Galtier, N., & Bierne, N. (2016). Shedding Light on the

Grey Zone of Speciation along a Continuum of Genomic Divergence. PLoS Biology, 14(12), 1–22.

https://doi.org/10.1371/journal.pbio.2000234

Rundle, H. D., & Nosil, P. (2005). Ecological speciation. Ecology Letters.

https://doi.org/10.1111/j.1461-0248.2004.00715.x

Ruokolainen, L., & Hanski, I. (2016). Stable coexistence of ecologically identical species: conspecific

aggregation via reproductive interference. Journal of Animal Ecology, n/a-n/a.

https://doi.org/10.1111/1365-2656.12490

Rutishauser, M. D., Bontadina, F., Braunisch, V., Ashrafi, S., & Arlettaz, R. (2012). The challenge

posed by newly discovered cryptic species: disentangling the environmental niches of long-

eared bats. Diversity and Distributions, 18(11), 1107–1119. https://doi.org/10.1111/j.1472-

4642.2012.00904.x

Ryberg, M. (2015). Molecular operational taxonomic units as approximations of species in the light

of evolutionary models and empirical data from Fungi. Molecular Ecology, 24(23), 5770–5777.

This article is protected by copyright. All rights reserved.


https://doi.org/10.1111/mec.13444

Sáez, A. G., & Lozano, E. (2005). Body doubles. Nature, 433(7022), 111.
Accepted Article
https://doi.org/10.1038/433111a

Safi, K., Cianciaruso, M. V., Loyola, R. D., Brito, D., Armour-Marshall, K., & Diniz-Filho, J. A. F. (2011).

Understanding global patterns of mammalian functional and phylogenetic diversity.

Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences,

366(1577), 2536–2544. https://doi.org/10.1098/rstb.2011.0024

Saleh, D., Laarif, A., Clouet, C., & Gauthier, N. (2012). Spatial and host-plant partitioning between

coexisting Bemisia tabaci cryptic species in Tunisia. Population Ecology, 54(2), 261–274.

https://doi.org/10.1007/s10144-012-0303-z

Scheffer, M., Vergnon, R., van Nes, E. H., Cuppen, J. G. M., Peeters, E. T. H. M., Leijs, R., & Nilsson, A.

N. (2015). The Evolution of Functionally Redundant Species; Evidence from Beetles. Plos One,

10(10), e0137974. https://doi.org/10.1371/journal.pone.0137974

Schlick-Steiner, B. C., Steiner, F. M., Seifert, B., Stauffer, C., Christian, E., & Crozier, R. H. (2010).

Integrative Taxonomy: A Multisource Approach to Exploring Biodiversity. Annual Review of

Entomology, 55, 421–438. https://doi.org/10.1146/annurev-ento-112408-085432

Schluter, D. (2000a). Ecological Character Displacement in Adaptive Radiation. The American

Naturalist, 156, S4–S16. https://doi.org/10.1086/303412

Schluter, D. (2000b). The Ecology of Adaptive Radiation. Oxford University Press.

Schluter, D. (2016). Speciation, ecological opportunity, and latitude. The American Naturalist, 187(1),

1–18. https://doi.org/10.1086/684193

Scriven, J. J., Whitehorn, P. R., Goulson, D., & Tinsley, M. C. (2016). Niche partitioning in a sympatric

cryptic species complex. Ecology and Evolution, n/a-n/a. https://doi.org/10.1002/ece3.1965

This article is protected by copyright. All rights reserved.


Siepielski, A. M., & McPeek, M. A. (2010). On the evidence for species coexistence: A critique of the

coexistence program. Ecology, 91(11), 3153–3164. https://doi.org/10.1890/10-0154.1


Accepted Article
Sinclair, B. J., Marshall, K. E., Sewell, M. A., Levesque, D. L., Willett, C. S., Slotsbo, S., … Huey, R. B.

(2016). Can we predict ectotherm responses to climate change using thermal performance

curves and body temperatures? Ecology Letters, 19(11), 1372–1385.

https://doi.org/10.1111/ele.12686

Sites, J. W., & Marshall, J. C. (2003). Delimiting species: a Renaissance issue in systematic biology.

