Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-26T10:29:35.774Z Has data issue: false hasContentIssue false

Evidence for enemy release in invasive common dace Leuciscus leuciscus in Ireland: a helminth community survey and systematic review

Published online by Cambridge University Press:  14 September 2020

P.A. Tierney*
Affiliation:
Department of Zoology, School of Natural Sciences, Trinity College Dublin, Dublin, Ireland
J.M. Caffrey
Affiliation:
INVAS Biosecurity Ltd., 82 Lakelands Close, Stillorgan, Co Dublin, Ireland
S.M. Matthews
Affiliation:
Department of Zoology, School of Natural Sciences, Trinity College Dublin, Dublin, Ireland
E. Costantini
Affiliation:
Department of Zoology, School of Natural Sciences, Trinity College Dublin, Dublin, Ireland
C.V. Holland
Affiliation:
Department of Zoology, School of Natural Sciences, Trinity College Dublin, Dublin, Ireland
*
Author for correspondence: P.A. Tierney, E-mail: tiernep1@tcd.ie
Rights & Permissions [Opens in a new window]

Abstract

Invasive species lose parasites in the process of invasion and tend to be less parasitized than conspecifics in the native range and sympatric native species in the invasive range (enemy release). We evaluated enemy release in an invasive freshwater fish in Ireland, common dace Leuciscus leuciscus, using helminth parasite community surveys at the core and front of the invasive range of common dace. Furthermore, we undertook a systematic literature review of helminth infection in common dace across its native range in Great Britain and Europe and invasive range in Ireland. The helminth parasite community survey revealed that invasive common dace were infected with fewer helminth species at the invasion front than at the core. Four helminth taxa – Acanthocephala, Monogenea, Digenea and Nematoda – were present in dace at the invasion core compared to only a single helminth species (Pomphorhynchus tereticollis) at the front. The systematic review revealed that invasive common dace in Ireland hosted fewer species of helminths than common dace in the native range. We report a total of three helminth species in common dace in Ireland compared to 24 in Great Britain and 84 in Continental Europe. Our results support the hypotheses that invasive populations are less parasitized than native populations and that more recently established populations host fewer parasites. However, we demonstrate that invasive species may continue to experience release from parasites long after initial invasion.

Type
Research Paper
Copyright
Copyright © The Author(s), 2020. Published by Cambridge University Press

Introduction

Invasive species are recognized globally as one of the largest threats to biodiversity as well as to human health, the environment and the economy (IUCN, 2000; Pimentel et al., Reference Pimentel, McNair and Janecka2001; Hulme, Reference Hulme2014; Lucy et al., Reference Lucy, Roy and Simpson2016). The success of invasive species may be partially explained by their capacity to escape natural enemies such as competitors, predators and parasites in the process of invasion, known as enemy release (Keane & Crawley, Reference Keane and Crawley2002; Colautti et al., Reference Colautti, Ricciardi, Grigorovich and MacIsaac2004; Liu & Stiling, Reference Liu and Stiling2006). Evidence is mixed as to whether and to what extent enemy release confers a competitive advantage to invasive species (Blossey & Notzold, Reference Blossey and Notzold1995; Lacerda et al., Reference Lacerda, Takemoto, Poulin and Pavanelli2013; Heger & Jeschke, Reference Heger, Jeschke, Jeschke and Heger2018). However, there is strong evidence that invasive species host fewer parasite species compared to conspecifics in their native range (Torchin et al., Reference Torchin, Lafferty, Dobson, McKenzie and Kuris2003; Goedknegt et al., Reference Goedknegt, Feis, Wegner, Luttikhuizen, Buschbaum, Camphuysen, van der Meer and Thieltges2016; Sarabeev et al., Reference Sarabeev, Balbuena and Morand2017; Schoeman et al., Reference Schoeman, Kruger, Secondi and du Preez2019).

Parasites are lost in the process of invasion due to a combination of failure to arrive with invading hosts and failure to persist in the new environment (Macleod et al., Reference Macleod, Paterson, Tompkins and Duncan2010). Non-native species must overcome a number of barriers to invasion: geographic barriers in order to be introduced to a novel range, growth and reproduction barriers in order to establish new populations, and dispersal barriers in order to invade new ranges (Blackburn et al., Reference Blackburn, Pyšek, Bacher, Carlton, Duncan, Jarošík, Wilson and Richardson2011). Parasites must overcome the same barriers to invasion as free-living organisms, often with the additional barrier of switching to native hosts in the introduced range, especially if introduced parasites have multi-host life cycles (Lymbery et al., Reference Lymbery, Morine, Kanani, Beatty and Morgan2014). Moreover, small founding populations of non-native hosts are likely to harbour only a subset of the parasite fauna of the source population (Hatcher & Dunn, Reference Hatcher, Dunn, MJ and AM2011; Lymbery et al., Reference Lymbery, Morine, Kanani, Beatty and Morgan2014). Among invaders that co-introduce parasites, freshwater fish are the most common and account for over half of documented cases, probably reflecting the high incidence of invasive fish in freshwater ecosystems (Lymbery et al., Reference Lymbery, Morine, Kanani, Beatty and Morgan2014). Parasites that are successfully co-introduced alongside their invasive hosts may go on to establish infections in native hosts, known as parasite spillover (Prenter et al., Reference Prenter, MacNeil, Dick and Dunn2004; Dunn & Hatcher, Reference Dunn and Hatcher2015; Sures et al., Reference Sures, Nachev, Grabner, Mehlhorn and Klimpel2019).

While parasites of invasive species are typically lost during the invasion process, invaders often acquire novel parasites as they are exposed to the local parasite community in their invasive range (Sheath et al., Reference Sheath, Williams, Reading and Robert Britton2015; Gendron & Marcogliese, Reference Gendron and Marcogliese2016; Loxton et al., Reference Loxton, Lawton, Stafford and Holland2016). Invasive species usually do not accumulate a sufficient number of local parasite species to replace the lost parasite diversity (Torchin & Mitchell, Reference Torchin and Mitchell2004; Loxton et al., Reference Loxton, Lawton, Stafford and Holland2016; Schoeman et al., Reference Schoeman, Kruger, Secondi and du Preez2019; Stuart et al., Reference Stuart, Paredis, Henttonen, Lawton, Ochoa Torres and Holland2020) (although see Gendron et al., Reference Gendron, Marcogliese and Thomas2012; Lacerda et al., Reference Lacerda, Takemoto, Poulin and Pavanelli2013). This may be due to low competency of invasive hosts for native parasites (Paterson et al., Reference Paterson, Lal, Dale, Townsend, Poulin and Tompkins2013; Tierney et al., Reference Tierney, Caffrey, Vogel, Matthews, Costantini and Holland2020) or the absence of host-specific parasites in the invaded range (Kennedy & Bush, Reference Kennedy and Bush1994; Torchin & Mitchell, Reference Torchin and Mitchell2004). However, it has been suggested that invasive species are more likely to be infected with parasite species that are directly transmitted (e.g. Monogenea, Myxosporea, Crustacea) or generalist parasite species that can utilize many species of intermediate hosts (Dobson & May, Reference Dobson, May, Mooney and Drake1986; Bauer, Reference Bauer1991; Torchin & Mitchell, Reference Torchin and Mitchell2004; Heger & Jeschke, Reference Heger, Jeschke, Jeschke and Heger2018).

Parasite diversity of an invasive host varies across its invasive range and is expected to be lowest at the expanding edge of the range (the invasion front) (White & Perkins, Reference White and Perkins2012; David et al., Reference David, Staentzel, Schlumberger, Perrot-Minnot, Beisel and Hardion2018; Stuart et al., Reference Stuart, Paredis, Henttonen, Lawton, Ochoa Torres and Holland2020). Host density at an invasion front is typically low, which inhibits parasite transmission, and parasites are likely to lag behind the establishment of their invasive hosts or be lost along the invasion corridor due to stochastic events or transience of hosts (Phillips et al., Reference Phillips, Kelehear, Pizzatto, Brown, Barton and Shine2010). Lags in parasite establishment along an invasion gradient have been observed in fish (Gaither et al., Reference Gaither, Aeby, Vignon, Meguro, Rigby, Runyon, Toonen, Wood and Bowen2013; David et al., Reference David, Staentzel, Schlumberger, Perrot-Minnot, Beisel and Hardion2018), amphibian (Phillips et al., Reference Phillips, Kelehear, Pizzatto, Brown, Barton and Shine2010) and small mammal invasions (Stuart et al., Reference Stuart, Paredis, Henttonen, Lawton, Ochoa Torres and Holland2020). This decrease in parasite infection along an invasion corridor leads to enhanced enemy release at the invasion front compared to the core of the invasive range (site of first introduction and expansion) (White & Perkins, Reference White and Perkins2012).

