Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-24T22:55:02.671Z Has data issue: false hasContentIssue false

Substrate selection of Christmas tree worms (Spirobranchus spp.) in the Gulf of Eilat, Red Sea

Published online by Cambridge University Press:  31 January 2017

Orly Perry*
Affiliation:
The Mina and Everard Goodman Faculty of Life Sciences, Bar-Ilan University, Ramat-Gan 52900, Israel
Yuval Sapir
Affiliation:
The Botanical Garden, School of Plant Sciences and Food Security, Tel Aviv University, Ramat Aviv, Tel Aviv 69978, Israel
Gad Perry
Affiliation:
Department of Natural Resource Management, Texas Tech University, Lubbock, Texas 79409, USA
Harry Ten Hove
Affiliation:
Naturalis Biodiversity Center, P.O. Box 9517, 2300 RA Leiden, the Netherlands
Maoz Fine
Affiliation:
The Mina and Everard Goodman Faculty of Life Sciences, Bar-Ilan University, Ramat-Gan 52900, Israel The Interuniversity Institute for Marine Science in Eilat, P.O. Box 469, Eilat 88103, Israel
*
Correspondence should be addressed to: O. Perry, The Mina and Everard Goodman Faculty of Life Sciences, Bar-Ilan University, Ramat-Gan 52900, Israel email: orlyperry1@gmail.com
Rights & Permissions [Opens in a new window]

Abstract

Christmas tree worms (Spirobranchus spp.) are prominent sessile organisms inhabiting hermatypic corals in tropical and sub-tropical reefs. Until recently, most of the larger Spirobranchus species were considered to be in obligatory associations with live hermatypic corals. However, recent studies indicate that some Spirobranchus species can build tubes on artificial substrate as well and that others may show preferences for using specific species of corals and hydrozoans as substrates. In the present study, we conducted a survey of Spirobranchus spp. substrate preference in the Gulf of Eilat. We found seven morphotaxa of Spirobranchus, of which two may be a single new species. We show that Spirobranchus taxa differ not only in their morphology, but also in their substrate use. Our results demonstrate that the ecological niche of Spirobranchus is species-specific, and a putative innate preference exists for some substrates.

Type
Research Article
Copyright
Copyright © Marine Biological Association of the United Kingdom 2017 

INTRODUCTION

Darwin (Reference Darwin1859) may have been the first to suggest that differences between closely related species may be the result of adaptation to reducing interspecific competition via selection of different habitats. In a series of articles beginning in the 1960s MacArthur and his colleagues further developed this idea, now most often referred to as resource partitioning or, more specifically, habitat partitioning (MacArthur & Levins, Reference MacArthur and Levins1964). Although early work primarily focused on terrestrial habitats, a recent review (Bowen et al., Reference Bowen, Rocha, Toonen and Karl2013) concluded that the ecological boundaries can be important arenas for evolutionary processes in marine and other ecosystems.

Levesque et al. (Reference Levesque, Juniper and Marcus2003) showed that resource partitioning can occur in Polychaeta. As with all members of the family Serpulidae, Spirobranchus Blainville, 1818 juveniles become sessile and build their own tube. The 37 Spirobranchus species currently recognized (Read et al., Reference Read, Fiege, Bellan, Read and Fauchald2016) are mostly distributed in subtropical and tropical zones, with S. triqueter (Linnaeus, Reference Linnaeus1767) even occurring in the Arctic (Pillai, Reference Pillai2009; Rzhavsky et al., Reference Rzhavsky, Kupriyanova, Sikorski and Dahle2014). The larger species are often recorded inhabiting corals, whereas smaller representatives may occur on almost any solid substrate. A substantial amount of research has been conducted on Spirobranchus giganteus sensu latissimo (Marsden, Reference Marsden1987; Nishi & Nishihira, Reference Nishi and Nishihira1996; Petitjean & Myers, Reference Petitjean and Myers2005; Rowley, Reference Rowley2008), often erroneously reported from all tropical regions as S. giganteus (Pallas, 1766), a species restricted to the Caribbean. The larger species of Spirobranchus are often named ‘Christmas tree worms’ for the bright colours and spiral arrangement of their radioles.

Frank & ten Hove (Reference Frank and ten Hove1992) hypothesized that morphology of branchial crowns may be correlated to different filtering strategies, which may be an indication of resource partitioning. However, despite the wide distribution and striking appearance of members of the genus, ecological processes such as habitat partitioning have not received much attention.

Most of the larger Spirobranchus species are associated with hermatypic corals (see below), their tubes embedded in the coral skeleton. Figure 1 shows different substrates colonized by Spirobranchus in the Gulf of Eilat. Growth rate and angle of the worm's tube are correlated with that of the coral in such a way that the opening of the tube will always stay on top of the surface (Nishi & Nishihira, Reference Nishi and Nishihira1999). Some taxa, such as the Caribbean S. giganteus and the Indo-West Pacific S. corniculatus, are thought to be obligate inhabitants of living corals (Hunte et al., Reference Hunte, Conlin and Marsden1990a; Marsden & Meeuwig, Reference Marsden and Meeuwig1990; Nishi, Reference Nishi1996). Members of the Spirobranchus giganteus complex are abundant on some coral species, rare or even absent from others. Coral preferences may differ between Spirobranchus taxa and thus between biogeographic regions. Dai & Yang (Reference Dai and Yang1995) found that in the coral reefs of Southern Taiwan, S. corniculatus mostly inhabit the corals Porites lutea Milne Edwards & Haime, 1851, P. lobata Dana, 1846, P. lichen Dana, 1846 and Montipora informis Bernard, 1897. In the bank reef off the West coast of Barbados the hexacoral species Diploria strigosa (Dana, 1846), Porites astreoides Lamarck, 1816 and the hydrozoan coral Millepora complanata Lamarck, 1816 are most heavily colonized by S. giganteus while Colpophyllia natans (Houttuyn, 1772), Dendrogyra cylindrus Ehrenberg, 1834, Dichocoenia stokesii Milne Edwards & Haime, 1848, Eusmilia fastigiata (Pallas, 1766) Meandrina meandrites (Linnaeus, 1758) and Mycetophyllia spp. Milne Edwards & Haime, 1848 were not colonized (Conlin, Reference Conlin1988; Hunte et al., Reference Hunte, Conlin and Marsden1990a). Specific searches for associated fauna in potential hosts (not only corals) may result in new association records (Hoeksema & ten Hove, Reference Hoeksema and ten Hove2016; Hoeksema et al., Reference Hoeksema, van Beusekom, ten Hove, Ivanenko, van der Meij and van Moorsel2016).