Trends in Ecology & Evolution, 18(9), 462–470. https://doi.org/10.1016/S0169-5347(03)00184-

Sites, J. W., & Marshall, J. C. (2004). Operational Criteria for Delimiting Species. Annual Review of

Ecology, Evolution, and Systematics, 35(1), 199–227.

https://doi.org/10.1146/annurev.ecolsys.35.112202.130128

Smith, K. L., Harmon, L. J., Shoo, L. P., & Melville, J. (2011). Evidence of Constrained Phenotypic

Evolution in a Cryptic Species Complex of Agamid Lizards. Evolution, 65(4), 976–992.

https://doi.org/10.1111/j.1558-5646.2010.01211.x

Soucek, D. J., Dickinson, A., Major, K. M., & McEwen, A. R. (2013). Effect of test duration and feeding

on relative sensitivity of genetically distinct clades of Hyalella azteca. Ecotoxicology, 22(9),

1359–66. https://doi.org/10.1007/s10646-013-1122-5

Stevens, G. C. (1989). The Latitudinal Gradient in Geographical Range : How so Many Species Coexist

in the Tropics. The American Naturalist, 133(2), 240–256.

Strand, M., & Sundberg, P. (2011). A DNA-based description of a new nemertean (phylum Nemertea)

species. Marine Biology Research, 7(1), 63–70. https://doi.org/Pii 930665099\rDoi

10.1080/17451001003713563

This article is protected by copyright. All rights reserved.


Stroud, J. T., & Losos, J. B. (2016). Ecological Opportunity and Adaptive Radiation. Annual Review of

Ecology and Systematics, (47), 507–532. https://doi.org/10.1146/annurev-ecolsys-121415-


Accepted Article
032254

Stuart, Y. E., & Losos, J. B. (2013). Ecological character displacement: glass half full or half empty?

Trends in Ecology & Evolution, 28, 402–408. https://doi.org/10.1016/j.tree.2013.02.014

Švara, V., Delić, T., Rađa, T., & Fišer, C. (2015). Molecular phylogeny of Niphargus boskovici

(Crustacea: Amphipoda) reveals a new species from epikarst. Zootaxa, 3994(3), 354–376.

Trontelj, P., Blejec, A., & Fišer, C. (2012). Ecomorphological convergence of cave communities.

Evolution; International Journal of Organic Evolution, 66(12), 3852–65.

https://doi.org/10.1111/j.1558-5646.2012.01734.x

Trontelj, P., Douady, C. J., Fišer, C., Gibert, J., Gorički, Š., Lefébure, T., … Zakšek, V. (2009). A

molecular test for cryptic diversity in ground water: how large are the ranges of macro-

stygobionts? Freshwater Biology, 54(4), 727–744. https://doi.org/10.1111/j.1365-

2427.2007.01877.x

Trontelj, P., & Fišer, C. (2009). Perspectives: Cryptic species diversity should not be trivialised.

Systematics and Biodiversity, 7(1), 1–3. https://doi.org/10.1017/S1477200008002909

Väinölä, R., Witt, J. D. S., Grabowski, M., Bradbury, J. H., Jazdzewski, K., & Sket, B. (2008). Global

diversity of amphipods (Amphipoda; Crustacea) in freshwater. Hydrobiologia, 595, 241–255.

https://doi.org/10.1007/s10750-007-9020-6

Venarsky, M. P., Anderson, F. E., & Wilhelm, F. M. (2009). Population genetic study of the U.S.

federally listed Illinois cave amphipod, Gammarus acherondytes. Conservation Genetics, 10(4),

915–921. https://doi.org/10.1007/s10592-008-9579-0

Verheye, M. L., Backeljau, T., & d’Udekem d’Acoz, C. (2016). Looking beneath the tip of the iceberg:

This article is protected by copyright. All rights reserved.


diversification of the genus Epimeria on the Antarctic shelf (Crustacea, Amphipoda). Polar

Biology, 39(5), 925–945. https://doi.org/10.1007/s00300-016-1910-5


Accepted Article
Villacorta, C., Canovas, F., Oromi, P., & Juan, C. (2009). Isolation and characterization of

microsatellite loci for the cavehopper Palmorchestia hypogaea ( Amphipoda : Talitridae ).