The cyprinid fish, common dace Leuciscus leuciscus (Linnaeus 1758), is native to continental Europe with the exception of Ireland, Spain, Italy and Greece (Welcomme, Reference Welcomme1988; Kottelat & Freyhof, Reference Kottelat and Freyhof2007). Common dace was first introduced to Ireland from Great Britain in 1889 by the accidental release of bait fish into the Munster Blackwater River, County Cork (Went, Reference Went1950). Common dace invaded the lower River Barrow in 1992 (Caffrey et al., Reference Caffrey, Hayden and Walsh2007), and by 2015 their range had spread approximately 90 km upstream (Delanty et al., Reference Delanty, Kelly and McLoone2017). Two helminth species (both native to Ireland) have previously been recorded in common dace in Ireland: the eel acanthocephalan Acanthocephalus clavula (Kane, Reference Kane1966) and the generalist acanthocephalan Pomphorhynchus tereticollis (Tierney et al., Reference Tierney, Caffrey, Vogel, Matthews, Costantini and Holland2020). The current study is the first survey of the helminth community of invasive common dace to be conducted in Ireland.

The distribution of invasive common dace in Ireland provides an opportunity to study helminth parasite community dynamics of an invasive species at the core and front of its invasive range. By sampling from the invasion core on the Munster Blackwater (over 120 years post-introduction) and from the invasion front on the upper River Barrow (less than five years post-colonization), we can utilize a ‘space for time’ substitution (e.g. Kołodziej-Sobocińska et al., Reference Kołodziej-Sobocińska, Brzeziński, Niemczynowicz and Zalewski2018) to study how the helminth parasite community of invasive common dace has changed over time since introduction. In the absence of longitudinal data across the invasion history of an invasive species, a ‘space for time’ substitution method allows us to infer the timing of events in an invasion by studying spatially separate populations at different stages of invasion. We then supplement our own data with information on the helminth species of common dace across their European range to assess the helminth parasite community between the native and invasive range of common dace.

We aim to answer three main research questions: (1) Is there evidence for enemy release – that is, is the parasite diversity of common dace in Ireland reduced compared to common dace in its native range? (2) Have common dace replaced lost diversity by acquiring local parasites in its invasive range? (3) Do we find enhanced enemy release at the invasion front?

At the macroscale, the recorded parasite diversity of a species is likely to vary across its native range due to variation in the diversity of other hosts and heterogeneity in research effort (Poulin et al., Reference Poulin, Presswell and Jorge2020). Host diversity is known to drive parasite diversity (Hechinger & Lafferty, Reference Hechinger and Lafferty2005; Thieltges et al., Reference Thieltges, Hof, Dehling, Brändle, Brandl and Poulin2011; Kamiya et al., Reference Kamiya, O'Dwyer, Nakagawa and Poulin2014; Johnson et al., Reference Johnson, Wood, Joseph, Preston, Haas and Springer2016). Due to species extirpations caused by glaciation events, northern and western Europe have a lower native biodiversity of freshwater fish than central and Mediterranean European regions (Reyjol et al., Reference Reyjol, Hugueny and Pont2007). As islands, Great Britain and Ireland have yet lower native freshwater fish biodiversity than continental Europe (Wheeler, Reference Wheeler1977). Combining the effect of enemy release and low host biodiversity, we expect helminth parasite diversity in invasive common dace to be lowest in Ireland, at an intermediate level in Great Britain and highest in Continental Europe.

Materials and methods

Invasive common dace were caught by electrofishing in July and August 2015 and 2017, from the core of the common dace invasive range on the Munster Blackwater and at the invasion front on the upper River Barrow (see Tierney et al., Reference Tierney, Caffrey, Vogel, Matthews, Costantini and Holland2020). Recent fish community surveys have recorded 11 species of fish in the Munster Blackwater comprising six native fish (brown trout Salmo trutta, Atlantic salmon Salmo salar, European eel Anguilla anguilla, lamprey Lampetra sp., three-spined stickleback Gasterosteus aculeatus and European flounder Platichthys flesus) and five non-native species (common dace, Eurasian minnow Phoxinus phoxinus, roach Rutilus rutilus, stoneloach Barbatula barbatula and gudgeon Gobio gobio) (Kelly et al., Reference Kelly, Connor, Matson, Feeney, Morrissey, Coyne and Rocks2014). In the River Barrow, recent large-scale surveys recorded 12 species comprising five native (brown trout, Atlantic salmon, European eel, lamprey and three-spined stickleback) and seven non-native species (common dace, Eurasian minnow, roach, stoneloach, gudgeon, European perch Perca fluviatilis and Northern pike Esox lucius), along with one non-native hybrid (bream × roach) (Delanty et al., Reference Delanty, Kelly and McLoone2017). Native twaite shad Alosa fallax and non-native freshwater bream Abramis brama, rudd Scardinius erythrophthalmus and tench Tinca tinca are also likely be present in the River Barrow (Delanty et al., Reference Delanty, Kelly and McLoone2017). Over the two sampling years, a total of 168 common dace were collected from the River Barrow and 119 common dace from the Munster Blackwater (table 1). Subsite sample size varied between years due to the distribution and availability of common dace in these rivers. The size range of the sampled dace were similar between regions (Gaussian Generalized Linear Model: df = 285, coefficient ± standard error (SE) = −0.02 ± 0.03, t = −0.6, P = 0.55) (table 1).

Table 1. Sample size and size range (standard length) of common dace Leuciscus leuciscus sampled (n = 287).

SD, standard deviation.

Fish were euthanized on site using anaesthesia by eugenol oil or percussive stunning and decerebration. Fish were bagged and transported on ice to the laboratory where they were frozen and stored at −20°C until dissection. The eyes, gills, alimentary tract, swim bladder and body cavity were examined for helminth parasites. Keys by Brown et al. (Reference Brown, Chubb and Veltkamp1986), Moravec (Reference Moravec1994), Gibson et al. (Reference Gibson, Jones and Bray2002), Jones et al. (Reference Jones, Bray and Gibson2005), Bray et al. (Reference Bray, Gibson and Jones2008) and Yamaguti (Reference Yamaguti1963) and the paper by Špakulová et al. (Reference Špakulová, Perrot-Minnot and Neuhaus2011) were used for parasite identification. The identification of P. tereticollis was additionally confirmed with molecular analysis (see Tierney et al., Reference Tierney, Caffrey, Vogel, Matthews, Costantini and Holland2020). Five digenean trematode specimens and three nematode specimens found in common dace were unidentifiable beyond class (Digenea) and phylum (Nematoda), respectively, because the characteristic features in these specimens were not sufficiently clear to allow for morphological identification.

Statistics were conducted using R (R Core Team, 2018). Helminth community structure was analysed at the component and infracommunity levels (Bush et al., Reference Bush, Lafferty, Lotz and Shostak1997) following Kennedy & Hartvigsen (Reference Kennedy and Hartvigsen2000). We described component community structure using total number of taxa and Simpson's Index of Diversity, calculated as:

$$D = 1-\;\displaystyle{{\mathop \sum \nolimits {\rm in\ }\lpar n-1\rpar } \over {N\lpar N-1\rpar }}$$

where n is the total number of individuals of taxon i and N is the total number of individuals of all taxa, using the R package vegan (Oksanen et al., Reference Oksanen, Blanchet and Friendly2019). The value of D ranges from zero to one and increases with increased diversity and evenness (Pielou, Reference Pielou1966; DeJong, Reference DeJong1975).

