<<

Zoological Journal of the Linnean Society

Towards the identifica tion of the ommastrephid paralarvae (: Cephalopoda): morphological description of three and a key to the Northeast Atlantic species

For Review Only Journal: Zoological Journal of the Linnean Society

Manuscript ID ZOJ-03-2016-2516.R2

Manuscript Type: Original Article

morphological comparison < Anatomy, morphology < Anatomy, larvae < Keywords: Development, North Atlantic < Geography, Mollusca < Taxa, Cephalopoda, identification key < Taxonomy, intraspecific variation < Taxonomy

Page 1 of 31 Zoological Journal of the Linnean Society

1 2 3 1 Abstract 4 5 2 Oceanic of the are an important fishing resource worldwide. 6 3 Although cumulative knowledge exists on their subadult and adult forms, little is known about 7 4 their young stages. Their hatchlings are among the smaller paralarvae. They are 8 5 characterized by the fusion of their tentacles into a proboscis and are very difficult to identify to 9 species level, especially in areas where more than one species coexist . Seven species are found 10 6 11 7 in the NE Atlantic. In this study, mature oocytes of coindetii , sagittatus and 12 8 were fertilized in vitro to obtain and describe hatchlings. Full descriptions 13 9 based on morphometric characters, chromatophore patterns, skin sculpture, and the structure of 14 10 proboscis suckers are provided based on live specimens. This information was combined with 15 11 previous descriptions of paralarvae , not necessarily based on DNA or known parentage, from 16 17 12 four other ommastrephid species distributed in the same area and a dichotomous key was 18 13 developed for theFor identification Review of paralarvae of the NE OnlyAtlantic. The most useful taxonomic 19 14 characters were: the relative size of the lateral and medial suckers of the proboscis , the 20 15 presence/absence of , and the arrangement of pegs on the proboscis suckers. This 21 16 key was successfully used to identify wild collected rhynchoteuthion paralarvae from the NE 22 17 Atlantic. Reliable identification of wild paralarvae can foster a better understanding of the 23 24 18 population dynamics and life cycles of ommastrephid squids. 25 26 19 27 Keywords: Morphology, NE Atlantic, Mediterranean Sea, marine taxonomy, in vitro 28 20 29 21 fertilization, Scanning Electron Microscopy. 30 31 22 32 33 23 34 35 24 36 37 25 38 39 26 40 41 27 42 43 28 44 45 29 46

47 30 48 31 49 50 32 51 52 33 53 54 34 55 56 35 57 58 59 60 1

Zoological Journal of the Linnean Society Page 2 of 31

1 2 3 36 Introduction 4 5 37 Ommastrephid squids are distributed in all the world´s oceans and their rapid growth and 6 38 abundance make them the most important cephalopod fishery resource (Arkhipkin et al., 2015). 7 39 Knowledge of early life stages is essential for understanding ecology and life cycles as well as 8 40 for assessment of fisheries. These oceanic squids are important prey and predators, occupying a 9 wide range of trophic levels in marine pelagic food webs (Coll et al., 2013). They are also 10 41 11 42 dominant prey in the diet of many top fish predators (Logan et al., 2013) as well as seabirds and 12 43 marine mammals (Boyle & Rodhouse, 2005). Despite the cumulative knowledge on the biology 13 44 of subadult and adult forms of several species of this family, present knowledge of early life 14 45 stages is fragmentary and limited. Hatchlings of this family are among the smaller of the 15 46 and show a characteristic morphology with tentacles fused into a proboscis. These 16 17 47 paralarvae are known as rhynchoteuthion. Identification of the early stages of several species is 18 48 considerably difficultFor (Nesis, 1979;Review Sweeney et al., 1992), Only especially in areas where more than 19 49 one species coexists (Gilly et al., 2006; RamosCastillejos et al., 2010). Based on wildcollected 20 50 ommastrephid planktonic stages from around the world, Nesis (1979) stated that it is possible to 21 51 identify rhynchoteuthion paralarvae to , but not species level. Nine of the 22 accepted 22 52 ommastrephid species (Jereb & Roper, 2010) occur in the N Atlantic Ocean. In the NE Atlantic 23 24 53 Ocean, seven ommastrephid species can be found: (Verany, 1839), Todaropsis 25 54 eblanae (Ball, 1841), Todarodes sagittatus (Lamarck, 1798), Ommastrephes bartramii 26 55 (Lesueur, 1821), pteropus (Steenstrup, 1855), pelagica (Bosc, 1802) 27 56 and antillarum Adam, 1957. Besides these , I. illecebrosus (Lesueur, 1821) and I. 28 57 oxygonius Roper, Lu & Mangold, 1969 occur in the NW Atlantic. Previous works have dealt 29 58 with the morphology of the paralarvae of I. illecebrosus (e. g. Roper & Lu, 1979; O´Dor et al., 30 31 59 1982). However, the taxonomic status of the NW Atlantic Illex spp. populations and the extent 32 60 of their distribution range are not resolved. Even adult individuals are difficult to identify by 33 61 morphological characters when more than one species occurs (Carlini, Kunkle & Vecchione, 34 62 2006). Thus, NW I. illecebrosus and I. oxygonius were not included in this work. 35 36 63 The rhynchoteuthion paralarvae of O. bartramii (Young & Hirota, 1990; Watanabe et al., 1996; 37 64 Vijai, Sakai & Sakurai (2015) and I. coindetii (Boletzky, Rowe & Aroles, 1973; Villanueva et 38 65 al., 2011) have been studied from hatchlings obtained during laboratory experiments. The 39 40 66 morphology of O. bartramii paralarvae from the NW Pacific was studied from in vitro 41 67 fertilizations (Watanabe et al., 1996; Vijai et al., 2015) and the morphology of some paralarvae 42 68 collected in the wild assigned to this species (Young & Hirota, 1990) is consistent with the 43 69 chromatophore pattern reported in both types of studies. Only descriptions of H. pelagica , O. 44 70 antillarum and S. pteropus paralarvae collected in the wild are available (Harman & Young, 45 71 1985; Sweeney et al., 1992). Nesis (1979) provided a description of wildcollected individuals 46 47 72 of T. sagittatus , T. angolensis Adam, 1962 and T. pacificus (Steenstrup, 1880). However, he 48 73 merged the characters of the three species into a single generic description and the specific 49 74 characters of T. sagittatus are not available. For I. coindetii , complete descriptions of the 50 75 morphology and chromatophore pattern of the paralarvae are lacking. Moreover, the hatchlings 51 76 of T. eblanae are undescribed. The current situation makes species identification of wild 52 77 rhynchoteuthion paralarvae from the NE Atlantic nearly impossible using morphological 53 54 78 characters (e. g. Zaragoza et al., 2015). This strongly limits the study of the ecology and biology 55 79 of rhynchoteuthions from the NE Atlantic (e.g. Moreno et al., 2009). 56 57 80 In to address this problem, we used in vitro fertilization methods to obtain live paralarvae 58 81 of the three most fished NE Atlantic ommastrephid species ( I. coindetii , T. sagittatus and T. 59 60 2

Page 3 of 31 Zoological Journal of the Linnean Society

1 2 3 82 eblanae ) and provide detailed descriptions of their morphology and chromatophore pattern. The 4 83 available knowledge on the morphology of rhynchoteuthions of the other four NE Atlantic 5 84 species ( O. bartramii , S. pteropus , H. pelagica and O. antillarum ) was reviewed from the 6 85 literature. Moreover, a key for the identification of the ommastrephid paralarvae from the NE 7 Atlantic is also provided, aiming to offer a tool for the study of the biology and population 8 86 9 87 dynamics of paralarval stages of ommastrephid squids. 10 11 88 Material and methods 12 13 89 Obtaining paralarvae through in vitro fertilizations 14 15 90 Adult squids of I. coindetii , T. sagittatus and T. eblanae were captured by local bottom trawlers 16 91 from Barcelona and Vilanova i la Geltrú, NW Mediterranean Sea, between May 2010 and April 17 92 2015. Illex coindetii and T. eblanae were captured from 120 to 350 m depth, and T. sagittatus 18 93 from 300 and 400For m depth. SpecialReview care was taken in selectingOnly the freshest squids captured 19 94 during the latest trawl of the day. Selected individuals were placed on crushed ice covered by a 20 95 plastic film and transported to the laboratory. Mature females with oocytes in oviducts (stage V 21 22 96 VI according to Brunetti, 1990) were selected for in vitro fertilizations following the general 23 97 methodology described by Villanueva et al. (2012) with minor modifications. In short, for I. 24 98 coindetti , the sperm source used was the bulbs of spermatangia attached to the internal of 25 99 the female; for T. eblanae spermatophores of the Needham´s sac and the vas deferens from male 26 100 individuals were used; for T. sagittatus , sperm from spermatophores or seminal receptacles and 27 spermatangia was used depending on availability. The oviducal glands used to make the 28 101 29 102 oviducal jelly were obtained from mature females of the same species collected previously. 30 103 Freezedried oviducal gland powder was stored at 80ºC until use. The fertilization percentage 31 104 of each female used during the experiments was estimated one day after fertilization, by 32 105 counting the number of fertilized eggs from a sample of 400500 eggs (Table S1). When more 33 106 than one female was used in an experiment, the mean fertilization rate of the experiment was 34 35 107 calculated and is indicated in Table S1. Sterile 60mm diameter polystyrene Petri dishes each 36 108 containing 2050 fertilized eggs were maintained in the dark at 15, 17 or 21 ºC using incubators 37 109 (see Table S1). Since vertical distribution of egg masses of the species studied is unknown, 38 110 temperature conditions during natural embryonic development remain uncertain. Therefore, the 39 111 temperatures chosen for egg incubation in the present study were based on the temperature 40 112 ranges of midwater layers in the Mediterranean Sea (Brasseur et al., 1996), where the egg 41 1 42 113 masses are expected to occur. Throughout the experiment, 25 mg l of two antibiotics, 43 114 ampicillin and streptomycin, were added to the filtered seawater (FSW) (Staaf et al., 2008). The 44 115 FSW with antibiotics was replaced daily using a binocular microscope and sterile plastic 45 116 pipettes. Dead embryos and those with abnormalities were removed and counted daily to 46 117 determine the survival rates until the hatchling stage. 47 48 118 In all the experiments, the paralarvae were cared for and euthanized with the ethical methods in 49 119 accordance with the European Union Directive 2010/63/EU. Paralarvae were anaesthetized 50 51 120 adding drops of 70% ethanol to Petri dishes containing approximately 12 ml of FSW and 52 121 overdoses of anaesthesia were used to euthanize them. The initial stages of anaesthesia started 53 122 very gradually, adding only a few drops of ethanol to the Petri dish containing the paralarvae 54 123 over a period of 1530 min, aiming to avoid signals of irritation such as body contractions or ink 55 124 ejection. 56 57 125 Observations of live paralarvae 58 59 60 3