Fig. 1. Different substrates colonized by Spirobranchus, reef of Eilat, Israel. (A) Spirobranchus gardineri associated with Cyphastrea sp. (B) S. gaymardi associated with Pocillopora sp. (C) Spirobranchus sp. associated with Stylophora sp. (D) S. cruciger associated with Millepora sp. (E) S. corniculatus associated with Acropora sp. (F) Spirobranchus sp. on artificial substrate. Arrow: operculum. (A, D, F) Photos: O. Perry; (B, C, E) photos: A. Hallakoun.

The co-occurrence of worm and coral may be of a mutualistic character. DeVantier et al. (Reference DeVantier, Reichelt and Bradbury1986) showed that the branchiae and especially the operculum of the worm can provide protection to the coral from predators. In some conditions the feeding behaviour may even enhance water flow close to the coral's surface (Strathmann et al., Reference Strathmann, Cameron and Strathmann1984), which might locally prevent bleaching (Ben-Tzvi et al., Reference Ben-Tzvi, Einbinder and Brokovich2006). However, settlement of Spirobranchus taxa on corals is an antagonistic interaction. Corals can be aggressive towards other organisms and compete for resources using sweeper tentacles (Genin & Karp, Reference Genin and Karp1994). Perhaps because of this, some of the more subordinate corals as Porites lutea, P. lobata and P. lichen were found to be most colonized by Spirobranchus corniculatus while the more aggressive corals such as Mycedium elephantotus (Pallas, 1766), Merulina ampliata (Ellis & Solander, 1786) and Galaxea astreata (Lamarck, 1816) were less colonized (Dai, Reference Dai1990; Dai & Yang, Reference Dai and Yang1995). Settling larvae of a species better adapted to cope with coral aggression are more likely to survive. Larval settlement preference has been demonstrated in Spirobranchus (Hunte et al., Reference Hunte, Conlin and Marsden1990a, Marsden & Meeuwig, Reference Marsden and Meeuwig1990) and may be an indicator of host-specific adaptations. An alternative view was proposed by Rowley (Reference Rowley2008), who suggested that S. corniculatus sensu stricto (as S. giganteus) contributes to the success of its hosts in several ways and the relationship should be thought of as mutualism rather than commensalism. Regardless, Hunte et al. (Reference Hunte, Conlin and Marsden1990a) demonstrated that substrate selection by planktonic larvae of S. giganteus was correlated with worm size, such that individuals located on the preferred species of coral reached a larger size. Thus, length can be a surrogate of high performance and provide evidence for adaptive host-specific interactions.

Living hermatypic corals are not the only substrate for the larger Spirobranchus worms. Recently, there have been records that Spirobranchus giganteus s. str. can also live on octocorals, rubble, and substrates such as oil buoys or pillars of a pier (Nygaard, Reference Nygaard2008; Skinner et al., Reference Skinner, Tenório, Penha and Soares2012; Hoeksema et al., Reference Hoeksema, Wah Lau and ten Hove2015). Although purporting to describe S. giganteus, Figure 1D of Skinner et al. (Reference Skinner, Tenório, Penha and Soares2012) depicts S. tetraceros (Schmarda, 1861). Nygaard (Reference Nygaard2008) may have been studying S. polycerus (Schmarda, 1861), a species with spiral branchiae as well and tubes of about 5 mm across (ten Hove, Reference ten Hove1970; Hoeksema & van Moorsel, Reference Hoeksema, van Moorsel and Hoeksema2016). The latter is very common on all kinds of hard substrates in the Antilles, including Bonaire (ten Hove, unpublished data).

The second (mainly) tropical species (but see below), Spirobranchus tetraceros, is an Indo-Pacific invasive in the Mediterranean (Ben-Eliahu & ten Hove, Reference Ben-Eliahu and ten Hove1992), and has been reported from Turkey to inhabit artificial as well as natural hard substrata (Çinar, Reference Çinar2006; Çinar et al., Reference Çinar, Kurt and Dağli2014). Spirobranchus tetraceros was also reported to inhabit artificial substrate and become an extensive fouler in the Suez Canal (Shalla & Holt, Reference Shalla and Holt1999; Selim et al., Reference Selim, Abdel Naby, Gab-Alla and Ghobashy2005). Moreover, S. tetraceros seems to be an opportunistic taxon as it was reported to explosively overgrow corals immediately after a period of stress in the Persian Gulf (Samimi Namin et al., Reference Samimi Namin, Risk, Hoeksema, Zohari and Rezai2010). In view of its exceptionally large distribution, from temperate Australia to tropical regions all over the globe, S. tetraceros most probably is a species-complex. Despite the body of knowledge accumulated on substrate of Spirobranchus spp., for some species, such as S. gardineri Pixell, Reference Pixell1913, hardly any substrate preference data are available. For the typical form (his variant types 1 and 2 in the meantime have both been named S. richardsmithi Pillai, Reference Pillai2009) the genera Millepora Linnaeus, 1758 spp., Psammocora Dana, 1846 spp. and Stylocoeniella Yabe & Sugiyama, 1935 spp. have been mentioned by Smith (Reference Smith1985: 40).