To describe infracommunity structure, we calculated mean taxonomic richness, maximum taxonomic richness, prevalence, mean intensity (Bush et al., Reference Bush, Lafferty, Lotz and Shostak1997), aggregation (mean-to-variance ratio, σ2/x̄̅; Wilson et al., Reference Wilson, Bjørnstad, Dobson, Merler, Poglayen, Randolph, Read, Skorping, Hudson, Rizzoli, Grenfell, Heesterbeek and Dobson2002) and mean and maximum Brillouin's Index of Diversity (Pielou, Reference Pielou1966). Brillouin's Index was calculated in R using the formula:

$$HB = \displaystyle{{\ln \lpar {N!} \rpar -\;\mathop \sum \nolimits {\rm ln}\lpar ni!\rpar } \over N}$$

where N is the total number of parasite individuals in a host and ni is the number of individuals of taxon i. Brillouin's index measures the diversity of a fully censused collection where HB increases with increased diversity (Pielou, Reference Pielou1966).

Prevalence of the acanthocephalan P. tereticollis was modelled with a binomial generalized linear mixed model (GLMM) and intensity of P. tereticollis was modelled with a negative binomial GLMM using the R package glmmTMB (Brooks et al., Reference Brooks, Kristensen, van Benthem, Magnusson, Berg, Nielsen, Skaug, Maechler and Bolker2017). Data on P. tereticollis combined intra-intestinal adult and extra-intestinal subadult parasite individuals (see Tierney et al., Reference Tierney, Caffrey, Vogel, Matthews, Costantini and Holland2020). Models incorporated invasion region (front or core) and host standard length as fixed effects, and subsite and year as random effects. Model fit was validated using the R package DHARMa (Hartig, Reference Hartig2019). All other helminth taxa detected in common dace occurred at too low prevalence and intensity to be fitted to a model.

A systematic literature search of helminth parasites of common dace across their European range was conducted in Web of Science using the search string (dace OR ‘Leuciscus leuciscus’) AND (parasit* OR helminth) and in Google Scholar using the search terms ‘Leuciscus leuciscus’, ‘dace’, ‘parasite’, ‘helminth’. Additional searches were conducted by replacing ‘Leuciscus leuciscus’ with ‘freshwater fish’. Records of non-helminth parasites, experimental infections and records of the subspecies Siberian dace, Leuciscus leuciscus baicalensis, were excluded. The Natural History Museum Host-Parasite database (Gibson et al., Reference Gibson, Bray and Harris2005) and references in the collected literature were also checked for additional relevant material that had not been captured in the literature search.

Results

Helminth community survey

Four helminth taxa were recorded in common dace across their invasive range in Ireland: one acanthocephalan species, one monogenean species and a number of nematode and digenean trematode individuals, which were unidentifiable to species level. All four taxa were present in common dace at the invasion core (although during different years) and one species was recorded at the invasion front. At the invasion core, we recorded three and two helminth taxa in 2015 and 2017, respectively (table 2). At the invasion front, we recorded one helminth species in 2015 but did not detect any helminths in 2017. Helminth community diversity in dace was greater at the invasion core than the invasion front, as measured by taxonomic richness, total number of taxa, Brillouin's Index and Simpson's Index (table 2). The maximum taxonomic richness per fish was two, which was recorded in both years at the invasion core. Here, Brillouin's and Simpson's diversity indices were lower in 2017 than 2015, but mean taxonomic richness increased from 2015 to 2017.

Table 2. Parameters of component and infracommunity structure in invasive common dace Leuciscus leuciscus from the front (Upper River Barrow) and core (Munster Blackwater) of its invasive range in Ireland.

SD, standard deviation.

The majority of common dace (82%) were uninfected. Among the four helminth taxa recorded in common dace, the acanthocephalan P. tereticollis was the most prevalent (table 3). The prevalence of P. tereticollis ranged from zero at the front in 2017 to 52% (95% confidence interval: 37–67) at the core in 2017. At the invasion core, prevalence of P. tereticollis was significantly higher than at the front (binomial GLMM: df = 282, coefficient ± SE = 6.02 ± 1.49, Z = 4.05, P < 0.001). Intensity of P. tereticollis infection did not differ significantly between the invasion core and front (negative binomial GLMM: df = 41, coefficient ± SE = 0.83 ± 1.30, Z = 0.63, P = 0.53). Both prevalence (binomial GLMM: df = 282, coefficient ± SE = 0.07 ± 0.01, Z = 5.71, P < 0.001) and intensity (negative binomial GLMM: df = 41, coefficient ± SE = 1.05 ± 0.17, Z = 6.35, P < 0.001) of P. tereticollis increased with host size (see also Tierney et al., Reference Tierney, Caffrey, Vogel, Matthews, Costantini and Holland2020). Pomphorhynchus tereticollis infection at the front was highly aggregated with a small number of infected fish (3.7%) hosting high worm burdens (mean intensity ± standard deviation (SD) = 48 ± 35) (table 3). Pomphorhynchus tereticollis was the only helminth that infected common dace populations at the invasion front. The other three helminth taxa all occurred in few hosts and in low numbers (prevalence less than 6% and intensity less than 2; table 3). The monogenean Discocotyle sagittata was represented by only a single individual, recorded in the core in 2017. The three nematode individuals and five digenean trematode individuals found in common dace were all recorded from the invasion core in 2015.

Table 3. Population biology parameters of helminth taxa recorded in common dace Leuciscus leuciscus at the front (Upper River Barrow) and core (Munster Blackwater) of their invasive range.

n, number of infected hosts; N, number of helminth individuals; CI, confidence interval; SE, standard error.

Systematic review

The literature search yielded 443 records of 109 helminth species from 20 European countries (see supplementary material for the full dataset). For the purposes of this study, locations were categorized into four broad biogeographical ranges: Ireland, Great Britain, Nordics and Continental Europe (excluding Nordics) (adapted from Reyjol et al., Reference Reyjol, Hugueny and Pont2007). We combined our own survey data for helminth species present with the previous record for A. clavula in common dace in Ireland to give a total of three helminth species recorded in invasive common dace in Ireland, representing two acanthocephalan species and one monogenean species. This number was low compared to the number of helminth species in the native range of common dace in Great Britain, the Nordics and Continental Europe (fig. 1). Digenean trematodes are the most commonly recorded helminth parasite species of common dace in Great Britain, the Nordics and Continental Europe (fig. 1). In the Nordics and Continental Europe, monogeneans are the next most common, while, in Great Britain, similar numbers of acanthocephalan, cestode and monogenean species are recorded.

Fig. 1. Total number of helminth species recorded in common dace Leuciscus leuciscus in its invasive range in Ireland and its native range in Great Britain, the Nordics and Continental Europe.

Although every effort was made to be comprehensive, it was not possible to gather information relating to specific site and sample size for every record due to some data being drawn from checklists that did not report such details. In total, we obtained site data for 163 records. The number of papers that recorded helminth species in common dace (as a rough proxy for sample effort) varied between countries and sites (supplementary table S1). We note that this approach is likely to underestimate sample efforts where records came from existing checklists, which were coded as one paper. Additionally, some papers included in the systematic review specifically studied only a single parasite species or taxon, including Thomas & Ollevier (Reference Thomas and Ollevier1992) on Anguillicola crassus (Belgium), Zrnčić et al. (Reference Zrnčić, Oraić, Mihaljević, Ćaleta, Zanella, Jelić and Jelić2009) on Posthodiplostomum cuticula in Croatia, Kirk & Lewis (Reference Kirk and Lewis1994) on Sanguilicola spp. in Great Britain (Anglian and Thames regions) and Perrot-Minnot et al. (Reference Perrot-Minnot, Guyonnet, Bollache and Lagrue2019) on Pomphorhynchus laevis and P. tereticollis in France. Where relatively comprehensive surveys of the helminth community have been conducted in common dace, the number of helminth species in the component community is similar across the native range. For example, in England, a total of ten species of helminth have been recorded in common dace in the River Avon (Kennedy, Reference Kennedy1974), 14 species in the River Kivijoki system in Finland (Ieshko et al., Reference Ieshko, Malakhova, Golitsyna, Lindholm, Heikkilä and Heikkilä1997), nine species in the River Morava in the Czech Republic (Gelnar et al., Reference Gelnar, Koubková, Pláňková and Jurajda1994) and ten species in the Tamis River region in Serbia (Djikanovic et al., Reference Djikanovic, Paunovic, Nikolic, Simonovic and Cakic2012) (supplementary table S1).