Zoological Journal of the Linnean Society Page 4 of 31

1 2 3 126 For accurate classification of embryonic stages, the scheme and definitions published for Illex 4 127 argentinus (Castellanos, 1960) (Sakai et al., 1998) and Todarodes pacificus (Watanabe et al., 5 128 1996) were directly applied to I. coindetii and T. sagittatus , respectively. The T. pacificus scale 6 129 was also adapted for T. eblanae . To avoid confusion between supposedly premature hatched 7 individuals and the expected normal hatching, hatchlings are defined here as individuals with a 8 130 9 131 welldeveloped ink sac, extensible proboscis and functional fins with a fin width nearly equal to 10 132 the head width. These developmental criteria can be found at stage XXX for the genus Illex 11 133 (Sakai et al., 1998; Villanueva et al., 2011) and stage XXXII for Todarodes and Todaropsis (see 12 134 Discussion). The descriptions were based on individuals of these stages. 13 14 135 The morphological description was based on measurements of several morphometric characters. 15 136 These measures were defined according to Roper & Voss (1983) as the ventral mantle length 16 17 137 (VML), the dorsal mantle length (DML), the total length (TL), the head length (HL), the head 18 138 width (HW), the Foreye diameter Review(ED), the funnel length (FuL), Only and the length of the second pair 19 139 of arms (AIIL). Two other characters were added to the morphometric descriptions. Since the 20 140 proboscis length was observed to change according to its contraction state, the total length 21 141 without the proboscis (TL w P) was also measured and was defined as the length of the 22 142 from the posterior tip of the mantle to the tip of the arm I. The proboscis width at the 23 24 143 base (PW) was defined as the maximum width of the proboscis at its base. The proboscis length 25 144 has been used as a taxonomic character (e. g. Diekmann, Piatkowski & Schneider, 2002). 26 145 However, this character varies highly with contraction state (Nesis, 1979; Sweeney et al., 1992 ; 27 146 Staaf et al., 2008) and usually changes throughout the ontogeny (Shea, 2005). Harman & Young 28 147 (1985) provided the proboscis length and the proboscis index of hawaiiensis 29 148 (Berry, 1912), H. pelagica and Sthenoteuthis oualaniensis (Lesson, 18301831). Although they 30 31 149 considered N. hawaiiensis to have a “typically short” proboscis and S. oualaniensis often with a 32 150 “very elongate” proboscis, both the proboscis length (Harman & Young, 1985: Fig. 2) and their 33 151 proboscis index overlap. Both characters also overlap with H. pelagica , as confirmed in other 34 152 posterior references (Sweeney et al., 1992; Diekmann et al., 2002). Thus, this character is not 35 153 very instructive and was not considered in our study. When resting on the Petri dishes, the 36 154 majority of the embryos or paralarvae lay on the dorsal surface of the body; thus, it was not 37 38 155 possible to measure the DML in many specimens. For this reason, the ratio between each 39 156 morphometric parameter and body size was obtained using the VML instead of the DML. The 40 157 terminology used for discriminating rhynchoteuthion types was: Type A, the ratio of sucker 41 158 sizes is greater than 2:1 (lateral suckers 200% or greater in size than the medial suckers), Type 42 159 B, this ratio is between 1.1:1 and 1.9:1 (lateral suckers larger than the medial suckers but below 43 160 200%) and Type C, this ratio is 1:1 (there is no size difference among the proboscis suckers). 44 45 161 These three categories are consistent with the proboscis sucker proportions of ommastrephid 46 162 paralarvae A, B, and C treated in Roper & Lu (1979). 47 48 163 For the description of the chromatophore pattern, the dorsal and ventral surfaces were depicted 49 164 in a schematic drawing of a rhynchoteuthion. Lateral views were only considered for the 50 165 description when hatchling chromatophores were visible only from the side view. For I. 51 166 coindetii and T. sagittatus , the lateral view was not included, since the lateral chromatophore 52 167 pattern is the sum of the dorsolateral and ventrolateral chromatophores. However, T. eblanae 53 54 168 has true lateral chromatophores in the midline of the lateral surface of the mantle, which are not 55 169 visible from dorsal or ventral views. The chromatophores of the head and mantle were assigned 56 170 to rows in an anteroposterior axis. For example, for the ventral mantle, the pattern 4 + 2 + 3 + 4 57 171 + 1 + 2 means: 4 chromatophores in the anterior margin of the mantle, 2 in the second row, 3 in 58 59 60 4

Page 5 of 31 Zoological Journal of the Linnean Society

1 2 3 172 the third row, 4 in the fourth row, 1 in the fifth row and 2 near the posterior tip of the mantle. 4 173 The mode of the pattern of each row was considered the most representative of the species and a 5 174 schematic drawing depicting this pattern is provided for each species. 6 7 175 Observations under scanning electron microscopy (SEM) 8 9 176 Paralarvae were euthanized with an overdose of anaesthesia prior to fixation. To avoid body 10 177 contraction of the paralarvae it is important to add the anaesthetic gradually before killing them, 11 178 starting anaesthesia with a few drops of ethanol. Once the three hearts stopped beating, the 12 13 179 individuals were fixed in 2.5% glutaraldehyde in seawater for 2448 h, washed in seawater 14 180 followed by dehydration in an increasing concentration of ethanol (20, 30 and 50%) and stored 15 181 in 70% ethanol in the dark at 4 °C. At the beginning of SEM preparation, the samples were 16 182 again dehydrated in an increasing concentration of ethanol (80, 90, and 96%) until they were 17 183 saturated in absolute ethanol. Each ethanol bath lasted 10 min. After complete dehydration in 18 the ethanol series,For the samples Review were dried to a critical point Only using CO as the transition liquid. 19 184 2 20 185 After the drying stage, samples were mounted on stubs with doublesided conductive sticky tape 21 186 to place them in the preferred position. The mounted samples were sputter coated with gold– 22 187 palladium. Finally, the samples were observed using a scanning electron microscope. 23 24 188 The general morphology and the number and arrangement of pegs in the proboscis and arm 25 189 suckers were examined. Nomenclature of proboscis suckers was as follows . The suckers were 26 190 named from 1 to 4 of each hemiproboscis tip, with 1 the most dorsal and 4 the most ventral. 27 Each side of the proboscis tip was named as R (right) or L (left), according to its position on the 28 191 29 192 anteroposterior axis of the body. For example: proboscis sucker R1 is the most dorsal sucker 30 193 from the right part of the proboscis tip. Suckers 2 corresponded to lateral suckers and suckers 1, 31 194 3 and 4 were medial suckers. Newly hatched rhynchoteuthion paralarvae only have one sucker 32 195 on arms I and II. Thus, no special nomenclature was necessary to designate each sucker. 33 34 196 Characters from both SEM and observations of live paralarvae were used to develop a 35 197 dichotomous key to facilitate the identification of the seven NE Atlantic rhynchoteuthion 36 37 198 paralarvae species. Morphological characters of O. bartramii, H. pelagica , S. pteropus and O. 38 199 antillarum were obtained from the literature (Harman & Young, 1985; Sweeney et al., 1992; 39 200 Young & Hirota, 1990; Sakurai et al., 1995; Diekmann et al., 2002; Vijai et al., 2015). It should 40 201 be noted that descriptions of the paralarvae of H. pelagica , S. pteropus and O. antillarum were 41 202 not confirmed by DNA or known parentage in these works . 42 43 203 Wild rhynchoteuthion paralarvae samples 44 45 204 Wild rhynchoteuthion paralarvae (n = 16) from the study area were collected using zooplankton 46 205 tows during the oceanic cruise LLUÇ3 (Palomera, Olivar & MoralesNin, 2005) in the summer 47 of 1999, NW Mediterranean; four oceanographic cruises conducted in the summers of 2003 and 48 206 49 207 2004 under the CACO research project (Sabatés et al., 2009), NW Mediterranean; and another 50 208 cruise near the Canary Islands in the spring of 2015 under the MAFIA research project using the 51 209 fishing net described by Meillat (2012). An additional paralarva collected during the project 52 210 FishJelly with a Bongo net with a 40 cm diameter opening and a mesh size of 300 µm in the 53 211 autumn of 2014 was also analysed from the NW Mediterranean. These zooplankton samples 54 55 212 were fixed in 5% formaldehyde buffered with sodium tetraborate. Ommastrephid paralarvae 56 213 were distinguished from the other cephalopod paralarvae by the presence of the proboscis. The 57 214 general morphology of the fixed specimens (chromatophores, size, number of arms, etc.) was 58 215 examined under the stereomicroscope prior to observation under SEM as described previously. 59 60 5