Conservation Genetics Resources, 1, 401–404. https://doi.org/10.1007/s12686-009-9093-9

Villacorta, C., Jaume, D., Oromí, P., & Juan, C. (2008). Under the volcano: phylogeography and

evolution of the cave-dwelling Palmorchestia hypogaea (Amphipoda, Crustacea) at La Palma

(Canary Islands). BMC Biology, 6(1), 7. https://doi.org/10.1186/1741-7007-6-7

Voda, R., Dapporto, L., Dinca, V., & Vila, R. (2015). Cryptic matters: Overlooked species generate

most butterfly beta-diversity. Ecography, 38(4), 405–409. https://doi.org/10.1111/ecog.00762

Warren, D. L., Glor, R. E., & Turelli, M. (2008). Environmental Niche Equivalency Versus

Conservatism: Quantitative Approaches To Niche Evolution. Evolution, 62(11), 2868–2883.

https://doi.org/10.1111/j.1558-5646.2008.00482.x

Weiss, M., & Leese, F. (2016). Widely distributed and regionally isolated! Drivers of genetic structure

in Gammarus fossarum in a human-impacted landscape. BMC Evolutionary Biology, 16(1), 153.

https://doi.org/10.1186/s12862-016-0723-z

Weiss, M., Niklas, J., Anna, M., & Florian, S. (2014). Molecular evidence for further overlooked

species within the Gammarus fossarum complex ( Crustacea : Amphipoda ). Hydrobiologia, 721,

165–184. https://doi.org/10.1007/s10750-013-1658-7

Wellborn, G. A., & Cothran, R. D. (2004). Phenotypic similarity and differntiation among sympatric

cryptic species in a freshwater amphipod species complex. Freshwater Biology, 49, 1–13.

Wellborn, G. A., & Cothran, R. D. (2007). Niche diversity in crustacean cryptic species:

complementarity in spatial distribution and predation risk. Oecologia, 154(1), 175–183.

This article is protected by copyright. All rights reserved.


https://doi.org/10.1007/s00442-007-0816-x

Weston, D. P., Poynton, H. C., Wellborn, G. A., Lydy, M. J., Blalock, B. J., Sepulveda, M. S., &
Accepted Article
Colbourne, J. K. (2013). Multiple origins of pyrethroid insecticide resistance across the species

complex of a nontarget aquatic crustacean, Hyalella azteca. Proceedings of the National

Academy of Sciences, 110(41), 16532–16537. https://doi.org/10.1073/pnas.1302023110

Westram, A., Baumgartner, C., Keller, I., & Jokela, J. (2011). Are cryptic host species also cryptic to

parasites? Host specificity and geographical distribution of acanthocephalan parasites infecting

freshwater Gammarus. Infection, Genetics and Evolution, 11(5), 1083–1090.

https://doi.org/10.1016/j.meegid.2011.03.024

White, K. N., Reimer, J. D., & Lorion, J. (2016). Preliminary analyses reveal strong genetic structure in

populations of Leucothoe vulgaris (Crustacea: Amphipoda: Leucothoidae) from Okinawa, Japan.

Systematics and Biodiversity, 14(1), 55–62. https://doi.org/10.1080/14772000.2015.1078856

Wiens, J. J. (2004). Speciation and ecology revisited: phylogenetic niche conservatism and the origin

of species. Evolution; International Journal of Organic Evolution, 58(1), 193–197.

https://doi.org/10.1016/j.echo.2007.08.009

Wiens, J. J. (2008). Commentary on Losos (2008): niche conservatism déjà vu. Ecology Letters,

11(10), 1004–1005. https://doi.org/10.1111/j.1461-0248.2008.01238.x

Wiens, J. J. (2011). The causes of species richness patterns across space, time, and clades and the

role of “ecological limits”. The Quarterly Review of Biology, 86(2), 75–96.

https://doi.org/10.1086/659883

Wiens, J. J., Ackerly, D. D., Allen, A. P., Anacker, B. L., Buckley, L. B., Cornell, H. V., … Stephens, P. R.

(2010). Niche conservatism as an emerging principle in ecology and conservation biology.

Ecology Letters, 13(10), 1310–1324. https://doi.org/10.1111/j.1461-0248.2010.01515.x

This article is protected by copyright. All rights reserved.