Discussion

Invasive common dace in Ireland had a lower diversity of helminth parasites compared to common dace in their native range and a lower diversity of helminth taxa at the invasion front compared to the invasion core. The total number of species recorded in dace was markedly lower in Ireland compared to Great Britain, the Nordics and Continental Europe where common dace are native. These findings support the hypothesis that common dace have lost parasites in the process of invasion and have experienced release from helminth parasites. Populations of common dace at the long-established invasion core had higher component and infracommunity diversity than those at the recently colonized invasion front. This supports the hypotheses that enemy release is enhanced at the invasion front and more recently established populations host less diverse parasite communities.

The lower number of total helminth species in common dace in Ireland compared to Great Britain is as expected. Holland & Kennedy (Reference Holland and Kennedy1997) found a reduced parasite fauna in most freshwater fish in Ireland compared to Great Britain and demonstrated that the helminth parasites found in Ireland are a subset of the British parasite fauna. In the literature review, we found fewer helminth parasites in Britain compared to Continental Europe, but the difference in scale and in sample effort between these areas must be noted. It would be expected that a geographically smaller area would have lower biodiversity (Rosenzweig, Reference Rosenzweig1995). Moreover, there are fewer papers reporting helminths in common dace in Great Britain than in Continental Europe, probably by virtue of fewer researchers specializing in fish parasitology. Nonetheless, when we compare Great Britain and Europe at a similar scale, the parasite community of common dace at the population level has similar richness across the native range. For example, the River Avon in England and the River Morava in the Czech Republic have both been well studied (see Kennedy, Reference Kennedy1974; Moravec, Reference Moravec2001) and have similar numbers of recorded helminth species. Our finding of enemy release in invasive common dace compared to Great Britain remains valid when examined at the scale of river/host population. Esch et al. (Reference Esch, Kennedy, Bush and Aho1988) studied three English rivers and found between eight and nine helminth species infecting common dace populations. This is in contrast to our findings of zero to four helminth taxa in total at the population level in Irish rivers. These considerations highlight the limitations of producing checklists of species: that biases are likely to exist in sampling effort, geographic coverage and possible inclusion of accidental infections (Holland & Kennedy, Reference Holland and Kennedy1997; Poulin et al., Reference Poulin, Besson, Morin and Randhawa2015, Reference Poulin, Presswell and Jorge2020; Poulin, Reference Poulin2019). Helminth parasite species richness is likely to have been underestimated in some regions due to lack of study and, in some cases (e.g. France, Perrot-Minnot et al., Reference Perrot-Minnot, Guyonnet, Bollache and Lagrue2019; and Croatia, Zrnčić et al., Reference Zrnčić, Oraić, Mihaljević, Ćaleta, Zanella, Jelić and Jelić2009), the available records focussing on a given helminth species or taxon rather than the complete helminth community.

We did not find evidence that common dace co-introduced novel helminth species to Ireland when introduced from Great Britain. Pomphorhynchus tereticollis is considered native to Europe, including the British Isles (Perrot-Minnot et al., Reference Perrot-Minnot, Špakulová, Wattier, Kotlík, Düşen, Aydoğdu and Tougard2018). It is hypothetically possible that genetically distinct strains of P. tereticollis could have been introduced from Great Britain with the introduction of common dace (O'Mahony et al., Reference O'Mahony, Bradley, Kennedy and Holland2004a, Reference O'Mahony, Kennedy and Hollandb). However, given that the Irish strain of the parasite is considered to be widespread and any founder population of genetically distinct strains co-introduced with dace would have been small, such an introduction would have been unlikely to impact the native P. tereticollis strain. Discocotyle sagittata is similarly widespread throughout Ireland and Great Britain (Kennedy, Reference Kennedy1974; Molloy et al., Reference Molloy, Holland and Poole1993; Holland & Kennedy, Reference Holland and Kennedy1997; Byrne et al., Reference Byrne, Holland, Poole and Kennedy2002). As the nematode and digenean trematode specimens recovered from common dace were unidentifiable, we cannot state whether they represent novel species to Ireland. However, considering that the prevalence and intensity of these taxa were so low, it is unlikely that these would represent co-introduced species maintained by the invasive population of common dace. Discocotyle sagittata, nematodes and digeneans were rare in common dace and each taxon was present only in one locality in one year, respectively. This could be an indication that these were accidental infections of parasites maintained by other hosts and not regular components of the dace parasite community (e.g. as with Crepidostomum sp. in eel in Thomas, Reference Thomas1958).

We found notably few monogeneans and digenean trematodes in common dace in Ireland, although they are among the most frequently reported taxa in the native range. The generalist digenean eye flukes of the family Diplostomidae are frequently acquired parasites of invasive fish species (Ondračková et al., Reference Ondračková, Dávidová, Blažek, Gelnar and Jurajda2009; Francová et al., Reference Francová, Ondračková, Polačik and Jurajda2011; Lacerda et al., Reference Lacerda, Takemoto, Poulin and Pavanelli2013; Tyutin et al., Reference Tyutin, Verbitsky, Verbitskaya and Medyantseva2013; Gendron & Marcogliese, Reference Gendron and Marcogliese2017). In Ireland, there was a notable absence of eye flukes in dace, despite eight species of the family Diplostomidae having been previously recorded in introduced cyprinids in Ireland (Holland & Kennedy, Reference Holland and Kennedy1997). Given that diplostomids are known to detrimentally impact the foraging ability and anti-predator response of their fish host (Crowden & Broom, Reference Crowden and Broom1980; Seppälä et al., Reference Seppälä, Karvonen and Valtonen2011; Lacerda et al., Reference Lacerda, Takemoto, Poulin and Pavanelli2013; Gopko et al., Reference Gopko, Mikheev and Taskinen2017), the release from eye flukes in Ireland may be advantageous for invasive common dace. Monogeneans, being directly transmitted, theoretically should be less likely to be lost during invasion since additional intermediate hosts are not required for their persistence. Host-specific monogenean parasites have been found to have been successfully co-introduced in a number of invasions and capable of host switching to native fish (Galli et al., Reference Galli, Strona, Benzoni, Crosa and Stefani2007; Sarabeev et al., Reference Sarabeev, Balbuena and Morand2018; Šimková et al., Reference Šimková, Řehulková, Rasoloariniaina, Jorissen, Scholz, Faltýnková, Mašová and Vanhove2019). Conversely, several studies have refuted that directly transmitted parasites make especially successful invaders. Lymbery et al. (Reference Lymbery, Morine, Kanani, Beatty and Morgan2014) found no association between parasite life cycle and host switching, and Lyndon & Kennedy (Reference Lyndon and Kennedy2001) found that the indirectly transmitted acanthocephalans are the most successful parasite colonizers of the British Isles. Holland & Kennedy (Reference Holland and Kennedy1997) also noted that monogeneans and digeneans are both poorly represented in the Irish helminth parasite fauna compared to that of Great Britain. The success of acanthocephalans as colonizers was mirrored in our study; two of the three species recorded in common dace in Ireland were acanthocephalans. In contrast, the Acanthocephala make up 19% of total parasite species in common dace in Great Britain and 6% of total parasite species in common dace in Continental Europe. Moreover, the acanthocephalan P. tereticollis was the only helminth to infect common dace at the invasion front.