Zoological Journal of the Linnean Society Page 6 of 31

1 2 3 216 The specimen from the MAFIA cruise was frozen and measurements were taken after 4 217 defrosting. All of these paralarvae were identified using the dichotomous key developed in the 5 218 present study. 6 7 219 Results 8 9 220 Morphological description of the rhynchoteuthion hatchlings from the NE Atlantic . 10 11 221 Morphology of rhynchoteuthion paralarvae is very similar among species : hatchlings usually 12 222 have only one sucker on both pairs of arms I and II, pair IV is a protuberance without suckers 13 223 and pair III is totally undeveloped . Morphometric measurements and indices for each species 14 15 224 are shown in Table 1. Morphometric comparisons between species were performed based on the 16 225 indices, rather than on the raw measures. A full description of the chromatophore pattern of 17 226 each species is provided in Table 2. The rhynchoteuthion species Type in relation to ratio of 18 227 proboscis sucker Forsize and the descriptionReview of the proboscis Onlyand arm sucker pegs is summarized in 19 228 Table 3. 20 21 229 Illex coindetii hatchlings 22 23 230 The general morphology of the hatchling (Fig. 1ad, 3ac) of this species does not diverge from 24 231 that described for the congeneric I. argentinus (Sakai et al., 1998). The mean VML is 1.41 ± 25 26 232 0.15 mm. On the head, there are 2 ventral chromatophores, one below each eye, and dorsally 27 233 there are two rows of chromatophores, the first formed by a single chromatophore anterior to the 28 234 eyes and the second formed by 3 chromatophores at the base of the head. In some individuals, 29 235 small dark brown pigmented dots appear on the tips of the arms and proboscis (Figs. 1b, d and 30 236 3b). Although uncommon, some individuals showed head chromatophores arranged 31 237 asymmetrically, generally placed on the lateral sides of the head. On the mantle, there are up to 32 33 238 6 rows ventrally, distributed as follows: 4 + 2 + 4 + 3+ 3+ 2; dorsally, there are up to 5 rows: 2 34 239 + 3 + 3 + 0 + 1. The 8 suckers are of similar size (35.1 ± 3.9 µm) (Type C rhynchoteuthion, Fig. 35 240 2a) and have only one row of pegs (12.9 ± 1.4) (Table 3, Fig. 2c), but additional asymmetrically 36 241 distributed pegs can be found outside of this row. The arm suckers measure 40.2 ± 2.3 µm in 37 242 diameter and bear two rows of pegs (Fig. 2b), the internal one with 12.1 ± 1.0 pegs, the external 38 one with 12.2 ± 1.4. A few pegs can appear externally to the external row (Fig. 2b). The skin is 39 243 40 244 smooth and does not have any special sculpture (Fig. 2d). When compared with the other 41 245 paralarvae described here (Table 1), the FuLI and PWI are larger, which indicates that the 42 246 funnel is comparatively larger than in the other two species (although the range of this index 43 247 overlaps with that of T. eblanae ); and the proboscis is wider, although some overlap exists. 44 45 248 Remarks: Moreno (2008: Fig. 4.18) tentatively identified rhynchoteuthion paralarvae based on 46 249 the drawings of Salman, Katagan & Benli (2003, see below for more details) and among these, 47 some were identified as Illex specimens. Her pictures clearly show a Type C rhynchoteuthion 48 250 49 251 without ocular or intestinal photophores. Thus, these specimens were either I. coindetii or T. 50 252 eblanae . However, the identification of these specimens could not be confirmed from the 51 253 available description and pictures. 52 53 254 Roura (2013) succeeded in molecularly identifying one I. coindetii paralarva out of the 15 54 255 barcoded from Cape Silleiro (northwestern coast of the Iberian Peninsula). However, he did not 55 256 provide any morphological description of these paralarvae. 56 57 257 Todarodes sagittatus hatchlings 58 59 60 6

Page 7 of 31 Zoological Journal of the Linnean Society

1 2 3 258 The general morphology of this hatchling (Fig. 1eh, 3df) is similar to the congeneric T. 4 259 pacificus (Watanabe et al., 1996; Puneeta et al., 2015). The mean VML is 1.64 ± 0.12 mm. On 5 260 the head, there are no ventral chromatophores. Dorsally there are two rows of chromatophores, 6 261 the first formed by a single chromatophore anterior to the eyes and the second formed by 3 7 chromatophores at the base of the head. On the mantle, there are up to 4 rows ventrally, 8 262 9 263 distributed as follows: 2 + 6 + 5 + 2; dorsally, there are up to 5 rows, with the following 10 264 configuration: 3 + 1 + 4 + 0 + 0. The lateral proboscis suckers are larger in size (41.2 ± 8.0 µm) 11 265 than the medial suckers (35.0 ± 6.7 µm) (Type B rhynchoteuthion, Fig. 2e). The ratio between 12 266 the size of the sucker diameter of the lateral suckers and the medial suckers is 1.2:1. Two rows 13 267 of pegs are found (Fig. 2g), the internal row with 18.0 ± 3 and the external with 17.0 ± 3 pegs. 14 15 268 There are no differences in the number or arrangement of the sucker pegs between lateral and 16 269 medial suckers. All pegs belonged to either of these two rows. The skin of this species shows a 17 270 hexagonlike structure under the stereomicroscope (Fig. 2h). The FuLI (Table 1) is 18 271 comparatively smallerFor than in I.Review coindetii and T. eblanae , showingOnly a shorter funnel. 19 20 272 Remarks: Salman et al. (2003: Fig. 7) identified 2 rhynchoteuthions from the Aegean Sea as I. 21 273 coindetii. However, the drawing of the proboscis tip clearly shows a type B rhynchoteuthion 22 274 paralarva. The value of the ratio of sucker sizes is 1.4:1. No photophores were drawn and H. 23 24 275 pelagica and O. antillarum have never been found in Mediterranean waters (Jereb & Roper, 25 276 2010), excluding the other Type B paralarva from the NE Atlantic area. Although the ratio of 26 277 sucker sizes does not fit accurately with the value obtained here for T. sagittatus (1.2:1), we 27 278 considered these two paralarvae as members of this species, rather than I. coindetii . Possible 28 279 sources of this variation could be: a) differences related to the developmental state, since those 29 280 drawn for Salman et al. (2003) show more than one sucker on each arm and are of a larger size 30 31 281 (2.53.5 mm ML); b) intraspecific or regional variation; c) differences related to the fixation 32 282 procedure, since plankton samples usually are not anaesthetized before fixation and some 33 283 contraction is expected to occur; d) lower levels of accuracy in the measurements from the 34 284 drawings. 35 36 285 Moreno (2008) described a rhynchoteuthion paralarva with lateral suckers larger than the medial 37 286 ones in Atlantic Iberian waters. It is not possible to take measurements from the pictures 38 287 (Moreno, 2008: Fig. 4.22c) due to the orientation of the . No evidence of eye photophores 39 40 288 is found based on the pictures. However, the chromatophore pattern of both dorsal and ventral 41 289 views of the head and mantle is visible. The chromatophore pattern of the ventral surface of the 42 290 mantle is 4 + 4 + 4 + 1 + 1 and for the dorsal surface 3 + 4 + 3 + 2. The dorsal chromatophore 43 291 pattern of the head is not fully visible, but in the ventral view no chromatophores appear. The 44 292 chromatophore pattern of this individual is consistent with that described for T. sagittatus in this 45 293 work, especially regarding the absence of ventral chromatophores on the head in the hatchlings 46 47 294 of this species , which is of high taxonomic significance . 48 49 295 An 18 mm TL juvenile individual assigned to T. sagittatus sampled in Sicily (Central 50 296 Mediterranean Sea) has been documented (Piatkowski et al., 2015: Fig. 16.4). The absence of 51 297 chromatophores on the ventral surface of the head is remarkable. Although this observation is 52 298 based only on one specimen and more observations are necessary, this fact stresses this 53 299 character as highly diagnostic for this species, possibly throughout its life as a rhynchoteuthion. 54 55 300 The rhynchoteuthion paralarvae of the congeneric species T. pacificus differ from T. sagitattus 56 301 in their smaller size (up to 1.4 mm ML, Puneeta et al., 2015: Fig. 9), the chromatophore pattern 57 58 59 60 7

Zoological Journal of the Linnean Society Page 8 of 31

1 2 3 302 (see Discussion) and the proboscis suckers, which are of equal size in T. pacificus (Puneeta et 4 303 al., 2015: Fig. 8). Therefore, this constitutes a Type C rhynchoteuthion. 5 6 304 Todaropsis eblanae hatchlings 7 8 305 This species is the largest rhynchoteuthion hatchling (Table 1) described to date and exhibits 9 306 more advanced development at the time of hatching than other previously described species 10 307 (Fig. 1il, 2I, 3gI). In addition to the arm pairs I, II and IV, the pair III arm primordia are 11 308 present as well (Fig. 2i). While arm pairs I and II possess one sucker, pairs III and IV have no 12 13 309 suckers. The mean VML is 2.16 ± 0.11 mm. On the head, there are 2 chromatophores ventrally, 14 310 one below each eye; dorsally there are three rows of chromatophores, the first formed by a 15 311 single chromatophore anterior to the eyes, the second formed by two chromatophores at the 16 312 level of the eyes and the third formed by 3 chromatophores at the base of the head. On the 17 313 mantle, there are up to 6 rows ventrally, distributed as follows: 7 + 5 + 0 + 0 + 5 + 2; dorsally, 18 there are up to 5For rows: 3 + 0 Review+ 5 + 0 + 2. Between 1 andOnly 4 true lateral chromatophores are 19 314 20 315 present on the mantle (Fig. 3i, Table 2). The 8 proboscis suckers are of similar size (54.3 ± 5.0 21 316 µm) (Type C rhynchoteuthion) and have two rows of pegs, the internal one with 18.1 ± 1.0 and 22 317 the external one with 22.0 ± 2.2 pegs (Fig. 3k), with up to 3 additional pegs asymmetrically 23 318 distributed outside of the external row. The arm suckers measure 63.0 ± 5.5 µm and bear two 24 319 rows of pegs (Fig. 2b), the internal one formed by 17.0 ± 0.8 and the external by 20.6 ± 1.5 25 26 320 pegs. Between 0 and 8 pegs arranged externally to the external row can be found 27 321 asymmetrically scattered. The skin of the species shows a hexagonlike structure under the 28 322 stereomicroscope (Fig. 2l). DMLI is smaller than in the other two species, which indicates a 29 323 similar length between DML and VML. Although there is some overlap with the other two 30 324 species, HWI is smaller, and thus the head is narrower. The AIILI shows a larger pair of arms II 31 325 than I. coindetii and T. sagittatus . 32 33 326 Remarks: Roura (2013) sequenced two wild rhynchoteuthion paralarvae which did not produce 34 35 327 a specieslevel match with any previously sequenced ommastrephid. He hypothesized that those 36 328 collected in the oceanic realm should be assigned to Todarodes sagitattus and those from the 37 329 shelf should be assigned to Todaropsis eblanae . The recent work of Gebhardt & Knebelsberger 38 330 (2015) provided available barcodes for these two species, which could be used to positively 39 331 identify wild rhynchoteuthions by DNA barcoding (Hebert et al., 2003). A BLAST (Altschul et 40 332 al., 1990) search show s that the sequence with the GenBank accession number LN614712, 41 42 333 uploaded and identified as T. eblanae by Roura, does represent an individual of this species. 43 44 334 Ommastrephes bartramii paralarvae (Fig. 3jk) 45 46 335 The following description is based on Sweeney et al. (1992), Young & Hirota (1990) Sakurai et 47 336 al. (1995) and Vijai et al. (2015). The hatchling bears the pairs of arms I, II and IV, the latter 48 337 devoid of suckers. On the head, there is a row of two ventral chromatophores; dorsally, there are 49 338 two rows formed by 1 and 2 chromatophores, respectively. On the mantle, the ventral surface 50 339 has up to 3 rows with the following formula: (34) + (01) + 1; dorsally there are up to 2 rows: 51 340 (45) + (01). Later (~3 mm DML), two ventrally centred chromatophores appear on the edge of 52 53 341 the mantle. The dorsal pattern becomes scatteredlike at size ~4 mm mantle length and ventrally 54 342 at ~6 mm mantle length. The ratio between the lateral and medial proboscis suckers is 2:1 (Type 55 343 A rhynchoteuthion). There are two rows of pegs on the proboscis suckers. There are more pegs 56 344 on the lateral suckers (~20 internal, ~27 external) than on the medial ones (~11 internal, ~14 57 58 59 60 8