Wiens, J. J., & Graham, C. H. (2005). Niche Conservatism: Integrating Evolution, Ecology, and

Conservation Biology. Annual Review of Ecology, Evolution, and Systematics, 36(1), 519–539.
Accepted Article
https://doi.org/10.1146/annurev.ecolsys.36.102803.095431

Wiens, J. J., & Penkrot, T. A. (2002). Delimiting species using DNA and morphological variation and

discordant species limits in spiny lizards (Sceloporus). Systematic Biology, 51(1), 69–91.

https://doi.org/10.1080/106351502753475880

Wiens, J. J., & Servedio, M. R. (2000). Species delimitation in systematics: inferring diagnostic

differences between species. Proceedings of the Royal Society of London - Series B: Biological

Sciences, 267(1444), 631–636. https://doi.org/10.1098/rspb.2000.1049

Witt, J. D. S., Threloff, D. L., & Hebert, P. D. N. (2006). DNA barcoding reveals extraordinary cryptic

diversity in an amphipod genus : implications for desert spring conservation. Molecular

Ecology, 15, 3073–3082. https://doi.org/10.1111/j.1365-294X.2006.02999.x

Yang, L., Hou, Z., & Li, S. (2013). Marine incursion into East Asia : a forgotten driving force of

biodiversity Marine incursion into East Asia : a forgotten driving force of biodiversity.

Proceedings of the Royal Society B: Biological Sciences, 280, 1–8.

Yang, Z., & Rannala, B. (2010). Bayesian species delimitation using multilocus sequence data.

Proceedings of the National Academy of Sciences, 107(20), 9264–9269.

https://doi.org/10.1073/pnas.0913022107

Yang, Z., & Rannala, B. (2014). Unguided species delimitation using DNA sequence data from

multiple loci. Molecular Biology and Evolution, 31(12), 3125–3135.

https://doi.org/10.1093/molbev/msu279

Yeates, D. K., Seago, A., Nelson, L., Cameron, S. L., Joseph, L., & Trueman, J. W. H. (2011). Integrative

taxonomy , or iterative taxonomy ? Systematic Entomology, 36, 209–217.

This article is protected by copyright. All rights reserved.


https://doi.org/10.1111/j.1365-3113.2010.00558.x

Zettler, M. L., Proffitt, C. E., Darr, A., Degraer, S., Devriese, L., Greathead, C., … Ysebaert, T. (2013).
Accepted Article
On the myths of indicator species: issues and further consideration in the use of static concepts

for ecological applications. PLoS ONE, 8(10), e78219.

https://doi.org/10.1371/journal.pone.0078219

Zhang, D. Y., Lin, K., & Hanski, I. (2004). Coexistence of cryptic species. Ecology Letters, 7(3), 165–

169. https://doi.org/10.1111/j.1461-0248.2004.00569.x

Zhang, J., Kapli, P., Pavlidis, P., & Stamatakis, A. (2013). A general species delimitation method with

applications to phylogenetic placements. Bioinformatics, 29(22), 2869–76.

https://doi.org/10.1093/bioinformatics/btt499

Zúñiga-Reinoso, Á., & Benítez, H. A. (2015). The overrated use of the morphological cryptic species

concept: An example with Nyctelia darkbeetles (Coleoptera: Tenebrionidae) using geometric

morphometrics. Zoologischer Anzeiger, 255, 47–53. https://doi.org/10.1016/j.jcz.2015.01.004

This article is protected by copyright. All rights reserved.


Table 1. Classification of methods used in DNA taxonomy to delimit cryptic species of amphipods

Methods Delimitation criteria Examples of analyses / References


Accepted Article
programs

Distance-based Intraspecific genetic Fixed molecular threshold, Lefébure et al., 2006;

methods variation << interspecific automated barcoding gap Puillandre et al., 2012

variation detection

Tree-based Point of transition from GMYC, PTP, BP&P, BFD, Pons et al., 2006; O’Meara,

methods species-level to BFD*, SpedeSTEM, Brownie 2010; Ence & Carstens, 2011;

population-level Grummer, Bryson & Reeder,

processes 2014; Yang & Rannala, 2010

Allele-sharing Mutual allelic exclusivity Haplowebs Flot et al., 2010

based

Population Reduced gene flow Nested clade analysis, Delimit differentiated

genetics among populations differentiation statistics populations rather than

species

GMYC: Generalized mixed Yule-coalescent models; PTP: Poisson tree processes models, BP&P:

Bayesian phylogenetics & phylogeography; BFD: Bayes factor delimitation (BFD* for genomic-scale

data).

This article is protected by copyright. All rights reserved.