The Gulf of Eilat is the northernmost coral reef, expanding northward from the Red Sea and the Indian Ocean. Eight nominal species of Spirobranchus have been recorded from the Red Sea and its northern Gulf of Eilat, of which seven from both areas (Table 1). The apparent absence of the Indo-West Pacific S. gaymardi from the Red Sea proper while occurring in its most northern gulf is probably due to insufficient collecting effort rather than reflecting a real difference.

Table 1. Spirobranchus species reported from the Red Sea and the Gulf of Eilat.

References: (1) Pixell (Reference Pixell1913), Mergner (Reference Mergner1979); (2) Gravier (Reference Gravier1906, as jousseaumei), Amoureux et al. (Reference Amoureux, Rullier and Fishelson1978, Figure 16); (3) Vine & Bailey-Brock (Reference Vine and Bailey-Brock1984, Figure 5B–G), Hassan (Reference Hassan1998); (4) Vine & Bailey-Brock (Reference Vine and Bailey-Brock1984); (5) ten Hove, personal observation; (6) Fauvel (Reference Fauvel1933); (7) Vine & Bailey-Brock (Reference Vine and Bailey-Brock1984); (8) Vine & Bailey-Brock (Reference Vine and Bailey-Brock1984), Hassan (Reference Hassan1998).

Table modified from ten Hove et al. (unpublished).

* Considered to be synonymous with S. corniculatus by Willette et al. (Reference Willette, Iñiguez, Kupriyanova, Starger, Vartman, Toha, Maralit and Barber2015).

P, published, identification confirmed; S, published, but ID has since been synonymized; U, previously unpublished; ?, published, questionable (Not yet verified).

However, species identification remains questionable as some taxa were identified based on too few specimens; other names were regarded to be synonyms by some authors. In order to examine the potential for habitat partitioning in Spirobranchus, our goal was to start resolving conflicting records about which species are found in the Eilat region and what hosts they inhabit. Thus, in this study we surveyed Spirobranchus spp. and their distribution on different substrates in the Gulf of Eilat.

MATERIALS AND METHODS

Collecting and preservation of specimens

The research was performed on the coral reef of Eilat, Gulf of Eilat, Red Sea, 29°33′N 34°57′E, Israel. A total of 189 specimens were collected by scuba from depths of 0.5–12 m (Israel Nature-Parks Authority [INPA] permit number 2014/40533). Seven substrate types were examined for Spirobranchus species (Table 2). Due to INPA limitations and to prevent damage to the reef, worms were sampled mainly from branching and encrusting corals and from artificial substrates such as pillars of a pier without coral association. After taking measurements (see below), a piece of the abdomen of each specimen was preserved in 100% ethanol and stored at −20°C for molecular analysis. The rest of each specimen was fixed in 4% formaldehyde (in SW) for 24 h, rinsed in filtered seawater, transferred to ethanol 70%, and stored at 4°C.

Table 2. Number of specimens of Spirobranchus spp. collected from different habitats.

Although we observed worms on massive coral genera such as Porites sp. Link, 1807 and Dipsastraea sp. Blainville, 1830, we did not sample them in order to minimize the damage to slow growing taxa.

Species identification

The Spirobranchus giganteus complex was tentatively resolved by Fiege & ten Hove (Reference Fiege and ten Hove1999, Figure 4), distinguishing 12 separate species mainly on the basis of the morphology of their opercula. However, the morphospecies of the Indo-West Pacific S. corniculatus complex in the strict sense as distinguished by Fiege & ten Hove, S. corniculatus (Grube, 1862), S. gaymardi (Quatrefages, 1865) and S. cruciger (Grube, 1862), are not distinguished by DNA (Willette et al., Reference Willette, Iñiguez, Kupriyanova, Starger, Vartman, Toha, Maralit and Barber2015).

For this study, identification of Spirobranchus taxa is based on the morphology of the operculum (as for instance in Willette et al., Reference Willette, Iñiguez, Kupriyanova, Starger, Vartman, Toha, Maralit and Barber2015; however see below), which is a hard (calcareous) cover sealing the opening of the tube when the worm retracts. The morphology of the operculum can change during growth and as a result its structure in juveniles can resemble the fully grown operculum of another species (ten Hove & Ben-Eliahu, Reference ten Hove and Ben-Eliahu2005). We concentrated our efforts on larger specimens, no longer subject to such ontogenetic changes. However, it is almost impossible to observe and differentiate between Spirobranchus taxa in situ because the operculum also functions as substrate for other organisms obscuring the opercular morphology (Figure 2). Overgrowth of opercula in serpulids is not uncommon, for example it was mentioned by Gambi (Reference Gambi1986) for Ditrupa arietina (O. F. Müller, 1776). Because identification in situ is difficult, and due to confusing and changing insights in the taxonomy of the larger Spirobranchus taxa, many studies followed Fauvel (Reference Fauvel1953) or Day (Reference Day1967) and applied their ‘widespread species’ concept of the nominal species Spirobranchus giganteus to material from Indo-West Pacific origin (Smith, Reference Smith and Hutchings1984; DeVantier et al., Reference DeVantier, Reichelt and Bradbury1986; Nishi & Nishihira, Reference Nishi and Nishihira1999; Floros et al., Reference Floros, Samways and Armstrong2005; Ben-Tzvi et al., Reference Ben-Tzvi, Einbinder and Brokovich2006; Rowley, Reference Rowley2008). However, ten Hove (Reference ten Hove1970), in a first attempt to unravel the ‘giganteus’ complex, drew the attention to the fact that there might be geographically more restricted taxa involved; he distinguished between two, possibly three subspecies: the Caribbean S. giganteus (Pallas, 1776), the Indo-West Pacific S. corniculatus and the Pacific Mid-American S. incrassatus Krøyer in Mörch, 1863. A further attempt was made by Fiege & ten Hove (Reference Fiege and ten Hove1999, especially figure 4), who mentioned 12 taxa which almost all previously had been included in Spirobranchus giganteus sensu latissimo.