Page 9 of 31 Zoological Journal of the Linnean Society

1 2 3 345 external). The skin shows a hexagonlike structure under the stereomicroscope (Vijai et al., 4 346 2015: Fig 7jm). 5 6 347 paralarvae (Fig. 3n) 7 8 348 The following descriptions are based on Sweeney et al. (1992). This species is characterized by 9 349 the presence of two equallysized intestinal photophores and another single on the 10 350 ventral surface of each eye. The chromatophore pattern has not been described for this species. 11 351 There are no size differences among the proboscis suckers (Type C rhynchoteuthion). The pegs 12 13 352 of the proboscis suckers are unknown. 14 15 353 Hyaloteuthis pelagica paralarvae (Fig. 3lm) 16 17 354 The following descriptions are based on Harman & Young (1985), Sweeney et al. (1992) and 18 355 Diekmann et al. (2002).For This paralarvaReview bears one central Only intestinal photophore and another on 19 356 the ventral surface of each eye. According to Diekmann et al. (2002: Table 5), the intestinal 20 357 photophore is visible in individuals of 1.5 mm ML or larger. The chromatophore pattern is 21 358 formed by highly scattered units on both the mantle and head. On the head, one chromatophore 22 359 is located on the ventral surface of each eye and another one is located on the dorsal surface of 23 24 360 the eye. On the head, dorsally there are four rows of chromatophores (excluding the ocular 25 361 chromatophores): 1 + 2 + 1. Individuals smaller than 2 mm of DML lack chromatophores on the 26 362 ventral surface of the mantle; later a single row of chromatophores on the anterior edge and four 27 363 chromatophores forming a diamondshape pattern between the fins appear. Dorsally on the 28 364 mantle, a single large chromatophore is present on the first third of the mantle surface during 29 365 most of its life as a rhynchoteuthion. The ratio of the proboscis suckers is between 1.25:1 and 30 31 366 1.5:1 (Type B rhynchoteuthion). There are two rows of pegs on the proboscis suckers. There are 32 367 more pegs on the lateral suckers (1416 internal, 1819 external) than on the medial ones (811 33 368 internal, 815 external). 34 35 369 Ornithoteuthis antillarum paralarvae (Fig. 3o) 36 37 370 The following description is based on Sweeney et al. (1992) and Diekmann et al. (2002). This 38 371 paralarva has single and round ventral photophores on the ventral surface of each eye and two 39 372 unequallysized intestinal photophores: the anterior one appears first (Sweeney et al., 1992), but 40 373 the posterior one grows larger (Diekmann et al., 2002: Table 5). The chromatophore pattern is 41 42 374 not known. The ratio of the proboscis suckers is up to 1.5:1 (Type B rhynchoteuthion). The pegs 43 375 of the proboscis suckers are not known. 44 45 376 Key for the NE Atlantic rhynchoteuthion paralarvae 46 47 377 The primary characters of this key are the size of the proboscis suckers, the presence/absence of 48 378 photophores and the arrangement of the proboscis sucker pegs. Although the first two characters 49 379 are observable under a standard stereomicroscope, the proboscis pegs should be observed by 50 380 specialized microscopy methods, such as SEM. Since the number and arrangement of the 51 381 chromatophores are variable during the ontogeny of the rhynchoteuthions, we only use these 52 53 382 characters when they are taxonomically significant (key step 3a). 54 55 383 1a. Ratio of sucker sizes of 2:1 (Type A rhynchoteuthion), two rows of pegs in proboscis 56 384 suckers. More pegs in the lateral proboscis suckers (~20 internal, ~27 external) than in the 57 385 medial ones (~11 internal, ~14 external). Photophores absent. 58 59 60 9

Zoological Journal of the Linnean Society Page 10 of 31

1 2 3 386 Ommastrephes bartramii 4 5 387 1b . Ratio of sucker sizes lower than 2:1 2 6 7 388 2a. Ratio of sucker sizes between 1.1:1 and 1.9:1 (Type B rhynchoteuthion) 3 8 9 389 2b . Ratio of sucker sizes of 1:1 (Type C rhynchoteuthion) 5 10 11 390 3a. Absence of intestinal or ocular photophores, two rows of pegs in proboscis suckers (1224 12 391 internal, 1222 external), no differences in the number of pegs between lateral and medial 13 392 proboscis suckers. No ventral head chromatophores. 14 15 393 Todarodes sagittatus 16 17 394 3b . With ocular and intestinal photophores. 4 18 For Review Only 19 395 4a . Presence of one central intestinal photophore. More pegs in the lateral proboscis suckers 20 396 (1416 internal, 1819 external) than in the medial ones (811 internal, 815 external). 21 22 397 Hyaloteuthis pelagica 23 24 398 4b . Presence of two unequal intestinal photophores, the posterior one larger than the first one, 25 399 proboscis sucker pegs unknown. 26 27 400 Ornithoteuthis antillarum 28 29 401 5a . Presence of two equallysized intestinal photophores and another single photophore on the 30 402 ventral surface of each eye, proboscis sucker pegs unknown. 31 32 403 Sthenoteuthis pteropus 33 34 404 5b . Without intestinal or ocular photophores 6 35 36 405 6a . Only one row of pegs on the proboscis suckers (1014, 112 free ones), skin without a 37 38 406 hexagonlike pattern. 39 40 407 Illex coindetii 41 408 6b . Two rows of pegs (1624 internal, 1727 external plus 03 free ones) on the proboscis 42 rd 43 409 suckers, presence of the 3 pair of arms at the time of hatching, skin with a hexagonlike 44 410 pattern. 45 46 411 Todaropsis eblanae 47 48 412 Test of the dichotomous key using wild rhynchoteuthion paralarvae 49 50 413 Fourteen of the 16 rhynchoteuthions examined were successfully identified with the aid of the 51 414 key: 9 belonged to I. coindetii , 4 to T. sagittatus and 1 to S. pteropus (Table 4, Fig. 4 ). Among 52 415 those not identified to the species level with the key there is a Type C paralarva with 2 rows of 53 416 dorsal head chromatophores (Fig. 4b) , compatible with the pattern of I. coindetii and not with T. 54 417 eblanae . Thus, this emphasizes the utility of the chromatophore pattern for the identification of 55 56 418 hatchlings when other characters are not available. The second unidentified rhynchoteuthion is a 57 419 Type C paralarva without photophores. It was not possible to determine whether it was I. 58 420 coindetii or T. eblanae . Fifteen of the specimens were obtained in the Mediterranean Sea: this is 59 60 10

Page 11 of 31 Zoological Journal of the Linnean Society

1 2 3 421 the first time that identification of Mediterranean ommastrephid paralarvae from formalin 4 422 preserved plankton samples is possible with certainty. In general, the most useful characters for 5 423 the identification of the paralarvae are: 1) the size of the proboscis suckers, 2) the 6 424 presence/absence of ocular or intestinal photophores and 3) the proboscis pegs. 7 8 425 Discussion 9 10 426 Previous studies have stressed the difficulties of identifying ommastrephid paralarvae from 11 427 plankton samples (Collins et al., 2002; Gilly et al., 2006; Moreno, 2008; Moreno et al., 2009; 12 13 428 Roura, 2013; Zaragoza et al., 2015). Morphological clusters could sometimes be identified 14 429 based only on the paralarvae morphology, but it was not possible to confirm the species (e. g. 15 430 Roper & Lu, 1979; Vecchione et al., 2001 ; Moreno, 2008). Other times, species have been 16 431 identified based on the adult characters that could be seen in the paralarvae, such as photophores 17 432 (e. g. Sweeney et al., 1992). Three different sources of paralarval species confirmation could be 18 used, namely: a) aquariumFor spawning, Review b) in vitro fertilization Only and c) molecular methods. The first 19 433 20 434 method is limited by the difficulties in catching, housing and maintaining broodstock and 21 435 inducing spawning (Durward et al., 1980; Bower & Sakurai, 1996). The last method is 22 436 constrained by the need for previous molecular data obtained from properly identified adults, 23 437 which may not exist for some species (Roura, 2013). Thus, the main advantage of using in vitro 24 438 fertilization is that the species identification of the paralarvae produced is ensured. Here, this 25 26 439 method was successfully used to describe the paralarvae morphology of three ommastrephid 27 440 species. 28 29 441 Towards the reliable morphological identification of rhynchoteuthion paralarvae: which are the 30 442 most useful characters? 31 32 443 The morphological description of I. coindetii , T. sagittatus and T. eblanae revealed new 33 444 taxonomic characters that permit the identification of rhynchoteuthion paralarvae collected in 34 445 plankton samples from the NE Atlantic. In the case of I. coindetii , when alive, they can be easily 35 446 identified by the absence of any special skin sculpture (Fig. 2d), a character shared with the 36 37 447 congeneric I. argentinus (Sakai et al., 1998: Fig. 4). Skin sculpture is present in fresh T. 38 448 sagittatus and T. eblanae (Fig. 2h, l, respectively), T. pacificus (Watanabe et al., 1996: Fig. 7; 39 449 Puneeta et al., 2015: Fig. 8), O. bartramii (Vijai et al., 2015: Fig 7jm) and Dosidicus gigas 40 450 (d’Orbigny, 1835: 50 [in 1834–1847]) (D. Staaf, pers. comm.). This skin sculpture is an optical 41 451 effect created by light that crosses vertical expansions of the basal membrane of the outer 42 452 epithelium between the mucous cells (F. Á. FernándezÁlvarez, unpubl. observation) . Thus, the 43 44 453 absence of this skin sculpture in hatchlings seems to be a synapomorphy of the genus Illex . 45 454 When dealing with fixed specimens, the presence of only one row of pegs in the proboscis 46 455 suckers allows I. coindetti to be differentiated from other sympatric rhynchoteuthions described 47 456 to date. 48 49 457 Todarodes sagittatus has a type B paralarva, a condition that differentiates this species from 50 458 sympatric ommastrephids, except for H. pelagica and O. antillarum , which bear ocular and 51 459 intestinal photophores, with the former bearing a peculiar chromatophore pattern. When we 52 53 460 compare T. sagittatus hatchlings with the congeneric species T. pacificus (Watanabe et al., 54 461 1996), the main difference is the absence of ventral head chromatophores in T. sagittatus . Based 55 462 on the chromatophore pattern depicted by Watanabe et al. (1996: Fig. 8), T. pacificus bears 5 56 463 rows on the ventral surface of the mantle, whereas T. sagittatus hatchlings have 34 (Table 2). 57 464 The specimens of T. pacificus depicted by Puneeta et al. (2015) are Type C rhynchoteuthions, 58 59 60 11