Helminth diversity in common dace was particularly low at the invasion front. This population of common dace had effectively no parasite diversity, being infected with only one species, the generalist acanthocephalan P. tereticollis (see Tierney et al., Reference Tierney, Caffrey, Vogel, Matthews, Costantini and Holland2020). Moreover, that species was present in only one year at low prevalence and intensity, raising the possibility that stochastic events may have caused it to disappear from the front population in the latter of the two sampling years. Additionally, given that parasites tend to be spatially aggregated in their environment (Sherrard-Smith et al., Reference Sherrard-Smith, Perkins, Chadwick and Cable2015), it is possible that differences in subsite sample sizes between 2015 and 2017 resulted in differences in detection between years. Nonetheless, our findings of low helminth diversity at the invasion front support theoretical predictions (Phillips et al., Reference Phillips, Kelehear, Pizzatto, Brown, Barton and Shine2010; White & Perkins, Reference White and Perkins2012) and other empirical studies (Gendron et al., Reference Gendron, Marcogliese and Thomas2012; Loxton et al., Reference Loxton, Lawton, Stafford and Holland2016; David et al., Reference David, Staentzel, Schlumberger, Perrot-Minnot, Beisel and Hardion2018; Stuart et al., Reference Stuart, Paredis, Henttonen, Lawton, Ochoa Torres and Holland2020) that demonstrate that parasite diversity is lowest in most recently invaded parts of the range. These findings also fit the colonization time hypothesis that areas recently colonized by host species have the lowest diversity of parasites (Guégan and Kennedy, Reference Guégan and Kennedy1993).

We expected that invasive species acquire local parasites over time (Kvach & Winkler, Reference Kvach and Winkler2011; Emde et al., Reference Emde, Rueckert, Palm and Klimpel2012; White & Perkins, Reference White and Perkins2012). However, parasite diversity in common dace was low even in the invasion core where common dace have been established for over 120 years, especially considering the possibility that dace may be merely an accidental host for some of the recorded taxa. The acquisition of local parasite species by invasive common dace seems markedly slow compared to other fish invaders – for example, Ponto–Caspian gobies. Gendron et al. (Reference Gendron, Marcogliese and Thomas2012) found that, following an initial period of parasite release, invasive round goby Neogobius melanostomus in the Great Lakes accumulated native parasites and had similar parasite diversity to native species by 15 years post-colonization. Francová et al. (Reference Francová, Ondračková, Polačik and Jurajda2011) found that round goby acquired native parasites and reported no difference between the parasite species richness of native and invasive round goby populations in the River Danube. In contrast, Kvach & Stepien (Reference Kvach and Stepien2008), studying invasive round goby and tubenose goby, Proterorhinus semilunaris, in the Great Lakes, found lower parasite species richness compared to the native Ponto–Caspian range and that most parasite species were rare, but observed no increase in parasite richness over ten years. Both Kvach & Stepien (Reference Kvach and Stepien2008) and Gendron et al. (Reference Gendron, Marcogliese and Thomas2012) recorded that the helminth parasite communities of invasive gobies were composed entirely of acquired native parasites with no evidence of goby specialists or helminths co-introduced by gobies.

The low helminth diversity of common dace in Ireland may be similarly due to a lack of co-introduced cyprinid-specific parasites and an absence of suitable parasites in the invaded range. Ireland has a depauperate freshwater fish fauna, dominated by salmonids and other anadromous fish (Wheeler, Reference Wheeler1977; Fahy, Reference Fahy1989). Furthermore, Ireland has no native cyprinids, although a number of cyprinid species such as bream, rudd and roach have been introduced prior to the 20th century (Fitzmaurice, Reference Fitzmaurice1984; Fitzsimons and Igoe, Reference Fitzsimons and Igoe2004). The River Barrow has a relatively rich community of other non-native cyprinid species (Delanty et al., Reference Delanty, Kelly and McLoone2017), which one might expect would promote greater parasite diversity in common dace. However, we observed markedly low parasite diversity in this river. Given that Ireland's cyprinid fauna are the result of introductions and presumably underwent parasite loss over the course of their invasions, there may be few cyprinid-specific helminth parasite species present in the local parasite community adapted to infect common dace. If cyprinid-specialist parasites are present, their transmission may be diluted by the presence of dominant native hosts such as salmonids. Adaptation of local parasites to introduced common dace may take some time. For example, the parasite community of introduced bream in Ireland comprises 16 species, similar to the total number of helminth species in its native British range (Holland & Kennedy, Reference Holland and Kennedy1997). However, this may be explained by the length of time since colonization; unlike common dace, bream are thought to have been present in Ireland since around the 5th century (Hayden et al., Reference Hayden, Pulcini, Kelly-Quinn, O'Grady, Caffrey, McGrath and Mariani2010). Furthermore, in the ecological context of Ireland where biodiversity is known to be low, it may not be possible for an invasive species to regain similar parasite richness to its native parasite diversity without the occurrence of subsequent invasion and co-introduction events to introduce suitable and/or specific parasites.

Conclusions

Invasive common dace in Ireland are infected with a considerably less diverse parasite community than common dace in their native ranges in Britain and Continental Europe, supporting the hypothesis that this invasive species has undergone enemy release. The generalist acanthocephalan P. tereticollis was the only helminth parasite species detected in the common dace populations at the invasion front, and all helminth taxa found at the invasion core, other than P. tereticollis, were rare, despite this population being established for over 120 years. Our findings provide evidence that not only do recently established populations host less diverse parasite communities, but that enemy release may persist in invasive populations long after establishment. This apparent slow acquisition of local parasites by common dace may be explained by the biogeographical and ecological context of Ireland, where biodiversity is low and native cyprinids are absent.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0022149X20000759

Acknowledgements

We thank Inland Fisheries Ireland staff for fish collection. Thanks to M. Williams, R. Byrne, A. Mooney, C. Patmore and S. Wheldon for field assistance, and N. Fitzgerald, A. Clark and J. Picard for laboratory assistance.

Financial support

This work was supported by Inland Fisheries Ireland (P.T., grant number 13194) and a Trinity College Dublin Postgraduate Research Studentship.

Conflicts of interest

None

Ethics standards

All activities contributing to this work comply with the ethical standards of the relevant national and institutional guides on wildlife research. Electrofishing and euthanasia of fish was carried out by trained Inland Fisheries Ireland staff. This work was approved by the Trinity College Dublin School of Natural Science's Research Ethics Committee.