Fig. 2. Operculum as substrate: Arrows show operculum with organisms on top. (A–C) operculum under binocular, (D–F) worms in situ. (A) Sponge; (B) Coral; (C) Spirorbis tube; (D–F) Unidentified organisms. Photos: O. Perry.

To facilitate identification, worms were collected and returned to the laboratory. We photographed each specimen using a Canon PowerShot G16 and examined each specimen using a Leica M165 FC binocular microscope and photographed it using a Leica DFC295 digital camera and LAS software. Worms' length was measured from the digital photographs, from base of radioles to end of pygidium, using image analysis ImageJ software (ver. 1.47v NIH). All specimens collected were categorized into groups (morphotypes) according to their opercular morphology, and were either compared to morphotype illustrations by Willette et al. (Reference Willette, Iñiguez, Kupriyanova, Starger, Vartman, Toha, Maralit and Barber2015, figure 2 – however, in this figure two morphotypes have been switched: figure 2a, b show Spirobranchus corniculatus,- figure 2c, d are of S. gaymardi) or attributed to morphologically better defined taxa such as S. tetraceros and S. gardineri. Unfamiliar morphologies which could not be ascribed to one of the above mentioned taxa were recorded as Spirobranchus sp. Corals were identified to genus level by their morphology using a coral fact sheets guide (http://coral.aims.gov.au/factsheet.jsp?speciesCode=0162). It is important to note that while taxonomy of the Spirobranchus corniculatus-complex is still not fully resolved, this paper is focusing on ecological interactions of morphotypes with their substrate.

Molecular work is currently being executed on the worms collected to underpin the taxonomy of the species.

Statistical analyses

We used contingency table and χ2 to test for association between substrate and identified morphotypes. To assess the effects of species identity and substrate type on worm length we used two-way analysis of variance. Preliminary observations revealed an apparent size difference between the two major morphotypes. Thus, in order to control for heritable differences between morphotypes, data were standardized within each species by subtracting the mean length and dividing by the standard deviation, to achieve mean of 0 and standard deviation of 1. All data analyses were carried out with R (R Development Core Team, 2012).

RESULTS

Morphological characterization

Specimens collected were divided into seven opercular morphology groups (Figure 3). Of the six species (eight nominal minus two confirmed synonyms, Spirobranchus corniculatus and S. gaymardi) reported from the Red Sea (Table 1) only five morphotypes (see below) were identified in this survey in the Gulf of Eilat. Among the 189 specimens collected, we identified the following taxa: S. corniculatus, S. cruciger, S. gaymardi (all three are considered part of the S. corniculatus complex s. str., see Fiege & ten Hove, Reference Fiege and ten Hove1999, Figure 4), S. gardineri and S. tetraceros. In addition, two not previously recognized opercular morphologies were observed among the collected samples. The seven groups of opercular morphology can be defined as follows: S. corniculatus (Figure 3A) has an oval (more or less egg-shaped) opercular plate with two laterodorsal spines arising from a short common base. These spines each have a small dorsal tine, a secondary spinule along the bend and are forked at the tip. Spirobranchus gaymardi (Figure 3B) differs from S. corniculatus by the larger dorsal spines and a small medioventral knob to a large forked medioventral spine, all arising from a short common base. The laterodorsal spines have well developed dorsal tines, meeting mid-dorsally. In addition the dorsal spines have one or two secondary spinules along the bend of the spine and are forked at the tip. Spirobranchus cruciger differs from S. gaymardi by the dorsal tines not expanded at their tips and not meeting mid-dorsally (Figure 3C).

Fig. 3. Seven opercular morphology groups of Spirobranchus spp. from Eilat (dorsal views). (A–C) S. corniculatus: (A) morphotype corniculatus s. str.; (B) morphotype gaymardi; (C) morphotype cruciger; (D) S. gardineri; (E) S. tetraceros; (F, G) Spirobranchus sp. Scale bar: 1 mm. Photos: O. Perry.

Spirobranchus gardineri (Figure 3D) has an oval opercular plate with one elongated almost central shaft with two dorsal and one forked midventral spines at the end, all pointing upwards. Spirobranchus tetraceros (Figure 3E) is recognized by an almost circular opercular plate with three pairs of antler-like spines arranged around the middle of the plate; each spine is forked at the tip (moreover, its radioles are arranged in two circles, not spirals; it is the only taxon here with anteriorly fringed peduncular wings, all others have smooth wings). Spirobranchus sp. two new morphotypes (Figure 3F, G) have a circular opercular plate with a pair of large dorsal antler-like spines arising from a short common base, each with a well developed dorsal tine (like those in morphotype gaymardi) well separated, and with two secondary spinules along the bend. There are two stout ventral spines, joined at their base, in some cases forked at their tips. A full description of this taxon is in preparation.