Zoological Journal of the Linnean Society Page 12 of 31

1 2 3 465 however more accurate information on the structure of the proboscis suckers, such as the 4 466 arrangement of pegs of the suckers, is lacking for comparison against T. sagittatus . 5 6 467 Todaropsis eblanae is the largest rhynchoteuthion hatchling described so far and the presence of 7 468 the third arm buds at hatching is a diagnostic character that differs from other described 8 469 rhynchoteuthion hatchlings. It can be easily differentiated from S. pteropus (also Type C 9 rhynchoteuthion) because it has no photophores. It can also be differentiated from I. coindetii 10 470 11 471 because T. eblanae has two rows of pegs in the proboscis suckers and a hexagonlike skin 12 472 sculpture. The morphological descriptions provided here for I. coindetii , T. sagittatus and T. 13 473 eblanae paralarvae were based on hatchlings and some characters are known to change during 14 474 their development (see below), such as morphometrics (Table 1) and the chromatophore pattern 15 475 (Table 2). However, the structure of the eight proboscis suckers (Table 3) is unlikely to change 16 17 476 during the rhynchoteuthion phase (see below). 18 Ommastrephes bartramiiFor is unlikely Review to be confused with Onlyother rhynchoteuthions present in the 19 477 20 478 NE Atlantic, as it is the only Type A rhynchoteuthion in these waters. However, in other areas 21 479 of its known distribution, this species overlaps with other species with this type of paralarva, 22 480 such as one of the species of the genus Ornithoteuthis Okada, 1927 (Wakabayashi et al., 2002): 23 481 (Sasaki, 1915). This species is sympatric with O. bartramii in both the 24 482 Pacific and Indian Oceans, but its rhynchoteuthion has fewer pegs on the lateral proboscis 25 26 483 suckers (Wakabayashi et al., 2002: Table 6). Moreover, Ornithoteuthis rhynchoteuthions have 27 484 two intestinal photophores (Sweeney et al., 1992; Diekmann et al., 2002: Table 5). It is not clear 28 485 if Nototodarus Pfeffer, 1912 species are Type B or A rhynchoteuthions or both (Sweeney et al., 29 486 1992). Since these species also lack photophores, they are the only species that may be 30 487 misidentified as O. bartramii in areas where their distribution overlaps. More research is needed 31 488 to clarify the morphology of Nototodarus paralarvae. 32 33 489 The presence of two intestinal photophores in S. pteropus can easily differentiate this species 34 35 490 from H. pelagica (which has a single round intestinal photophore) in the NE Atlantic, in 36 491 addition to the proboscis suckers (Type C in the first species, Type B in the second one). 37 492 Ornithoteuthis antillarum also has two intestinal photophores, however the presence of large 38 493 lateral suckers and the unequal size of the intestinal photophores in O. antillarum distinguishes 39 494 the two species. 40 41 495 The presence of a single intestinal photophore differentiates H. pelagica from the other 42 496 rhynchoteuthions from the NE Atlantic. The proboscis sucker pegs of H. pelagica were studied 43 44 497 by Harman & Young (1985), showing a pattern with ~14 internal, ~17 external pegs in the 45 498 lateral suckers and ~9 internal, ~10 external in the medial ones. However, another type B 46 499 species with a single intestinal photophore, Eucleoteuthis luminosa (Sasaki, 1915), is sympatric 47 500 throughout the distribution area of H. pelagica with the exception of the entire North Atlantic . 48 501 For both species, the proboscis sucker pegs have been described (Harman & Young, 1985 for H. 49 502 pelagica and Wakabayashi et al., 2002 and GranadosAmores et al., 2013 for E. luminosa ), and 50 51 503 the number of pegs in the proboscis suckers seems to be higher than in H. pelagica , although 52 504 some overlap exists (Wakabayashi et al., 2002: Table 4). Wakabayashi et al. (2006) and 53 505 GranadosAmores et al. (2013) molecularly identified paralarvae of E. luminosa , validating its 54 506 description . 55 56 507 Ornithoteuthis antillarum can be easily differentiated from other Type B rhynchoteuthions by 57 508 its two unequallysized photophores. This character also differentiates this species from E. 58 59 60 12

Page 13 of 31 Zoological Journal of the Linnean Society

1 2 3 509 luminosa in the S Atlantic. The congeneric O. volatilis is a Type A rhynchoteuthion 4 510 (Wakabayashi et al., 2002), while O. antillarum is a Type B rhynchoteuthion (Diekmann et al, 5 511 2002). 6 7 512 Based on the information presented above, it is possible to evaluate the reliability of 8 513 identifications of rhynchoteuthions based on each taxonomic character. Although the 9 chromatophore pattern led to the correct identification of a T. sagittatus specimen from the 10 514 11 515 literature and an I. coindetii specimen from our wild collected paralarvae (see above and Table 12 516 4, Fig. 4bc), it should be noted that this pattern changes during the ontogeny (e.g. Young & 13 517 Hirota, 1990). The same occurs with morphometrics: the shape of the body of rhynchoteuthion 14 518 paralarvae changes with the size of the animal (e.g. Young & Hirota, 1990, Vidal, 1994 or Gilly 15 519 et al., 2006) and the different morphometric indices change during ontogeny (RamosCastillejos 16 17 520 et al., 2010). Although differences between the number of pegs on arm suckers do exist between 18 521 species (Table 3),For comparisons Revieware only possible between Onlyparalarvae of similar sizes (in the case 19 522 of the three rhynchoteuthion species described here, hatchlings and the immediately following 20 523 stages). Moreover, while the paralarva grows, more suckers are added to the arms, making it 21 524 difficult to identify each single sucker. As can be seen in RamosCastillejos et al. (2010) and 22 525 Wakabayashi et al. (2002), the number of pegs per row on arm suckers increases while the 23 24 526 paralarva grows. Thus, we do not recommend relying only on chromatophore pattern, 25 527 morphometrics and arm sucker pegs, without the support of others characters for identification 26 528 purposes. A combination of both proboscis suckers and photophores seems to be the most 27 529 reliable combination. Photophores are easy to find under the stereomicroscope, both in fresh and 28 530 fixed specimens. However, no member of the subfamilies Illicinae Posselt, 1891 or 29 531 Adam, 1960 is known to possess photophores. Together these subfamilies represent 64% of the 30 31 532 species biodiversity among the Ommastrephidae. Moreover, this character should be considered 32 533 with caution since it is not known when photophores appear in some species (Sweeney et al., 33 534 1992; Diekmann et al., 2002) and at least in D. gigas some variation for photophore appearance 34 535 during ontogenetic development is suspected (Gilly et al., 2006 found paralarvae of this species 35 536 with the typical photophores present in the juvenile). Again, by using a binocular microscope it 36 537 is easy to differentiate between rhynchoteuthion Types A, B and C. When species of the same 37 38 538 Type coexist (I. coindetii and T. eblanae in NE Atlantic), the best approach is to study the 39 539 proboscis pegs by SEM (Table 4, Fig. 4a ). While an individual is a rhynchoteuthion, they have 40 540 the same 8 proboscis suckers until the proboscis splits and the animal becomes a juvenile (Shea, 41 541 2005; Wakabayashi et al., 2002). As RamosCastillejos et al. (2010: Fig. 9) have shown for D. 42 542 gigas , the number of internal and external pegs in the 8 proboscis suckers remains the same 43 543 throughout the rhynchoteuthion stage. 44 45 544 The search for the hatchling stage in ommastrephids 46 47 545 An unresolved discussion remains regarding which of the embryological stages described for 48 49 546 ommastrephid squids represents the actual hatchling. This question was raised by Watanabe et 50 547 al. (1996), who observed a twostage delay between hatching from aquariaspawning eggs and 51 548 those from in vitro fertilization in T. pacificus . They hypothesize that the hatching stage should 52 549 be addressed between stages XXV and XXVII, because Hoyle´s organ was visible during these 53 550 stages. A similar observation was reported by Staaf et al. (2008) for D. gigas . Recently, Puneeta 54 551 et al. (2015) observed hatchlings of stage XXXI from aquariumspawned T. pacificus . In short, 55 56 552 although several experiments with eggs obtained by in vitro fertilization and by aquaria 57 553 spawning in ommastrephids have been performed, the question is unresolved. 58 59 60 13