References

Bauer, ON (1991) Spread of parasites and diseases of aquatic organisms by acclimatization: a short review. Journal of Fish Biology 39, 679686.CrossRefGoogle Scholar
Blackburn, TM, Pyšek, P, Bacher, S, Carlton, JT, Duncan, RP, Jarošík, V, Wilson, JRU and Richardson, DM (2011) A proposed unified framework for biological invasions. Trends in Ecology and Evolution 26, 333339.CrossRefGoogle ScholarPubMed
Blossey, B and Notzold, R (1995) Evolution of increased competitive ability in invasive nonindigenous plants: a hypothesis. Journal of Ecology 83, 887889.CrossRefGoogle Scholar
Bray, RA, Gibson, D and Jones, A (2008) Keys to the trematoda, Volume 3. Wallingford, Oxfordshire, UK, CAB International Publishing and The Natural History Museum.CrossRefGoogle Scholar
Brooks, ME, Kristensen, K, van Benthem, KJ, Magnusson, A, Berg, CW, Nielsen, A, Skaug, HJ, Maechler, M and Bolker, BM (2017) glmmTMB balances speed and flexibility among packages for zero-inflated generalized linear mixed modeling. The R Journal 9, 378400.CrossRefGoogle Scholar
Brown, AF, Chubb, JC and Veltkamp, CJ (1986) A key to the species of Acanthocephala parasitic in British freshwater fishes. Journal of Fish Biology 28, 327334.CrossRefGoogle Scholar
Bush, AO, Lafferty, KD, Lotz, JM and Shostak, AW (1997) Parasitology meets ecology on its own terms: Margolis et al. revisited. The Journal of Parasitology 83, 575583.CrossRefGoogle Scholar
Byrne, CJ, Holland, CV, Poole, R and Kennedy, CR (2002) Comparison of the macroparasite communities of wild and stocked brown trout (Salmo trutta L.) in the west of Ireland. Parasitology 124, 435445.CrossRefGoogle Scholar
Caffrey, JM, Hayden, B and Walsh, T (2007) Dace (Leuciscus leuciscus L.): an invasive fish species in Ireland. Irish Freshwater Fisheries, Ecology and Management 5. Dublin, Ireland, Central Fisheries Board.Google Scholar
Colautti, RI, Ricciardi, A, Grigorovich, IA and MacIsaac, HJ (2004) Is invasion success explained by the enemy release hypothesis? Ecology Letters 7, 721733.CrossRefGoogle Scholar
Crowden, AE and Broom, DM (1980) Effects of the eyefluke, Diplostomum spathaceum, on the behaviour of dace (Leuciscus leuciscus). Animal Behaviour 28, 287294.CrossRefGoogle Scholar
David, GM, Staentzel, C, Schlumberger, O, Perrot-Minnot, MJ, Beisel, JN and Hardion, L (2018) A minimalist macroparasite diversity in the round goby of the Upper Rhine reduced to an exotic acanthocephalan lineage. Parasitology 145, 10201026.CrossRefGoogle Scholar
DeJong, TM (1975) A comparison of three diversity indices based on their components of richness and evenness. Oikos 26, 222.CrossRefGoogle Scholar
Delanty, K, Kelly, FL, McLoone, P, et al. (2017) Fish stock assessment of the river barrow catchment 2015. Dublin, Inland Fisheries Ireland.Google Scholar
Djikanovic, V, Paunovic, M, Nikolic, V, Simonovic, P and Cakic, P (2012) Parasitofauna of freshwater fishes in the Serbian open waters: a checklist of parasites of freshwater fishes in Serbian open waters. Reviews in Fish Biology and Fisheries 22, 297324.CrossRefGoogle Scholar
Dobson, A and May, R (1986) Patterns of invasions by pathogens and parasites. pp. 5876in Mooney, H and Drake, J (Eds) Ecology and biological invasions of North America and Hawaii. Berlin, Springer-Verlag.CrossRefGoogle Scholar
Dunn, AM and Hatcher, MJ (2015) Parasites and biological invasions: parallels, interactions, and control. Trends in Parasitology 31, 189199.CrossRefGoogle ScholarPubMed
Emde, S, Rueckert, S, Palm, HW and Klimpel, S (2012) Invasive Ponto-Caspian amphipods and fish increase the distribution range of the Acanthocephalan Pomphorhynchus tereticollis in the River Rhine. PLoS ONE 7, e53218.CrossRefGoogle ScholarPubMed
Esch, GW, Kennedy, CR, Bush, AO and Aho, JM (1988) Patterns in helminth communities in freshwater fish in Great Britain: alternative strategies for colonization. Parasitology 96, 519532.CrossRefGoogle ScholarPubMed
Fahy, E (1989) Conservation and management of brown trout, Salmo trutta, in Ireland. Freshwater Biology 21, 99109.CrossRefGoogle Scholar
Fitzmaurice, P (1984) The effects of freshwater fish introductions into Ireland. EIFAC technical paper 42 (Supplement 2), 448457.Google Scholar
Fitzsimons, M and Igoe, F (2004) Freshwater fish conservation in the Irish republic: a review of pressures and legislation impacting on conservation efforts. Biology and Environment 104, 1732.Google Scholar
Francová, K, Ondračková, M, Polačik, M and Jurajda, P (2011) Parasite fauna of native and non-native populations of Neogobius melanostomus (Pallas, 1814) (Gobiidae) in the longitudinal profile of the Danube River. Journal of Applied Ichthyology 27, 879886.CrossRefGoogle Scholar
Gaither, MR, Aeby, G, Vignon, M, Meguro, YI, Rigby, M, Runyon, C, Toonen, RJ, Wood, CL and Bowen, BW (2013) An invasive fish and the time-lagged spread of its parasite across the Hawaiian archipelago. PloS One 8, e56940.CrossRefGoogle ScholarPubMed
Galli, P, Strona, G, Benzoni, F, Crosa, G and Stefani, F (2007) Monogenoids from freshwater fish in Italy, with comments on alien species. Comparative Parasitology 74, 264272.CrossRefGoogle Scholar
Gelnar, M, Koubková, B, Pláňková, H and Jurajda, P (1994) Report on metazoan parasites of fishes of the River Morava with remarks on the effects of water pollution. Helminthologia 31, 4756.Google Scholar
Gendron, AD and Marcogliese, DJ (2016) Reduced survival of a native parasite in the invasive round goby: Evidence for the dilution hypothesis? Aquatic Invasions 11, 189198.CrossRefGoogle Scholar
Gendron, AD and Marcogliese, DJ (2017) Enigmatic decline of a common fish parasite (Diplostomum spp.) in the St. Lawrence River: evidence for a dilution effect induced by the invasive round goby. International Journal for Parasitology: Parasites and Wildlife 6, 402411.Google Scholar
Gendron, AD, Marcogliese, DJ and Thomas, M (2012) Invasive species are less parasitized than native competitors, but for how long? The case of the round goby in the Great Lakes-St. Lawrence Basin. Biological Invasions 14, 367384.CrossRefGoogle Scholar
Gibson, D, Jones, A and Bray, RA (2002) Keys to the trematoda, Volume 1. Wallingford, Oxfordshire, UK, CAB International Publishing and The Natural History Museum.CrossRefGoogle Scholar
Gibson, D, Bray, RA and Harris, EA (2005) The Natural History Museum host-parasite database. London. Available at www.nhm.ac.uk/research-curation/scientific-resources/taxonomy-systematics/host-parasites/index.html (accessed 1 February 2020).Google Scholar
Goedknegt, MA, Feis, ME, Wegner, KM, Luttikhuizen, PC, Buschbaum, C, Camphuysen, K, van der Meer, J and Thieltges, DW (2016) Parasites and marine invasions: ecological and evolutionary perspectives. Journal of Sea Research 113, 1127.CrossRefGoogle Scholar
Gopko, M, Mikheev, VN and Taskinen, J (2017) Deterioration of basic components of the anti-predator behavior in fish harboring eye fluke larvae. Behavioral Ecology and Sociobiology 71, 68.CrossRefGoogle Scholar
Guégan, J-F and Kennedy, C (1993) Maximum local helminth parasite community richness in British freshwater fish: a test of the colonization time hypothesis. Parasitology 106, 91100.CrossRefGoogle ScholarPubMed
Hartig, F (2019) DHARMa: residual diagnostics for hierarchical (multi-level/mixed) regression models. R package version 0.2.6. Available at https://CRAN.R-project.org/, accessed 22 June 2020.Google Scholar
Hatcher, MJ and Dunn, AM (2011) Parasites and invasions. pp. 224264in MJ, Hatcher and AM, Dunn (Eds) Parasites in ecological communities: from interactions to ecosystems. Cambridge, Cambridge University Press.CrossRefGoogle Scholar
Hayden, B, Pulcini, D, Kelly-Quinn, M, O'Grady, M, Caffrey, J, McGrath, A and Mariani, S (2010) Hybridisation between two cyprinid fishes in a novel habitat: genetics, morphology and life-history traits. BMC Evolutionary Biology 10, 111.CrossRefGoogle Scholar