Species–substrate association

Nearly half of all specimens collected belong to the Spirobranchus corniculatus complex s. str. The complex was more abundantly found on corals than on artificial substrate (Table 2), but showed no preference for any particular coral species. Spirobranchus tetraceros and S. gardineri were both uncommon (N = 10, 5.3% and N = 5, 2.7%, respectively) and both were found exclusively in association with corals. Spirobranchus tetraceros was found on Pocillopora Lamarck, 1816 spp., Stylophora Schweigger, 1820 spp., Seriatopora Lamarck, 1816 spp. and Acropora Oken, 1815 spp., whereas S. gardineri was found exclusively on Cyphastrea Milne Edwards & Haime, 1848 spp., a new record in addition to the three genera mentioned by Smith (Reference Smith1985). Spirobranchus sp. was found mainly on artificial substrate (N = 51, 61.5%). Overall, the distribution of Spirobranchus taxa over substrate types differed significantly (χ218 = 295.1, P < 0.001; Figure 4). We only found sufficient numbers of the S. corniculatus complex and the undescribed morphotype (Spirobranchus sp.) for analyses of size. The average length of Spirobranchus sp. and S. corniculatus complex was 14.5 and 21 mm, respectively (Figure 5A).

Fig. 4. Relative abundance of Spirobranchus spp. in different substrates.

Fig. 5. The relationship between substrate and length in Spirobranchus associated with corals and artificial substrate. (A) Average length of Spirobranchus corniculatus complex (grey) and Spirobranchus sp. (black). (B) Standardized lengths of Spirobranchus corniculatus complex (grey) and Spirobranchus sp. (black) differed significantly as a function of substrate.

Spirobranchus sp. was significantly larger when settled on artificial substrate than on corals, whereas individuals of the S. corniculatus complex were larger on corals and smaller on artificial substrate (Figure 5B).

DISCUSSION

Only five of the eight nominal Spirobranchus species previously reported from the Gulf of Eilat were found in this study. Some apparent absences can be explained by misidentification in the past, revisions in taxonomy that have occurred in the intervening years, and difference in sampling effort. Spirobranchus latiscapus, for instance, generally has been reported from dredged material and deeper water, not sampled by us. We cannot overrule the possibility of species disappearance from the area as a result of anthropogenic causes or climate changes (Rilov, Reference Rilov2016). Given the relatively limited collection efforts in this study, it is premature to interpret our results as an indication that a particular species no longer appears in the Gulf of Eilat. Data on substrates used by several of the Spirobranchus spp. are missing or limited.

We found five specimens of S. gardineri in this study, all of them associated with live Cyphastrea. The low abundance of S. tetraceros in this study is slightly surprising, having been reported to be a very abundant Lessepsian migrant in the Suez Canal and Mediterranean (Selim et al., Reference Selim, Abdel Naby, Gab-Alla and Ghobashy2005). However, records from the Red Sea never mention such a massive occurrence (e.g. Ben-Eliahu & ten Hove, Reference Ben-Eliahu and ten Hove2011). Another interesting difference is that the species fouls artificial substrate in the Suez Canal and Alexandria Harbour, but was only found on corals in the present study. Either it has a large ecological plasticity, which is suggested by Samimi Namin et al. (Reference Samimi Namin, Risk, Hoeksema, Zohari and Rezai2010), or we are dealing with a complex of species only to be distinguished with genetics. All three morphotypes of the Spirobranchus corniculatus complex were identified in this study: S. corniculatus, S. cruciger and S. gaymardi. Contrary to previous records that S. corniculatus complex is an obligate symbiont of living corals (Nishi, Reference Nishi1996), this is the first record of S. corniculatus complex s. str. inhabiting an artificial substrate in addition to hermatypic corals. About 20% of these worms were found on artificial substrate in our study (Figure 4). It appears that the association with corals is not obligatory to the S. corniculatus complex s. str. in the Gulf of Eilat, all opercular morphologies belonging to the complex (thus S. corniculatus, cruciger and gaymardi) were found on corals as well as on artificial substrate. This may be attributed to differences in collecting efforts in previous studies or to differentiation in the Red Sea of the S. corniculatus complex. Genetic work (that has already started) may be able to shed light on this seemingly dissimilar behaviour. Two new and previously unrecognized morphologies were found which might represent a single new species. These forms showed a preference for artificial substrate, on which more than 60% had settled.

In S. giganteus s. str., the association with favourable substrate may influence fitness (Hunte et al., Reference Hunte, Marsden and Conlin1990b). In a laboratory experiment, Hunte et al. (Reference Hunte, Marsden and Conlin1990b) showed that worms that were found in their preferred habitat attained larger size. This appears to be supported by our study: worms of two species not only favoured different habitats, but they also apparently attained a larger size in their preferred habitats (S. corniculatus complex is more abundant and larger on corals; Spirobranchus sp. more abundant and larger on artificial substrate). Thus, worms of the S. corniculatus complex are not negatively affected by coral host defences, a finding consistent with the view that the worms may be commensal symbionts (Rowley, Reference Rowley2008). The opposite trend was found for Spirobranchus sp., preferring artificial substrates, making additional research on the relationships between different worm species and their various hosts and habitats necessary.

Spirobranchus worms sampled here showed high levels of plasticity in substrate selection. Worms were found on all six coral genera sampled, as well as on artificial substrate, though two species (S. gardineri, S. tetraceros) were only found on corals. The high level of plasticity in settling on a variety of substrates shown by Spirobranchus corniculatus s.str. as well as probably new species of Spirobranchus in the Gulf of Eilat may also explain the relative abundance of the genus in general.