BOX 1 Amphipods as a model for study of cryptic species

Amphipods (Crustacea: Malacostraca: Peracarida) consist of ca. 10,000 species that are classified
Accepted Article
into six orders (Horton et al., 2016; Väinölä et al., 2008). There is no comprehensive molecular

phylogeny of amphipods and clade-specific phylogenetic studies indicate that morphology alone

neither accurately depicts species diversity nor phylogenetic relationships (Corrigan, Horton,

Fotherby, White, & Hoelzel, 2013; C. Fišer, Sket, & Trontelj, 2008; Hou et al., 2011; Lörz & Held,

2004; Verheye, Backeljau, & d’Udekem d’Acoz, 2016). Species-level taxonomy is demanding.

Amphipods do not have copulatory apparatus and the species were traditionally delimited on a basis

of different appendage shapes, body shapes and setal patterns; reproductive barriers were rarely if

ever studied. Within-population variation is often high in contrast to small between-species

differences.

Amphipods thrive in most aquatic habitats and some even inhabit terrestrial habitats (Ketmaier &

Pavesi, 2013) (Fig. 2). Many species (80%) are marine, living in the supra-littoral zone to the hadal

zone, where they represent an important part of the community (Jamieson, Fujii, Mayor, Solan, &

Priede, 2010). About 45% of freshwater species are exclusive to groundwater (Väinölä et al., 2008), a

highly constrained environment that has long been thought to promote morphological stasis and

convergence (Lefébure et al., 2006). Most amphipods are benthic, however, and one order

(Hyperiida) is pelagic. Benthic species display highly diverse ecologies (e.g., free-roaming species,

burrowers, tube builders, sedentary ambush-predators, and commensals on whales, mollusks and

bryozoans) and use a broad range of feeding strategies (e.g., filter feeding, shredding, scavenging

and predation) (Alexander, Dick, & O’Connor, 2013; Macneil, Dick, & Elwood, 1997). Local

abundances can be high; hence they are also a key link in food webs (Pinchuk, Coyle, Farley, &

Renner, 2013). Lastly, amphipods are routinely used in ecotoxicological tests (Feckler et al., 2012;

Marmonier et al., 2013), the results of which may be biased if model species contain cryptic species

with different sensitivities to pollutants.

This article is protected by copyright. All rights reserved.


In the absence of free pelagic larvae, most amphipod species have limited distributional ranges

(Myers & Lowry, 2009). Yet, some tens of species, including some cryptic species (Cristescu, Witt,
Accepted Article
Grigorovich, Hebert, & MacIsaac, 2004; Pilgrim, Blum, Reusser, Lee, & Darling, 2013; Pilgrim &

Darling, 2010), have substantially expanded their distributional ranges in recent decades (Grabowski,

Bacela, & Konopacka, 2007; Jazdzewski, Konopacka, & Grabowski, 2004). Some amphipods are

invasive species and pose a threat to native fauna (Dick & Platvoet, 2000; Dodd et al., 2014).

Amphipods are an ideal model organism for linking evolutionary processes contributing to cryptic

diversity with local ecological dynamics, global biodiversity patterns and applied ecology. Cryptic

species have been found in all ecological groups of amphipods. The number of cryptic species within

nominal species varies up to several tens of species (Havermans, 2016; Katouzian et al., 2016;

Mamos et al., 2016; Witt et al., 2006). Ecological diversity of amphipods allows for testing

generalities in processes generating cryptic diversity under different ecological settings. Limited

distributional ranges at the species level coupled with ubiquitous distribution at the genus level

provide natural replicates for analysis. As a key functional group in ecosystems, amphipods naturally

link the issue of morphological evolution with local ecological dynamics, ecosystem functioning and

nature conservation.

BOX 2. Co-occurrence of cryptic species

Knowlton (1993) suggested that sympatry and local co-occurrence is common among cryptic species

in the sea. Recent studies from many animal phyla largely support this view; no less than 20% of the

cryptic amphipod species were found to live in sympatry or even syntopy.

Assuming that cryptic species significantly overlap in their ecological niche space, niche

differentiation cannot explain their coexistence (Chesson, 2000). Several studies addressed this issue

using various model organisms, including rotifers (Montero-Pau, Ramos-Rodríguez, Serra, & Gómez,

This article is protected by copyright. All rights reserved.