Hechinger, RF and Lafferty, KD (2005) Host diversity begets parasite diversity: bird final hosts and trematodes in snail intermediate hosts. Proceedings of the Royal Society B 272, 10591066.CrossRefGoogle ScholarPubMed
Heger, T and Jeschke, JM (2018) Enemy release hypothesis. pp. 92102in Jeschke, JM and Heger, T (Eds) Invasion biology: hypotheses and evidence. Boston, Massachusetts, CABI.CrossRefGoogle Scholar
Holland, CV and Kennedy, CR (1997) A checklist of parasitic helminth and crustacean species recorded in freshwater fish from Ireland. Biology and Environment 97, 225243.Google Scholar
Hulme, PE (2014) Invasive species challenge the global response to emerging diseases. Trends in Parasitology 30, 267270.CrossRefGoogle ScholarPubMed
Ieshko, EP, Malakhova, RP and Golitsyna, NB (1997) Ecological characteristics of lake fish parasite fauna formation in the River Kivijoki system. pp. 311318in Lindholm, T, Heikkilä, R and Heikkilä, M (Eds) Ecosystems, fauna and flora of the Finnish-Russian Nature Reserve Friendship. Helsinki, Finnish Environment Institute.Google Scholar
IUCN (2000) IUCN guidelines for the prevention of biodiversity loss caused by invasive alien species. Gland, Switzerland, IUCN.Google Scholar
Johnson, PTJ, Wood, CL, Joseph, MB, Preston, DL, Haas, SE and Springer, YP (2016) Habitat heterogeneity drives the host-diversity-begets-parasite-diversity relationship: evidence from experimental and field studies. Ecology Letters 19, 752761.CrossRefGoogle ScholarPubMed
Jones, A, Bray, RA and Gibson, D (2005) Keys to the trematoda. Volume 2. Wallingford, Oxfordshire, UK, CAB International Publishing and The Natural History Museum.CrossRefGoogle Scholar
Kamiya, T, O'Dwyer, K, Nakagawa, S and Poulin, R (2014) Host diversity drives parasite diversity: meta-analytical insights into patterns and causal mechanisms. Ecography 37, 689697.CrossRefGoogle Scholar
Kane, MB (1966) Parasites of Irish fishes. Scientific Proceedings of the Royal Dublin Society B 1, 205220.Google Scholar
Keane, RM and Crawley, MJ (2002) Exotic plant invasions and the enemy release hypothesis. Trends in Ecology and Evolution 17, 164170.CrossRefGoogle Scholar
Kelly, FL, Connor, L, Matson, R, Feeney, R, Morrissey, E, Coyne, J and Rocks, K (2014) Sampling fish for the water framework directive - summary report 2013. Dublin,Inland Fisheries Ireland.Google Scholar
Kennedy, CR (1974) A checklist of British and Irish freshwater fish parasites with notes on their distribution. Journal of Fish Biology 6, 613644.CrossRefGoogle Scholar
Kennedy, CR and Bush, AO (1994) The relationship between pattern and scale in parasite communities: a stranger in a strange land. Parasitology 109, 187196.CrossRefGoogle Scholar
Kennedy, CR and Hartvigsen, RA (2000) Richness and diversity of intestinal metazoan communities in brown trout Salmo trutta compared to those of eels Anguilla anguilla in their European heartlands. Parasitology 121, 5564.CrossRefGoogle ScholarPubMed
Kirk, RS and Lewis, JW (1994) The distribution and host range of species of the blood fluke Sanguinicola in British freshwater fish. Journal of Helminthology 68, 315318.CrossRefGoogle ScholarPubMed
Kołodziej-Sobocińska, M, Brzeziński, M, Niemczynowicz, A and Zalewski, A (2018) High parasite infection level in non-native invasive species: it is just a matter of time. Ecography 41, 12831294.CrossRefGoogle Scholar
Kottelat, M and Freyhof, J (2007) Handbook of European freshwater fishes. Berlin, Publications M. Kottelat, Cornol, Switzerland/J. Freyhof.Google Scholar
Kvach, Y and Stepien, CA (2008) Metazoan parasites of introduced round and tubenose gobies in the Great Lakes: support for the “enemy release hypothesis.”. Journal of Great Lakes Research 34, 2335.CrossRefGoogle Scholar
Kvach, Y and Winkler, HM (2011) The colonization of the invasive round goby Neogobius melanostomus by parasites in new localities in the southwestern Baltic Sea. Parasitology Research 109, 769780.CrossRefGoogle ScholarPubMed
Lacerda, ACF, Takemoto, RM, Poulin, R and Pavanelli, GC (2013) Parasites of the fish Cichla piquiti (Cichlidae) in native and invaded Brazilian basins: release not from the enemy, but from its effects. Parasitology Research 112, 279288.CrossRefGoogle Scholar
Liu, H and Stiling, P (2006) Testing the enemy release hypothesis: a review and meta-analysis. Biological Invasions 8, 15351545.CrossRefGoogle Scholar
Loxton, KC, Lawton, C, Stafford, P and Holland, CV (2016) Reduced helminth parasitism in the introduced bank vole (Myodes glareolus): more parasites lost than gained. International Journal for Parasitology: Parasites and Wildlife 5, 175183.Google ScholarPubMed
Lucy, FE, Roy, H, Simpson, A, et al. (2016) INVASIVESNET towards an international association for open knowledge on invasive alien species. Management of Biological Invasions 7, 131139.CrossRefGoogle Scholar
Lymbery, AJ, Morine, M, Kanani, HG, Beatty, SJ and Morgan, DL (2014) Co-invaders: the effects of alien parasites on native hosts. International Journal for Parasitology: Parasites and Wildlife 3, 171177.Google ScholarPubMed
Lyndon, AR and Kennedy, CR (2001) Colonisation and extinction in relation to competition and resource partitioning in acanthocephalans of freshwater fishes of the British Isles. Folia Parasitologica 48, 3746.CrossRefGoogle ScholarPubMed
Macleod, CJ, Paterson, AM, Tompkins, DM and Duncan, RP (2010) Parasites lost - do invaders miss the boat or drown on arrival? Ecology Letters 13, 516527.CrossRefGoogle ScholarPubMed
Molloy, S, Holland, C and Poole, R (1993) Helminth parasites of brown and sea trout Salmo trutta L. from the west coast of Ireland. Biology and Environment: Proceedings of the Royal Irish Academy 93B, 137142.Google Scholar
Moravec, F (1994) Parasitic nematodes of freshwater fishes of Europe. Dordrecht, the Netherlands, Kluwer Academic Publishers.Google Scholar
Moravec, F (2001) Checklist of the metazoan parasites of fishes of the Czech Republic and the Slovak Republic (1873-2000). Czech Republic, Academia, Praha.Google Scholar
Oksanen, J, Blanchet, FG and Friendly, M, et al. (2019) vegan: Community Ecology Package. R package version 2.5-6. Available at https://CRAN.R-project.org/package=vegan.Google Scholar
O'Mahony, EM, Bradley, DG, Kennedy, CR and Holland, CV (2004a) Evidence for the hypothesis of strain formation in Pomphorhynchus laevis (Acanthocephala): an investigation using mitochondrial DNA sequences. Parasitology 129, 341347.CrossRefGoogle Scholar
O'Mahony, EM, Kennedy, CR and Holland, CV (2004b) Comparison of morphological characters in Irish and English populations of the acanthocephalan Pomphorhynchus laevis (Müller, 1776). Systematic Parasitology 59, 147157.CrossRefGoogle Scholar
Ondračková, M, Dávidová, M, Blažek, R, Gelnar, M and Jurajda, P (2009) The interaction between an introduced fish host and local parasite fauna: Neogobius kessleri in the middle Danube River. Parasitology Research 105, 201208.CrossRefGoogle ScholarPubMed
Paterson, RA, Lal, A, Dale, M, Townsend, CR, Poulin, R and Tompkins, DM (2013) Relative competence of native and exotic fish hosts for two generalist native trematodes. International Journal for Parasitology: Parasites and Wildlife 2, 136143.Google ScholarPubMed
Perrot-Minnot, MJ, Špakulová, M, Wattier, R, Kotlík, P, Düşen, S, Aydoğdu, A and Tougard, C (2018) Contrasting phylogeography of two Western Palaearctic fish parasites despite similar life cycles. Journal of Biogeography 45, 101115.CrossRefGoogle Scholar
Perrot-Minnot, MJ, Guyonnet, E, Bollache, L and Lagrue, C (2019) Differential patterns of definitive host use by two fish acanthocephalans occurring in sympatry: Pomphorhynchus laevis and Pomphorhynchus tereticollis. International Journal for Parasitology: Parasites and Wildlife 8, 135144.Google ScholarPubMed
Phillips, BL, Kelehear, C, Pizzatto, L, Brown, GP, Barton, D and Shine, R (2010) Parasites and pathogens lag behind their host during periods of host range advance. Ecology 91, 872881.CrossRefGoogle ScholarPubMed