ACKNOWLEDGEMENTS

The authors dedicate this paper to the memory of the late Dr Nechama Ben-Eliahu, a central figure in the Polychaeta world. We are grateful to her for sharing her vast knowledge with the authors. We would like to show our deep appreciation to Tamar Feldstein (Tel-Aviv University), Dan Perry (Arava Institute for Environmental studies) and Noga Stambler (Bar-Ilan University) for providing helpful comments on the manuscript and scientific guidance. We thank Muriel Dray, Eynav Cohen, Ellie Foran, Ayelet Hallakoun, Assaf Partzelan, Yaniv Shmuel, Ido Shefy, Dror Komet, Jessica Bellworthy, Yoav Balaban, Gabriella Sanka and Irena Kolesnikova (Interuniversity institute for Marine Sciences in Eilat) for assisting in the field and in the lab. Oren Levy and Noa Blecher-Simon (Bar-Ilan University) provided facilities and advice in lab work. Efrat Gavish and Ariel Chipman (Hebrew University of Jerusalem) kindly provided access to the National Natural History Collections and Bert W. Hoeksema (Naturalis Biodiversity Center) to their collections in Leiden. Assaf Zvuloni (The Israel Nature & Parks Authority) assisted with permits. We thank Elena Kupriyanova (Australian Museum Research Institute) for information on field and lab work as well as useful taxonomy discussions.

This research is part of the requirements for a PhD thesis for Orly Perry at Bar Ilan University.

FINANCIAL SUPPORT

The Interuniversity institute for Marine Sciences in Eilat provided logistics, diving and boating services. A Martin Fellowship enabled Orly Perry's visit to Naturalis Biodiversity Center, Leiden.