2011; Montero-Pau & Serra, 2011; Ortells, Gomez, & Serra, 2003), nematodes (De Meester, Derycke,

Bonte, & Moens, 2011; De Meester et al., 2016; Derycke et al., 2008), different amphipods (Cothran,
Accepted Article
Henderson, et al., 2013; Cothran, Noyes, & Relyea, 2015; Cothran, Stiff, Chapman, Wellborn, &

Relyea, 2013; Dionne, Vergilino, Dufresne, Charles, & Nozais, 2011; Eisenring et al., 2016; Ž. Fišer et

al., 2015; Wellborn & Cothran, 2004), chironomids (Pfenninger & Nowak, 2008), bugs (Saleh, Laarif,

Clouet, & Gauthier, 2012), bumble bees (Scriven, Whitehorn, Goulson, & Tinsley, 2016), fig wasps (D.

Y. Zhang, Lin, & Hanski, 2004) and bats (Ashrafi, Beck, Rutishauser, Arlettaz, & Bontadina, 2011;

Nicholls & Racey, 2006; Rutishauser, Bontadina, Braunisch, Ashrafi, & Arlettaz, 2012). These studies

unveiled emerging commonalities in co-occurrence patterns and unsolved issues with important

implications for nature conservation.

With little exception, studies showed that co-occurring cryptic species are ecologically differentiated.

The degree of species differentiation varies from case to case. One study showed strong differential

preferences for foraging habitat (Nicholls & Racey, 2006), whereas other cases found differences are

more subtle (Ž. Fišer et al., 2015; Wellborn & Cothran, 2004). Co-occurrence is based on different

trade-offs; i.e., i) when species disperse along environmental gradients and locally outcompete each

other (Eisenring et al., 2016; Ž. Fišer et al., 2015; Pfenninger & Nowak, 2008; Rutishauser et al.,

2012), ii) when temporal fluctuations in environmental factors alter competitive strength of co-

occurring species over daily, seasonal or annual scales (Montero-Pau et al., 2011; Montero-Pau &

Serra, 2011; Ortells et al., 2003), iii) where selective predation regulates the strength of the stronger

competitor (Wellborn & Cothran, 2004, 2007), and iv) when, under various scenarios, heterospecific

breeding interference favors population growth when species is rare (Ruokolainen & Hanski, 2016;

D. Y. Zhang et al., 2004).

This review brought up a general issue; i.e., whether and how statistically significant differences in

biological traits can be translated into ecologically meaningful differences. Some authors attempted

to solve the puzzle by testing whether species co-occurrences are random or non-randomly more

This article is protected by copyright. All rights reserved.


frequent/rare with respect to distributional data (Ž. Fišer et al., 2015; Pfenninger & Nowak, 2008;

Rutishauser et al., 2012; Saleh et al., 2012). The best evidence for coexistence was a demonstration
Accepted Article
that a species’ population can grow if initially present in the system at low abundance (Siepielski &

McPeek, 2010). This test requires laboratory controlled experiments that are, in the case of cryptic

species, quite demanding. We are aware of only two model systems that tested for co-existence

under laboratory conditions. In the amphipod Hyalella, none of the three co-occurring species

showed positive population growth when they were introduced at low abundance in experimental

mesocosms (Cothran et al., 2015). In contrast, interactions among species of the nematode Pelloditis

marina either facilitated or suppressed population growth (De Meester et al., 2011). An ambiguous

relationship between the degree of species differentiation and population growth in a community

has important implications for many questions; e.g., can we predict species responses to

environmental change (Sinclair et al., 2016). This question extends beyond the issue of co-occurring

cryptic species. For example, if local factors mediate long-term co-occurrence of species which

outcompete each other in mesocosms, it can be expected that altered environmental conditions

may reshape ecological space and intensify competitive interactions between initially non-

competing species (Lortie & Callaway, 2006; Poisot, Canard, Mouillot, Mouquet, & Gravel, 2012).

Glossary

Brownian motion: A model of trait evolution, sometimes called random walk, which assumes

neutral and gradual evolution of traits due to genetic drift. Change in mean phenotype is

expected to be non-directional and occurs at a constant rate, whereas variance among

species is linearly related to the amount of time since divergence.

This article is protected by copyright. All rights reserved.


Orstein-Uhlenbeck model: A model of trait evolution, sometimes also called a “rubber-band”, which

is an alternative model describing the response of a phenotype to stabilizing selection.


Accepted Article
Evolutionary changes are constrained by a constant force towards an adaptive optimum.

White noise model: A model of trait evolution representing a process in which variation of

phenotypic traits revolves constantly around moving trait optima, which generates

evolutionary phenotypic change that is independent from phylogenetic relationships.