Zoological Journal of the Linnean Society Page 14 of 31

1 2 3 554 In naturallyhatched individuals the development of the characteristics that allow them to swim 4 555 and avoid predators is expected. However, individuals at stages XXIX or younger never have a 5 556 fully developed ink sac, fins or statoliths. Thus, correct swimming is not possible at this time 6 557 and these individuals can only perform horizontal movements on the bottom of the Petri dish, 7 probably by using cilia on the skin (R. Villanueva, unpubl. observation). Additionally, in the 8 558 9 559 case of T. pacificus (Watanabe et al., 1996), T. sagittatus and T. eblanae , fins are not fully 10 560 developed until stage XXXII. In the present study, the criteria to select a stage as hatchling was 11 561 a paralarva capable of swimming, with a fully developed ink sac, statoliths and fins as wide as 12 562 the head. This definition corresponds to stages beyond XXX ( sensu Sakai et al., 1998) for Illex 13 563 and XXXII ( sensu Watanabe et al., 1996) for Todarodes and Todaropsis . Taking into account 14 15 564 these criteria, any stage below XXXII should not be addressed as the hatchling for O. bartramii , 16 565 based on the description by Vijai et al. (2015). In addition to these arguments, individuals less 17 566 developed than those considered here are never reported from plankton samples (for instance, a 18 567 paralarva withoutFor fins). The presenceReview of these prematurely Only hatched individuals, in our opinion, 19 568 does not indicate the natural hatchling stage and is a probable consequence of suboptimal 20 569 artificial conditions in the laboratory (Villanueva et al., 2011). Better culture conditions would 21 22 570 likely produce hatchlings that are morphologically more similar to those produced in the wild as 23 571 the results reported by Puneeta et al. (2015) suggest. 24 25 572 Conclusions 26 27 573 Comprehensive knowledge on the life cycle of ommastrephid squids is hindered by the fact that 28 574 many aspects of the first stages of their life cycle remain a mystery, especially regarding the 29 575 paralarvae. For instance, we still do not know the main diet of early rhynchoteuthions as their 30 576 stomachs are usually empty or contain unrecognized food (e.g. Uchikawa et al., 2009; 31 577 CamarilloCoop et al., 2013). Vidal & Haimovici, (1998) have found microorganisms (ciliates, 32 33 578 flagellates and bacteria) with mucus inside the digestive tract of early rhynchoteuthions and 34 579 copepod appendages on the digestive tracts of paralarvae > 3.7 mm ML ; together these 35 580 observations indicate that ommastrephid paralarvae certainly adopt different feeding strategies 36 581 as they grow (Uchikawa et al., 2009; Shea, 2005) . This lack of knowledge inevitably leads to a 37 582 total failure when any culture experiment is attempted (e. g. Yatsu, Tafur & Maravi, 1999; Staaf 38 583 et al., 2008). Thus, the little knowledge available on paralarvae beyond yolk consumption 39 40 584 proceeds from paralarvae collected in the wild, including the morphology of posthatchlings (e. 41 585 g. Young & Hirota, 1990), the bathymetric layers suitable for their development (e. g. Roper & 42 586 Lu, 1979), their distribution (e.g. Staaf et al., 2013) and the tentative length of the paralarval 43 587 period (e. g. Uchikawa et al., 2009). Clearly, if our knowledge in all these matters relies on wild 44 588 collected paralarvae, a wellestablished taxonomic knowledge of each species for reliable 45 589 identification is necessary. Here, we fill a gap in the knowledge of NE Atlantic 46 47 590 rhynchoteuthions by describing the morphology of three of the seven species and used this new 48 591 knowledge in conjunction with the literature to develop the first dichotomous key covering all 49 592 the ommastrephid squid paralarvae present in a wide area. Moreover, this methodology was 50 593 applied to identify 16 plankton sampled ommastrephid paralarvae and, for the first time, data on 51 594 properly identified paralarvae was provided for the Mediterranean Sea. This key will permit the 52 595 identification of rhynchoteuthions from plankton samples collected in the NE Atlantic, thus 53 54 596 providing a useful tool for future studies on the planktonic life and population dynamics of 55 597 ommastrephid squids. 56 57 598 Acknowledgements 58 59 60 14

Page 15 of 31 Zoological Journal of the Linnean Society

1 2 3 599 We greatly appreciate the help with squid collection of the crew of the following trawlers: Pilar 4 600 Rus , La Seu , Avi Pau , Pilar Peña , Montserrat , Silvia i Ben (Vilanova i la Geltrú) and Maireta 5 601 IV (Barcelona); and Elisabeth CuestaTorralvo for her support during collecting samples. 6 602 Thanks to José Manuel Fortuño for his dedication and kindness during the SEM sessions. 7 Thanks are due to Ana Sabatés and Vanesa Raya, for providing the wild rhynchoteuthion 8 603 9 604 paralarvae collected during the research project CACO (REN200201339/MAR) and Laura 10 605 Recasens and Pilar Olivar for those collected on the LLUÇ3 cruise. María Pascual kindly 11 606 provided us with the wild paralarva from the project FishJelly. The Sthenoteuthis pteropus 12 607 specimen was collected on board the RV Hesperides during the research project MAFIA 13 608 (CTM201239587C0403), funded by the Spanish Ministry of Economy and Competitiveness 14 15 609 (MINECO). Thanks to Dr. Louise Allcock and three anonymous reviewers for their comments . 16 610 FAFA was supported by a predoctoral fellowship of the MINECO (BES2013063551). EAGV 17 611 was supported by the Brazilian National Research CouncilCNPq (Grants no. 207680/20140, 18 612 250017/20130). For This study wasReview funded by the research Only projects CALOCEAN1 (AGL2009 19 613 11546) and CALOCEAN2 (AGL201239077) from the MINECO. 20 21 614 References 22 23 615 Adam W. 1957. Notes sur les Céphalopodes. XXIII – Quelques espèces des Antilles. Bulletin de 24 616 l’Institut Royal des Sciences Naturelles de Belgique 33: 110. 25 26 617 Adam W. 1960. Notes sur les Cephalopodes XXIV: Contribution a la connaissance de 27 l’hectocotyle chez les Ommastrephidae. Bulletin de l’Institut Royal des Sciences Naturelles de 28 618 29 619 Belgique 36: 110. 30 31 620 Adam W. 1962. Céphalopodes de l’Archipel du CapVert, de l’Angola et du Mozambique. 32 621 Memorias da Junta de Investigacoes do Ultramar (2ª Série) 33: 964. 33 34 622 Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. 1990. Basic local alignment search 35 623 tool. Journal of Molecular Biology 215: 403410. 36 37 624 Arkhipkin AI, Rodhouse PGK, Pierce GJ, Sauer W, Sakai M, Allcock L, Arguelles J, Bower JR, 38 625 Castillo G, Ceriola L, Chen CS, Chen X, DiazSantana M, Downey N, González AF, Granados 39 626 Amores J, Green CP, Guerra A, Hendrickson LC, Ibáñez C, Ito K, Jereb P, Kato Y, Katugin 40 627 ON, Kawano M, Kidokoro H, Kulik VV, Laptikhovsky VV, Lipinski MR, Liu B, Mariátegui L, 41 42 628 Marin W, Medina A, Miki K, Miyahara K, Moltschaniwskyj N, Moustahfid H, Nabhitabhata J, 43 629 Nanjo N, Nigmatullin CM, Ohtani T, Pecl G, Perez JAA, Piatkowski U, Saikliang P, Salinas 44 630 Zavala CA, Steer M, Tian Y, Ueta Y, Vijai D, Wakabayashi T, Yamaguchi T, Yamashiro C, 45 631 Yamashita N, Zeidberg LD. 2015. World Squid Fisheries. Reviews in Fisheries Science & 46 632 Aquaculture 23: 92252. 47 48 633 Ball R. 1841. On a species of Loligo , found on the shore of Dublin Bay. Proceedings of the 49 634 Royal Irish Academy 1: 362364. 50 51 635 Berry SS. 1912. A catalogue of Japanese Cephalopoda. Proceedings of the Academy of Natural 52 53 636 Sciences of Philadelphia 64: 380444. 54 55 637 Boletzky SV, Rowe L, Aroles L. 1973. Spawning and development of the eggs, in the 56 638 laboratory, of Illex coindetii (Mollusca: Cephalopoda). Veliger 15: 257258. 57 58 59 60 15

Zoological Journal of the Linnean Society Page 16 of 31

1 2 3 639 Bosc LAG. 1802. Histoire naturelle des vers contenant leur descriptionet leurs moeurs; avec 4 640 figures dessinees d’apres nature . Paris: Déterville. 5 6 641 Bower JR, Sakurai Y. 1996. Laboratory observations on Todarodes pacificus (Cephalopoda: 7 642 Ommastrephidae) egg masses. American Malacological Bulletin 13: 6571. 8 9 643 Boyle PR, Rodhouse PG. 2005. Cephalopods. Ecology and Fisheries . Oxford: Blackwell 10 644 Science Ltd. 11 12 645 Brasseur P, Beckers JM, Brankart JM, Schoenauen R. 1996. Seasonal temperature and salinity 13 646 fields in the Mediterranean Sea: Climatological analyses of a historical data set. Deep-Sea 14 15 647 Research I 43: 159192. 16 17 648 Brunetti NE. 1990. An scale for identification of stages of sexual maturity in the Argentine 18 649 squid ( Illex argentinusFor). Frente Review Maritimo 7: 4551. Only 19 20 650 CamarilloCoop S, SalinasZavala CA, Lavaniegos BE, Markaida U. 2013. Food in early life 21 651 stages of Dosidicus gigas (Cephalopoda: Ommastrephidae) from the Gulf of California, 22 652 Mexico. Journal of the Marine Biological Association of the United Kingdom 93: 19031910. 23 24 653 Carlini DB, Kunkle LK, Vecchione M. 2006. A molecular systematic evaluation of the squid 25 654 genus Illex (Cephalopoda: Ommastrephidae) in the North Atlantic Ocean and Mediterranean 26 655 Sea. Molecular Phylogenetics and Evolution 41: 496502. 27 28 656 Castellanos ZJA. 1960. Una nueva especie de calamar Argentino, Ommastrephes argentinus sp. 29 30 657 nov. (Mollusca, Cephalopoda). Neotropica 6: 5558. 31 32 658 Coll M, Navarro J, Olson RJ, Christensen V. 2013. Assessing the trophic position and 33 659 ecological role of squids in marine ecosystems by means of foodweb models. Deep-Sea 34 660 Research II 95: 2136. 35 36 661 Collins MA, Yau C, Boyle PR, Friese D, Piatkowski U. 2002. Distribution of cephalopods from 37 662 plankton surveys around the British Isles. Bulletin of Marine Science 71: 239254. 38 39 663 Diekmann R, Piatkowsky U, Schneider M. 2002. Early life and juvenile cephalopods around 40 664 seamounts of the subtropical eastern North Atlantic: Illustrations and a key for their 41 665 identification. Berichte aus dem Institut für Meereskunde an der Christian-Aabrechts- 42 43 666 Universität Kiel 326: 142. 44 D’Orbigny A. 1835 (18341847). Mollusques. Voyage dans l’Amerique Meridionale 5: 1758. 45 667 46 Durward RD, Vessey E, O'Dor RK, Amaratunga T. 1980. Reproduction in the squid , Illex 47 668 48 669 illecebrosus : first observation in captivity and implications for the life cycle. International 49 670 Commision for the Nortwest Atlantic Fisheries Selected Papers 6: 713. 50 51 671 Gebhardt K, Knebelsberger T. 2015. Identification of cephalopod species from the North and 52 672 Baltic Seas using morphology, COI and 18S rDNA sequences. Helgoland Marine Research 69: 53 673 259271. 54 55 674 Gilly WF, Elliger CA, Salinas CA, CamarillaCoop S, Bazzino G, Beman M. 2006. Spawning 56 675 by jumbo squid Dosidicus gigas in San Pedro Mártir Basin, Gulf of California, Mexico. Marine 57 676 Ecology Progress Series 313: 125133. 58 59 60 16