Pielou, EC (1966) The measurement of diversity in different types of biological collections. Journal of Theoretical Biology 13, 131144.CrossRefGoogle Scholar
Pimentel, D, McNair, S, Janecka, J, et al. (2001) Economic and environmental threats of alien plant, animal, and microbe invasions. Agriculture, Ecosystems and Environment 84, 120.CrossRefGoogle Scholar
Poulin, R (2019) Best practice guidelines for studies of parasite community ecology. Journal of Helminthology 93, 811.CrossRefGoogle ScholarPubMed
Poulin, R, Besson, AA, Morin, MB and Randhawa, HS (2015) Missing links: testing the completeness of host-parasite checklists. Parasitology 143, 114122.CrossRefGoogle ScholarPubMed
Poulin, R, Presswell, B and Jorge, F (2020) The state of fish parasite discovery and taxonomy: a critical assessment and a look forward. International Journal for Parasitology.https://doi.org/10.1016/j.ijpara.2019.12.009.CrossRefGoogle Scholar
Prenter, J, MacNeil, C, Dick, JTA and Dunn, AM (2004) Roles of parasites in animal invasions. Trends in Ecology and Evolution 19, 385390.CrossRefGoogle ScholarPubMed
R Core Team (2018) R: a language and environment for statistical computing. Vienna, Austria, R Foundation for Statistical Computing.Google Scholar
Reyjol, Y, Hugueny, B, Pont, D, et al. (2007) Patterns in species richness and endemism of European freshwater fish. Global Ecology and Biogeography 16, 6575.CrossRefGoogle Scholar
Rosenzweig, ML (1995) Species diversity in space and time. Cambridge, Cambridge University Press.CrossRefGoogle Scholar
Sarabeev, V, Balbuena, JA and Morand, S (2017) Testing the enemy release hypothesis: abundance and distribution patterns of helminth communities in grey mullets (Teleostei: Mugilidae) reveal the success of invasive species. International Journal for Parasitology 47, 687696.CrossRefGoogle ScholarPubMed
Sarabeev, V, Balbuena, JA and Morand, S (2018) Invasive parasites are detectable by their abundance-occupancy relationships: the case of helminths from Liza haematocheilus (Teleostei: Mugilidae). International Journal for Parasitology 48, 793803.CrossRefGoogle Scholar
Schoeman, AL, Kruger, N, Secondi, J and du Preez, LH (2019) Repeated reduction in parasite diversity in invasive populations of Xenopus laevis: a global experiment in enemy release. Biological Invasions 21, 13231338.CrossRefGoogle Scholar
Seppälä, O, Karvonen, A and Valtonen, ET (2011) Eye fluke-induced cataracts in natural fish populations: is there potential for host manipulation? Parasitology 138, 209214.CrossRefGoogle ScholarPubMed
Sheath, DJ, Williams, CF, Reading, AJ and Robert Britton, J (2015) Parasites of non-native freshwater fishes introduced into England and Wales suggest enemy release and parasite acquisition. Biological Invasions 17, 22352246.CrossRefGoogle Scholar
Sherrard-Smith, E, Perkins, SE, Chadwick, EA and Cable, J (2015) Spatial and seasonal factors are key determinants in the aggregation of helminths in their definitive hosts: Pseudamphistomum truncatum in otters (Lutra lutra). International Journal for Parasitology 45, 7583.CrossRefGoogle Scholar
Šimková, A, Řehulková, E, Rasoloariniaina, JR, Jorissen, MWP, Scholz, T, Faltýnková, A, Mašová, Š and Vanhove, MPM (2019) Transmission of parasites from introduced tilapias: a new threat to endemic Malagasy ichthyofauna. Biological Invasions 21, 803819.CrossRefGoogle Scholar
Špakulová, M, Perrot-Minnot, M-JJ and Neuhaus, B (2011) Resurrection of Pomphorhynchus tereticollis (Rudolphi, 1809) (Acanthocephala: Pomphorhynchidae) based on new morphological and molecular data. Helminthologia 48, 268277.CrossRefGoogle Scholar
Stuart, P, Paredis, L, Henttonen, H, Lawton, C, Ochoa Torres, CA and Holland, CV (2020) The hidden faces of a biological invasion: parasite dynamics of invaders and natives. International Journal for Parasitology 50, 111123.CrossRefGoogle ScholarPubMed
Sures, B, Nachev, M and Grabner, D (2019) The River Rhine as hotspot of parasite invasions. pp. 409429in Mehlhorn, H and Klimpel, S (Eds) Parasite and disease spread by major rivers on earth. Parasitology research monographs, Vol 12. Cham, Springer.CrossRefGoogle Scholar
Thieltges, DW, Hof, C, Dehling, DM, Brändle, M, Brandl, R and Poulin, R (2011) Host diversity and latitude drive trematode diversity patterns in the European freshwater fauna. Global Ecology and Biogeography 20, 675682.CrossRefGoogle Scholar
Thomas, JD (1958) Studies on Crepidostomum metoecus (Braun) and C. farionis (Müller), parasitic in Salmo trutta L. and S. salar L. in Britain. Parasitology 48, 336352.CrossRefGoogle Scholar
Thomas, K and Ollevier, F (1992) Paratenic hosts of the swimbladder nematode Anguillicola crassus. Diseases of Aquatic Organisms 13, 165174.CrossRefGoogle Scholar
Tierney, PA, Caffrey, JM, Vogel, S, Matthews, SM, Costantini, E and Holland, CV (2020) Invasive freshwater fish (Leuciscus leuciscus) acts as a sink for a parasite of native brown trout Salmo trutta. Biological Invasions 22, 22352250.CrossRefGoogle Scholar
Torchin, ME and Mitchell, CE (2004) Parasites, pathogens, and invasions by plants and animals. Frontiers in Ecology and the Environment 2, 183190.CrossRefGoogle Scholar
Torchin, ME, Lafferty, KD, Dobson, AP, McKenzie, VJ and Kuris, AM (2003) Introduced species and their missing parasites. Nature 421, 628630.CrossRefGoogle ScholarPubMed
Tyutin, AV, Verbitsky, VB, Verbitskaya, TI and Medyantseva, EN (2013) Parasites of alien aquatic animals in the upper Volga basin. Russian Journal of Biological Invasions 4, 5459.CrossRefGoogle Scholar
Welcomme, RL (1988) International introductions of inland aquatic species. Technical paper, number 294. Rome, Italy, Food and Agriculture Organization of the United Nations.Google Scholar
Went, AEJ (1950) Notes on the introduction of some freshwater fish into Ireland. Journal of the Department of Agriculture 47, 38.Google Scholar
Wheeler, A (1977) The origin and distribution of the freshwater fishes of the British Isles. Journal of Biogeography 4, 124.CrossRefGoogle Scholar
White, TA and Perkins, SE (2012) The ecoimmunology of invasive species. Functional Ecology 26, 13131323.CrossRefGoogle Scholar
Wilson, K, Bjørnstad, O, Dobson, A, Merler, S, Poglayen, G, Randolph, S, Read, A and Skorping, A (2002) Heterogeneities in macroparasite infections: patterns and processes. pp. 644in Hudson, PJ, Rizzoli, A, Grenfell, BT, Heesterbeek, H and Dobson, AP (Eds) The ecology of wildlife diseases. Oxford, Oxford University Press.Google Scholar
Yamaguti, S (1963) Systema Helminthum. Volume IV. Monogenea and Aspidocotylea. New York, Interscience Publishers Inc.Google Scholar
Zrnčić, S, Oraić, D, Mihaljević, Ž, Ćaleta, M, Zanella, D, Jelić, D and Jelić, M (2009) First observation of Posthodiplostomum cuticola (Nordmann, 1832) metacercariae in cypriniformes from Croatia. Helminthologia 46, 112116.CrossRefGoogle Scholar
Figure 0

Table 1. Sample size and size range (standard length) of common dace Leuciscus leuciscus sampled (n = 287).

Figure 1

Table 2. Parameters of component and infracommunity structure in invasive common dace Leuciscus leuciscus from the front (Upper River Barrow) and core (Munster Blackwater) of its invasive range in Ireland.

Figure 2

Table 3. Population biology parameters of helminth taxa recorded in common dace Leuciscus leuciscus at the front (Upper River Barrow) and core (Munster Blackwater) of their invasive range.

Figure 3

Fig. 1. Total number of helminth species recorded in common dace Leuciscus leuciscus in its invasive range in Ireland and its native range in Great Britain, the Nordics and Continental Europe.

Supplementary material: File

Tierney et al. supplementary material

Tierney et al. supplementary material 1

Download Tierney et al. supplementary material(File)
File 96.8 KB
Supplementary material: File

Tierney et al. supplementary material

Tierney et al. supplementary material 2

Download Tierney et al. supplementary material(File)
File 131.9 KB