Morphological crypsis: Extreme similarity of morphological phenotypes of two or more species. In

this review we use this term for those species whose morphological differentiation cannot

be told apart without prior molecular analysis.

Polyphyletic group: A group of organisms derived from different ancestors.

Monophyletic group: A group of organisms composed of a common ancestor and all its descendant

species.

Paraphyletic group: A group of organisms derived from a common ancestor, which excludes some

descendants.

Grinellian ecological niche. A concept of ecological niche grounded on species requirements for a

specific habitat; a niche is parametrized with conditions in which a species persists and

breeds.

Eltonian ecological niche: A concept of ecological niche defined by the place a species takes in its

biotic environment; i.e., “its relationship towards food and enemies” and the functional

role it plays in the ecosystem.

Integrative taxonomy: An interdisciplinary approach in delimiting species, based on confrontation of

genetic, morphological, behavioral and ecological evidence of lineage separation.

Species concept vs. species taxon: Several authors argued that much of the confusion related to the

term species is of semantic nature. The “species concept” treats species as evolutionary

entities. “Species taxon” has practical value. It is an elementary unit in biodiversity

This article is protected by copyright. All rights reserved.


research, and individuated by delimitation methods. The correspondence between species

concept and species taxon is unclear, and depends on the context of speciation. For this
Accepted Article
reason the “species taxon” is a hypothesis of the evolutionary entities emerging through

the process of speciation.

Figures

FIGURE 1 The three sections of a research agenda for integrating cryptic species in biodiversity

science. For each section, we provide the main topics addressed (headings) as well as quantitative

information retrieved from the literature (italics). Arrows show how the three sections are

integrated into biodiversity science. The quest for causal mechanisms directs taxonomic practices.

Integrative taxonomy fosters understanding of the evolutionary mechanisms behind morphological

crypsis and integration of cryptic species into biodiversity research. The latter may in turn reveal

how the three mechanisms contribute to biodiversity across space and environmental gradients.

FIGURE 2 (as part of BOX 1). A snapshot of Amphipoda diversity. Lyssianasoids (A) live in the deep

sea at depths below 5000 m depth; ampeliscids (B) commonly burrow in soft sediments; corophiids

(C) are filter feeders that live in self-built tubes; caprellids (D) mimic plants on which they live;

leucothoids (E) and cyamids (F), respectively, live in association with sponges and whales; several

lines of talitroids (G) successfully colonized semi-terrestrial supra-littoral habitats, while hyperiids (H)

live pelagial, often within gelatinose plankton. Cryptic species were found in all these ecological

groups. They naturally pose the question whether the relative importance of evolutionary

mechanisms underlying morphological crypsis differs across ecological settings. All photos made by

Hans Hillewaert and used with permission (licensed under Creative Commons, photos available at

https://www.flickr.com/photos/bathyporeia/albums/72157639365477036).

This article is protected by copyright. All rights reserved.


FIGURE 3 Age distribution of 43 cryptic species of amphipods, extracted from 29 out of 120 reviewed

publications. Please note that the methods used for estimation of divergence vary between studies,
Accepted Article
see text.

FIGURE 4 An overview of practices in taxonomy to delimit cryptic species of amphipods. A) Number

of molecular markers (n = 76 studies); B) Molecular methods to species delimitation (n = 76 studies);

C) Non-molecular data used to test cryptic species hypotheses inferred from DNA taxonomy (n = 55

described species containing cryptic species complexes).

FIGURE 5 Rates of species description and discovery of cryptic species of amphipods. Upper panel

shows rates of species description between 2000 and 2015 (source: Amphipoda Newsletters vol. 30-

39). Middle panel shows cryptic species discovery rate (source Supplementary Information) over the

same time period. Lower panel shows that the proportion of discovered but undescribed cryptic

species increases through time.

FIGURE 6 A general methodology for integrating multiple species criteria into the analysis of large-

scale biodiversity patterns. Individuals can be assigned to different species hypotheses according to

different criteria. Alternative datasets can be mapped in space, and regions where different

speciation mechanisms operate can be pinpointed and tested for causal environmental factors using

multi-model inferences and variance partitioning. PTP: Poisson tree processes models; GMYC:

generalized mixed Yule-coalescent approaches; BP&P: Bayesian phylogenetics & phylogeography;

AICc: Akaike’s information criterion corrected for small sample size.

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.

You might also like