References

REFERENCES

Amoureux, L., Rullier, F. and Fishelson, L. (1978) Systématique et écologie d'annélides polychètes de la presqu'il du Sinai. Israel Journal of Zoology 27, 57163.Google Scholar
Ben-Eliahu, M.N. and ten Hove, H.A. (1992) Serpulids (Annelida: Polychaeta) along the Mediterranean coast of Israel – new population build-ups of Lessepsian migrants. Israel Journal of Zoology 38, 3553.Google Scholar
Ben-Eliahu, M.N. and ten Hove, H.A. (2011) Serpulidae (Annelida: Polychaeta) from the Suez Canal – from a Lessepsian migration perspective (a monograph). Zootaxa 2848, 1147.Google Scholar
Ben-Tzvi, O., Einbinder, S. and Brokovich, E. (2006) A beneficial association between a polychaete worm and a scleractinian coral? Coral Reefs 25, 9898.Google Scholar
Bowen, B.W., Rocha, L.A., Toonen, R.J. and Karl, S.A. (2013) The origins of tropical marine biodiversity. Trends in Ecology and Evolution 28, 359366.Google Scholar
Çinar, M.E. (2006) Serpulid species (Polychaeta: Serpulidae) from the Levantine coast of Turkey (eastern Mediterranean), with special emphasis on alien species. Aquatic Invasions 1, 223240.CrossRefGoogle Scholar
Çinar, M.E., Kurt, Ş.G. and Dağli, E. (2014) Checklist of Annelida from the coasts of Turkey. Turkish Journal of Zoology 38, 734764.CrossRefGoogle Scholar
Conlin, B.E. (1988) Aspects of habitat selection by a tropical Serpulid Polychaete Spirobranchus giganteus (Pallas). MSc, McGill University, Canada.Google Scholar
Dai, C.F. (1990) Interspecific competition in Taiwanese corals with special reference to interactions between alcyonaceans and scleractinians. Marine Ecology Progress Series 60, 291297.CrossRefGoogle Scholar
Dai, C.F. and Yang, H.P. (1995) Distribution of Spirobranchus giganteus corniculatus (Hove) on the coral reefs of Southern Taiwan. Zoological Studies 34, 117125.Google Scholar
Darwin, C.R. (1859) On the origin of species by natural selection, or the preservation of favoured races in the struggle for life. London: John Murray.CrossRefGoogle Scholar
Day, J.H. (1967) Part 2. Sedenteria. A monograph on the polychaeta of Southern Africa. London: The British Museum (Natural History).Google Scholar
DeVantier, L.M., Reichelt, R.E. and Bradbury, R.H. (1986) Does Spirobranchus giganteus protect host Porites from predation by Acanthaster planci: predator pressure as a mechanism of coevolution? Marine Ecology Progress Series 32, 307310.Google Scholar
Fauvel, P. (1933) Mission Robert Ph. Dollfus en Égypte. (Décembre 1927–Mars 1929). Mémoires présentés à l'Institut d’Égypte et publiés sous les auspices de sa Majesté Fouad Ier, Roi d’Égypte 21, 3183.Google Scholar
Fauvel, P. (1953) Annelida Polychaeta. The fauna of India, including Pakistan, Ceylon, Burma and Malaya. Allahabad: Indian Press.Google Scholar
Fiege, D. and ten Hove, H.A. (1999) Redescription of Spirobranchus gaymardi (Quatrefages, 1866) (Polychaeta: Serpulidae) from the Indo-Pacific with remarks on the Spirobranchus giganteus complex. Zoological Journal of the Linnean Society 126, 355364.Google Scholar
Floros, C.D., Samways, M.J. and Armstrong, B. (2005) Polychaete (Spirobranchus giganteus) loading on South African corals. Aquatic Conservation: Marine and Freshwater Ecosystems 15, 289298.Google Scholar
Frank, U. and ten Hove, H.A. (1992) In vitro exposure of Spirobranchus giganteus and S. tetraceros (Polychaeta, Serpulidae) to various turbidities; branchial morphologies and expression of filtering strategies? Oebalia 18, 4552.Google Scholar
Gambi, M.C. (1986) Ecological implications in the tube morphology and epibiota of a population of Ditrupa arietina (Polichaeta, Serpulidae). Rapports de la Commission Internationale pour l'Exploration scientifique de la Mer Méditerranée 30, 19.Google Scholar
Genin, A. and Karp, L. (1994) Effects of flow on competitive superiority in scleractinian corals. Limnology and Oceanography 39, 913924.Google Scholar
Gravier, C. (1906) Sur les Annélides Polychètes de la Mer Rouge (Serpulides). Bulletin du Museum Histoire Naturelle Paris 12, 110115.Google Scholar
Hassan, M. (1998) Modification of carbonate substrata by bioerosion and bioaccretion on coral reefs of the Red Sea. Thesis Kiel (1997), als Ms. gedrückt. Aachen: Shaker Verlag, vi + 126 pp.Google Scholar
Hoeksema, B.W. and ten Hove, H.A. (2016) The invasive sun coral Tubastraea coccinea hosting a native Christmas tree worm at Curaçao, Dutch Caribbean. Marine Biodiversity. doi: 10.1007/s12526-016-0472-7.Google Scholar
Hoeksema, B.W., van Beusekom, M., ten Hove, H.A., Ivanenko, V.N., van der Meij, S.E.T. and van Moorsel, G.W.N.M. (2016) Helioseris cucullata as a host coral at St. Eustatius, Dutch Caribbean. Marine Biodiversity. doi: 10.1007/s12526-016-0599-6.Google Scholar
Hoeksema, B.W. and van Moorsel, G.W.N.M. (2016) Stony corals of St. Eustatius. In Hoeksema, B.W. (ed.) Marine biodiversity survey of St. Eustatius, Dutch Caribbean, 2015. Leiden: Naturalis Biodiversity Centre, and Bennebroek: ANEMOON Foundation, pp. 3237.Google Scholar
Hoeksema, B.W., Wah Lau, Y. and ten Hove, H.A. (2015) Octocorals as secondary hosts for Christmas tree worms off Curaçao. Bulletin of Marine Science 91, 489490.Google Scholar
Hunte, W., Conlin, B.E. and Marsden, J.R. (1990a) Habitat selection in the tropical polychaete Spirobranchus giganteus 1. Distribution on corals. Marine Biology 104, 8792.Google Scholar
Hunte, W., Marsden, J.R. and Conlin, B.E. (1990b) Habitat selection in the tropical polychaete Spirobranchus giganteus 3. Effects of coral species on body size and body proportions. Marine Biology 104, 101107.CrossRefGoogle Scholar
Levesque, C., Juniper, K.S. and Marcus, J. (2003) Food resource partitioning and competition among alvinellid polychaetes of Juan de Fuca Ridge hydrothermal vents. Marine Ecology Progress Series 246, 173182.CrossRefGoogle Scholar
Linnaeus, C. (1767) Systema naturae per regna tria naturae, secundum classes, ordines, genera, species cum characteribus, differentiis, synonymis, locis. Ed.12, 1, 2, pp. 533–1327. Holmiae, Salvius L.Google Scholar
MacArthur, R. and Levins, R. (1964) Competition, habitat selection and character displacement in a patchy environment. Proceedings of the National Academy of Sciences USA 51, 12071210.CrossRefGoogle Scholar
Marsden, J.R. (1987) Coral preference behaviour by planktotrophic larvae of Spirobranchus giganteus corniculatus (Serpulidae: Polychaeta). Coral Reefs 6, 7174.CrossRefGoogle Scholar
Marsden, J.R. and Meeuwig, J. (1990) Preferences of planktotrophic larvae of the tropical serpulid Spirobranchus giganteus (Pallas) for exudates of corals from a Barbados reef. Journal of Experimental Marine Biology and Ecology 137, 95104.Google Scholar
Mergner, H. (1979) Quantitative ökologische analyse eines rifflagunenareals bei Aqaba (Golf von Aqaba, Rotes Meer). Helgoländer wissenschafliche Meeresuntersuchungen 32, 476507.CrossRefGoogle Scholar
Nishi, E. (1996) Serpulid polychaetes associated with living and dead corals at Okinawa Island, Southwest Japan. Publications of the Seto Marine Biological Laboratory 37, 305318.Google Scholar
Nishi, E. and Nishihira, M. (1996) Age-estimation of the Christmas tree worm Spirobranchus giganteus (Pomlychaeta, Serpulidae) living buried in the coral skeleton from the coral-growth band of the host coral. Fisheries Science 62, 400403.Google Scholar
Nishi, E. and Nishihira, M. (1999) Use of annual density banding to estimate longevity of infauna of massive corals. Fisheries Science 65, 4856.Google Scholar
Nygaard, L. (2008) Size distribution of Spirobranchus giganteus in Bonaire: is there a benefit of recruitment to live coral? Physis Journal of Marine Science 3, 2530.Google Scholar
Petitjean, S.E. and Myers, A.E. (2005) Age, characterization, and distribution of Spirobranchus giganteus on Paraiso reef. Epistimi 8, 14.Google Scholar
Pillai, T.G. (2009) Descriptions of new serpulid polychaetes from the Kimberleys of Australia and discussion of Australian and Indo-West Pacific species of Spirobranchus and superficially similar taxa. Records of the Australian Museum 61, 93199.Google Scholar
Pixell, H.L.M. (1913) Polychaeta of the Indian Ocean, together with some species from the Cape Verde Islands. Transactions of the Linnean Society (Zoology) London 16, 6992.Google Scholar
R Development Core Team (2012) R: a language and environment for statistical computing. Vienna: R Foundation for Statistical Computing.Google Scholar
Read, G., Fiege, D. and Bellan, G. (2016) Spirobranchus Blainville, 1818. In Read, G. and Fauchald, K. (eds) World Polychaeta database. World Register of Marine Species. http://www.marinespecies.org/aphia.php?p=taxdetails&id=129582.Google Scholar
Rilov, G. (2016) Multi-species collapses at the warm edge of a warming sea. Scientific Reports 6, 36897.Google Scholar
Rowley, S.J. (2008) A critical evaluation of the symbiotic association between tropical tube-dwelling Polychaetes and their hermatypic coral hosts, with a focus on Spirobranchus giganteus (Pallas, 1766). The Plymouth Student Scientist 1, 335353.Google Scholar
Rzhavsky, A.V., Kupriyanova, E.K., Sikorski, A.V. and Dahle, S. (2014) Calcareous tubeworms (Polychaeta, Serpulidae) of the Arctic Ocean. Moscow: KMK Scientific Press.Google Scholar
Samimi Namin, K., Risk, M.J., Hoeksema, B.W., Zohari, Z. and Rezai, H. (2010) Coral mortality and serpulid infestations associated with red tide, in the Persian Gulf. Coral Reefs 29, 509.CrossRefGoogle Scholar
Selim, S.A., Abdel Naby, F., Gab-Alla, A. and Ghobashy, A. (2005) Gametogenesis and spawning of Spirobranchus tetraceros (Polychaeta, Serpulidae) in Abu Kir Bay, Egypt. Mediterranean Marine Science 6, 8997.Google Scholar
Shalla, S.H. and Holt, T.J. (1999) The Lessepsian migrant Pomatoleios kraussii (Annelida, Polychateta) – recent formation of dense aggregations in Lake Timsah and the Bitter Lakes (Suez Canal, Egypt). Egyptian Journal of Biology 1, 133137.Google Scholar
Skinner, L.F., Tenório, A.A., Penha, F.L. and Soares, D.C. (2012) First record of Spirobranchus giganteus (Pallas, 1766) (Polychaeta, Serpulidae) on Southeastern Brazillian coast: new biofouler and free to live without corals? Pan-American Journal of Aquatic Science 7, 117124.Google Scholar
Smith, R.S. (1984) Development and settling of Spirobranchus giganteus (Polychaeta; Serpulidae). In Hutchings, P.A. (ed.) Proceedings of the First International Polychaete Conference. Sydney: The Linnean Society of New South Wales, pp. 461483.Google Scholar
Smith, R.S. (1985) Photoreceptors of serpulid polychaetes. PhD thesis, James Cook University, North Queensland.Google Scholar
Strathmann, R.R., Cameron, R.A. and Strathmann, M.F. (1984) Spirobranchus giganteus (Pallas) breaks a rule for suspension-feeders. Journal of Experimental Marine Biology and Ecology 79, 245249.Google Scholar
ten Hove, H.A. (1970) Serpulinae (Polychaeta) from the Caribbean: 1 – The genus Spirobranchus. Studies on the Fauna of Curaçao and other Caribbean Islands 32, 157.Google Scholar
ten Hove, H.A. and Ben-Eliahu, M.N. (2005) On the identity of Hydroides priscus Pillai 1971 – Taxonomic confusion due to ontogeny in some serpulid genera (Annelida: Polychaeta: Serpulidae). Senckenbergiana Biologica 85, 119.Google Scholar
Vine, J.P. and Bailey-Brock, H.J. (1984) Taxonomy and ecology of coral reef tube worms (Serpulidae, Spirorbidae) in the Sudanese Red Sea. Zoological Journal of the Linnean Society 80, 135156.Google Scholar
Willette, D.A., Iñiguez, A.R., Kupriyanova, E.K., Starger, C.J., Vartman, T., Toha, A.H., Maralit, B.A. and Barber, P.H. (2015) Christmas tree worms of Indo-Pacific coral reefs: untangling the Spirobranchus corniculatus (Grube, 1862) complex. Coral Reefs 34, 899904.Google Scholar
Figure 0