Page 17 of 31 Zoological Journal of the Linnean Society

1 2 3 677 GranadosAmores J, Ramos JE, SalinasZavala CA, CamarilloCoop S. 2013. Eucleoteuthis 4 678 luminosa (cephalopoda: ommastrephidae) off the west coast of the Baja California Peninsula, 5 679 Mexico. Journal of Shellfish Research 32: 341346. 6 7 680 Harman RF, Young RE. 1985. The larvae of ommastrephid squids (Cephalopoda, Teuthoidea) 8 681 from Hawaiian waters. Vie Milieu 35: 211222. 9 10 682 Hebert PD, Cywinska A, Ball SL, deWaard JR. 2003. Biological identifications through DNA 11 683 barcodes. Proceedings of the Royal Society B: Biological Sciences 270: 313321. 12 13 684 Jereb P, Roper CFE. 2010. Cephalopods of the world. An annotated and illustrated catalogue of 14 15 685 cephalopod species known to date. Vol 2. Myopsid and Oegopsid Squids . Rome: FAO. 16 17 686 Lamarck JB. 1798. Extrait d’un mémoire sur le genre de la Séche, du Calmar et Poulpe, 18 687 vulgairement nommés,For Polypes Review de Mer. Bulletin des Sciences, Only par la Société Philomatique de 19 688 Paris 2: 129131. 20 21 689 Lesson RP. 18301831. Voyage autour du monde: exécuté par ordre du Roi, sur la corvette de 22 690 Sa Majesté, la Coquille, pendant les années 1822, 1823, 1824, et 1825 . Paris: A. Bertrand. 23 24 691 Lesueur CA. 1821. Descriptions of several new species of cuttlefish. Journal of the Academy of 25 692 Natural Sciences of Philadelphia 2: 86101. 26 27 693 Logan JM, Toppin R, Smith S, Galuardi B, Porter J, Lutcavage ME. 2013. Contribution of 28 694 cephalopod prey to the diet of large pelagic fish predators in the central North Atlantic Ocean. 29 30 695 Deep-Sea Research II 95: 7482. 31 32 696 Meillat M. 2012. Essais du chalut mésopélagos pour le programme MYCTO 3DMAP de l'IRD, 33 697 à bord du Marion Dufresne. Du 10 au 21 Août 2012. Rapport de mission Marion Dufresne. 34 698 Unpublished report, Ifremer. 35 36 699 Moreno A. 2008. Geographic variation in cephalopod life history traits. Unpublished 37 700 “Dissertação apresentada para accesso à categoria de Investigador Auxiliar” Thesis, Unidade de 38 701 Recursos Marinhos e Sustentabilidade LIPIMAR. 39 40 702 Moreno A, Santos A, Piatkowski U, Santos AMP, Cabral H. 2009. Distribution of cephalopod 41 703 paralarvae in relation to the regional oceanography of the western Iberia. Journal of Plankton 42 43 704 Research 31: 7391. 44 Nesis K. 1979. Squid larvae of the family Ommastrephidae (Cephalopoda). Zoologicheskii 45 705 46 706 Zhurnal 58: 1730. 47 48 707 O'Dor RK, Balch N, Foy EA, Hirtle RWM, Johnston DA, Amaratunga T. 1982. Embryonic 49 708 development of the squid, , and effect of temperature on development rates. 50 709 Journal of the Northwest Atlantic Fisheries Scien ce 3: 4145. 51 52 710 Okada YK. 1927. Contribution à l’etude des céphalopodes lumineaux (notes préliminaires). IV. 53 711 Ommastrephes volatilis Sasaki est une forme lumineaux; établissement d’un nouveau genre: 54 712 Ornithoteuthis . Bulletin de l’Institut Océanographique de Monaco 494: 1316. 55 56 713 Palomera I, Olivar MP, MoralesNin B. 2005. Larval development and growth of the European 57 714 hake in North Western Mediterranean. Scientia Marina 69: 251258. 58 59 60 17

Zoological Journal of the Linnean Society Page 18 of 31

1 2 3 715 Pfeffer G. 1912. Die Cephalopoden der PlanktonExpedition. Zugleich eine monographische 4 716 übersicht der Oegopsiden Cephalopoden. Ergebnisse der Plankton-Expedition der Humboldt- 5 717 stiftung 2: 1815. 6 7 718 Piatkowski U, Zumholz K, Lefkaditou E, Oesterwind D, Jereb P, Pierce GJ, Allcock L. 2015. 8 719 Todarodes sagittatus . . Cephalopod biology and fisheries in Europe: II. 9 Species Accounts. ICES Cooperative Research Report 235: 193205. 10 720 11 721 Posselt HI. 1891. Todarodes sagittatus (Lmk.) Stp., En anatomisk studie med Bemaerkinger om 12 13 722 Slaegtskabsforholdet mellem Ommatostrephfamiliens Genera. Videnskabelige Meddelelser fra 14 723 den Naturhistoriske Forening I Kjobenhavn 1890: 301359. 15 16 724 Puneeta P, Vijai D, Yoo HK, Matsui H, Sakurai Y. 2015. Observations on the spawning 17 725 behavior, egg masses and paralarval development of the ommastrephid squid Todarodes 18 726 pacificus in a laboratoryFor mesocosm. Review Journal of Experimental Only Biology 218: 38253835. 19 20 727 RamosCastillejos JE, SalinasZavala CA, CamarilloCoop S, EnríquezParedes LM. 2010. 21 728 Paralarvae of the jumbo squid, Dosidicus gigas . Invertebrate Biology 129: 172183. 22 23 729 Roper CFE, Lu CC. 1979. Rhynchoteuthion larvae of ommastrephid squids of the western 24 730 North Atlantic, with the first description of larvae and juveniles of Illex illecebrosus . 25 26 731 Proceedings of the Biological Society of Washington 91: 10391059. 27 28 732 Roper CFE, Lu CC, Mangold K. 1969. A new species of Illex from the Western Atlantic and 29 733 distributional aspects of other Illex species Cephalopoda: ). Proceedings of Biological 30 734 Society of Washington 82: 295322. 31 32 735 Roper CFE, Voss GL. 1983. Guidelines for taxonomic descriptions of cephalopod species. In: 33 736 Roper CFE, Lu CC, Hochberg FG, eds. Proceedings of the workshop on the biology and 34 737 resource potential of cephalopods. Memoirs of National Museum Victoria, Melbourne, Australia 35 738 44: 4963. 36 37 739 Roura A. 2013. Ecology of planktonic cephalopod paralarvae in coastal upwelling systems. 38 39 740 Unpublished D. Phil. Thesis, Vigo University. 40 41 741 Sabatés A, Salat J, Raya V, Emelianov M, SeguraNoguera M. 2009. Spawning environmental 42 742 conditions of Sardinella aurita at the northern limit of its distribution range, the western 43 743 Mediterranean. Marine Ecology Progress Series 385: 227236. 44 45 744 Sakai M, Brunetti NE, Elena B, Sakurai Y. 1998. Embryonic development and hatchlings of 46 745 derived from artificial fertilization. South African Journal of Marine Science 47 746 20: 255265. 48 49 747 Sakurai Y, Young RE, Hirota J, Mangold K, Vecchione M, Clarke MR, Bower J. 1995. 50 748 Artificial fertilization and development through hatching in the oceanic squids Ommastrephes 51 52 749 bartramii and Sthenoteuthis oualaniensis (Cephalopoda: Ommastrephidae). Veliger 38: 185 53 750 191. 54 55 751 Salman A, Katagan T, Benli HA. 2003. Vertical distribution and abundance of juvenile 56 752 cephalopods in Aegean Sea. Scientia Marina 67: 167176. 57 58 59 60 18

Page 19 of 31 Zoological Journal of the Linnean Society

1 2 3 753 Sasaki M. 1915. On three interesting new oegopsids from the Bay of Sagami. Journal of the 4 754 College of Agriculture, Tohoku Imperial University, Sapporo 6: 131150. 5 6 755 Shea EK. 2005. Ontogeny of the fused tentacles in three species of ommastrephid squids 7 756 (Cephalopoda, Ommastrephidae). Invertebrate Biology 124: 2538. 8 9 757 Staaf DJ, CamarilloCoop S, Haddock SHD, Nyack AC, Payne J, SalinasZavala CA, Seibel 10 758 BA, Trueblood L, Widmer C, Gilly WF. 2008. Natural egg mass deposition by the Humboldt 11 759 squid ( Dosidicus gigas ) in the Gulf of California and characteristics of hatchlings and 12 13 760 paralarvae. Journal of the Marine Biological Association of the United Kingdom 88: 759770. 14 15 761 Staaf DJ, Redfern JV, Gilly WF, Watson W, Balance LT. 2013. Distribution of ommastrephid 16 762 paralarvae in the eastern tropical Pacific. Fishery Bulletin 111: 7889. 17 18 763 Steenstrup J. 1855.For Kjaeber afReview en kolossal Blaeksprutte. Only Oversigt over det Kongelige Danske 19 764 Videnskabernes Selskabs Forhandlinger 1855: 199200. 20 21 765 Steenstrup J. 1880. De Ommatostrephagtige Blaeksprutter indbyrdes Forhold. Oversigt over det 22 766 Kongelige Danske Videnskabernes Selskabs Forhandlinger 1880: 73110. 23 24 767 Sweeney MJ, Roper CFE, Mangold KM, Clarke MR, Boletzky SV. 1992. "Larval" and juvenile 25 768 cephalopods: A manual for their identification . Washington DC: Smithsonian Institution Press. 26 27 769 Uchikawa K, Sakai M, Wakabayashi T, Ichii T. 2009. The relationship between paralarval 28 770 feeding and morphological changes in the proboscis and beaks of the 29 30 771 Ommastrephes bartramii . Fisheries Science 75: 317323. 31 32 772 Vecchione M, Roper CFE, Sweeney MJ, Lu CC. 2001. Distribution of paralarval cephalopods 33 773 in the western North Atlantic with observations on paralarval development. NOAA Technical 34 774 Report NMFS 152: 154. 35 36 775 Verany JB. 1839. Memoire sur six nouvelles espèces de Céphalopodes trouves dans la 37 776 Mediterranee a Nice. Memorie della Reale Accademia delle Scienze di Torino, (series 2) 1: 91 38 777 98. 39 40 778 Vidal EAG. 1994. Relative growth of paralarvae and juveniles of Illex argentinus (Castellanos, 41 779 1960) in southern Brazil. Antarctic Science 6: 275282. 42 43 780 Vidal EAG, Haimovici M. 1998. Feeding and the possible role of the proboscis and mucus 44 cover in the ingestion of microorganisms by rhynchoteuthion paralarvae (Cephalopod: 45 781 46 782 Ommastrephidae). Bulletin of Marine Science 63: 305316. 47 48 783 Vijai D, Sakai M, Sakurai Y. 2015. Embryonic and paralarval development following artificial 49 784 fertilization in the neon flying squid Ommastrephes bartramii . Zoomorphology 134: 417430. 50 51 785 Villanueva R, Quintana D, Petroni G, Bozzano A. 2011. Factors influencing the embryonic 52 786 development and hatchling size of the oceanic squid Illex coindetii following in vitro 53 787 fertilization. Journal of Experimental Marine Biology and Ecology 407: 5462. 54 55 788 Villanueva R, Staaf DJ, Argüelles J, Bozzano A, CamarilloCoop S, Nigmatullin CM, Petroni 56 789 G, Quintana D, Sakai M, Sakurai Y, SalinasZavala CA, De SilvaDávila R, Tafur R, 57 58 59 60 19