Fig. 1. Different substrates colonized by Spirobranchus, reef of Eilat, Israel. (A) Spirobranchus gardineri associated with Cyphastrea sp. (B) S. gaymardi associated with Pocillopora sp. (C) Spirobranchus sp. associated with Stylophora sp. (D) S. cruciger associated with Millepora sp. (E) S. corniculatus associated with Acropora sp. (F) Spirobranchus sp. on artificial substrate. Arrow: operculum. (A, D, F) Photos: O. Perry; (B, C, E) photos: A. Hallakoun.

Figure 1

Table 1. Spirobranchus species reported from the Red Sea and the Gulf of Eilat.

Figure 2

Table 2. Number of specimens of Spirobranchus spp. collected from different habitats.

Figure 3

Fig. 2. Operculum as substrate: Arrows show operculum with organisms on top. (A–C) operculum under binocular, (D–F) worms in situ. (A) Sponge; (B) Coral; (C) Spirorbis tube; (D–F) Unidentified organisms. Photos: O. Perry.

Figure 4

Fig. 3. Seven opercular morphology groups of Spirobranchus spp. from Eilat (dorsal views). (A–C) S. corniculatus: (A) morphotype corniculatus s. str.; (B) morphotype gaymardi; (C) morphotype cruciger; (D) S. gardineri; (E) S. tetraceros; (F, G) Spirobranchus sp. Scale bar: 1 mm. Photos: O. Perry.

Figure 5

Fig. 4. Relative abundance of Spirobranchus spp. in different substrates.

Figure 6

Fig. 5. The relationship between substrate and length in Spirobranchus associated with corals and artificial substrate. (A) Average length of Spirobranchus corniculatus complex (grey) and Spirobranchus sp. (black). (B) Standardized lengths of Spirobranchus corniculatus complex (grey) and Spirobranchus sp. (black) differed significantly as a function of substrate.