Zoological Journal of the Linnean Society Page 20 of 31

1 2 3 790 Yamashiro C, Vidal EAG. 2012. A laboratory guide to in vitro fertilization of oceanic squids. 4 791 Aquaculture 342–343: 125133. 5 6 792 Wakabayashi T, Saito K, Tsuchiya K, Segawa S. 2002. Descriptions of Eucloteuthis luminosa 7 793 (Sasaki, 1915) and Ornithoteuthis volatilis (Sasaki, 1915) paralarvae in the northwestern 8 794 Pacific. Venus 60: 237260. 9 10 795 Wakabayashi T, Suzuki N, Sakai M, Ichii T, Chow S. 2006. Identification of ommastrephid 11 796 squid paralarvae collected in northern Hawaiian waters and phylogenetic implications for the 12 13 797 family Ommastrephidae using mtDNA analysis. Fisheries Science 72: 494502. 14 15 798 Watanabe K, Sakurai Y, Segawa S, Okutani T. 1996. Development of the ommastrephid squid 16 799 Todarodes pacificus , from fertilized egg to rhynchoteuthion paralarva. American Malacological 17 800 Bulletin 13: 7388. 18 For Review Only 19 801 Yatsu A, Tafur R, Maravi C. 1999. Embryos and rhynchoteuthion paralarvae of the jumbo 20 802 flying squid Dosidicus gigas (Cephalopoda) obtained through artificial fertilization from 21 803 Peruvian waters. Fisheries Science 65: 904908. 22 23 804 Young RE, Hirota J. 1990. Description of Ommastrephes bartramii (Cephalopoda: 24 805 Ommastrephidae) paralarvae with evidence for spawing in Hawaiian waters. Pacific Science 44: 25 26 806 7180. 27 28 807 Zaragoza N, Quetglas A, Hidalgo M, ÁlvarezBerastegui D, Balbín R, Alemany F. 2015. 29 808 Effects of contrasting oceanographic conditions on the spatiotemporal distribution of 30 809 Mediterranean cephalopod paralarvae. Hydrobiologia 749: 114. 31 32 810 33 34 811 Figure captions 35 36 812 Figure 1 . (ad) Illex coindetii . a) Ventral view, aged 354 h and incubated at 17 ºC. b) Dorsal 37 813 view, aged 427 h and incubated at 17 ºC. c) Lateral view, aged 262 h and incubated at 21 ºC. d) 38 814 Ventral view of an individual with expanded chromatophores, aged 236 h and incubated at 17 39 815 ºC. (eh) Todarodes sagittatus . e) Ventral view, aged 364 h and incubated at 15 ºC. f) Dorsal 40 41 816 view, aged 358 h and incubated at 17 ºC. g) Lateral view, aged 360 h and incubated at 17 ºC. h) 42 817 Ventral view of an individual with expanded chromatophores, aged 336 h and incubated at 17 43 818 ºC. (il) Todaropsis eblanae . i) Ventral view, aged 475 h and incubated at 17 ºC. j) Dorsal view, 44 819 aged 500 h and incubated at 15 ºC. k) Lateral view, aged 498 h and incubated at 15 ºC. l) 45 820 Ventral view of an individual with expanded chromatophores, aged 649 h and incubated at 15 46 821 ºC. ac, eg, ik, specimens anaesthetized with ethanol, which potentially causes chromatophore 47 48 822 contraction. d, h, l, individuals without anaesthesia. Scale bars: 1 mm. 49 50 823 51 52 824 Figure 2 . (ad) Illex coindetii . a) SEM image of the ventral view of the head, aged 270 h and 53 825 incubated at 17 ºC. b) SEM image of the arm I sucker, aged 270 h and incubated at 17 ºC. c) 54 826 SEM image of a proboscis sucker, aged 454 h and incubated at 17 ºC. d) Detail of the ventral 55 827 skin of an anaesthetized specimen, aged 329 h and incubated at 21 ºC. (eh) Todarodes 56 828 sagittatus . e) SEM image of proboscis tip, showing the differences between the lateral and 57 829 medial suckers, aged 361 h and incubated at 15 ºC. f) SEM image of the left arm I sucker, aged 58 59 60 20

Page 21 of 31 Zoological Journal of the Linnean Society

1 2 3 830 361 h and incubated at 15 ºC. g) SEM image of a proboscis sucker, aged 361 h and incubated at 4 831 15 ºC. h) Detail of the ventral skin of an anaesthetized specimen, aged 383 h and incubated at 15 5 832 ºC. (il) Todaropsis eblanae . i) SEM image of the ventrolateral view of the head of a paralarva, 6 833 aged 477 h and incubated at 15 ºC, the III and IV pairs of arm stumps are visible. j) SEM image 7 of the sucker of the left arm I, aged 475 h and incubated at 15 ºC. k) SEM image of a proboscis 8 834 9 835 sucker, aged 477 h and incubated at 15 ºC. l) Detail of the ventral skin of an anaesthetized 10 836 specimen, aged 498 h and incubated at 15 ºC. Scale bars: a, e, i, 100 µm; b, c, f, g, k, 20 µm; j, 11 837 50 µm; d, h, l, 0.5 mm. 12 13 838 14 15 839 Figure 3 . Schematic drawing of the chromatophore and photophore pattern of the seven North 16 840 eastern Atlantic rhynchoteuthions. (ac) Illex coindetii . a) Ventral view. b) Dorsal view. c) 17 841 Lateral view. (df) Todarodes sagittatus . d) Ventral view. e) Dorsal view. f) Lateral view. (gi) 18 Todaropsis eblanaeFor. g) Ventral Review view. h) Dorsal view. Only i) Lateral view. (jk) Ommastrephes 19 842 20 843 bartramii . j) Ventral view. k) Dorsal view. (lm) Hyaloteuthis pelagica . l) Ventral view. m) 21 844 Dorsal view. n) Sthenoteuthis pteropus , ventral view. o) Ornithoteuthis antillarum , ventral 22 845 view. Grey chromatophores of af depict those seen in both dorsal and ventral views. Concentric 23 846 black and white circles on ln depict ocular and intestinal photophores. Chromatophore pattern 24 847 of af is based on the mode of the chromatophore pattern (see table 2); jk based on Sweeney et 25 26 848 al. (1992), Young & Hirota, (1990), Sakurai et al. (1995) and Vijai et al. (2015). Photophore and 27 849 chromatophore pattern of lm based on Harman & Young (1985) and Sweeney et al. (1992). 28 850 Photophore pattern of n based on Sweeney et al. (1992), of o based on Sweeney et al (1992) and 29 851 Diekmann et al. (2002). The chromatophore pattern of S. pteropus (n) and O. antillarum (o) are 30 852 not known. 31 32 853 Figure 4 . Main taxonomic characters used to identify wildcollected rhynchoteuthions by the 33 854 dichotomous key provided here. (ac) lllex coindetii . a) 2.4 mm DML. SEM image of the 34 35 855 proboscis suckers, showing a single row of pegs. bc) 1.03 mm DML. b) Dorsal view of the 36 856 head showing the chromatophore pattern 1 + 3. c) Ventral view of the head showing one row of 37 857 two chromatophores. (d) Todarodes sagittatus , 2.20 mm DML. SEM image of the proboscis 38 858 suckers showing lateral suckers larger than the medial sucker. (eg) Sthenoteuthis pteropus , 7.71 39 859 mm DML. e) Dorsal view of the specimen. f) Ventral view of the head showing the ocular 40 860 photophores. g) Ventral view of the specimen with the mantle opened to show the two equally 41 42 861 sized intestinal photophores. Scale bars: a, d: 0.1 mm; bc: 0.5 mm; eg: 1 mm. 43 44 862 45 46 863 Tables 47 48 49 50 51 52 53 54 55 56 57 58 59 60 21

Zoological Journal of the Linnean Society Page 22 of 31

1 2 3 4 5 6 Table 1. Morphometric parameters measured (mm) and morphometric parameter ratios in relation to the VML (%). Abbreviations: VML: ventral mantle 7 length, DML: dorsal mantle length, TL: total length, TL w P: total length without the proboscis, HL: head length, HW: head width, ED: eye diameter, FuL: 8 9 funnel length, PW: proboscis width at the base, AIIL: Arm II length, DMLI: dorsal mantle length index, TLI: total length index, TL w PI: total length without 10 the proboscis index, HLI: head length index, HWI: head width index, EDI: eye diameter index, FuLI: funnel length index, PWI: proboscis width at the base 11 index, AIILI: Arm II Length Index. 12 13 For Review Only 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 22 43

44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Page 23 of 31 Zoological Journal of the Linnean Society

1 2 3 4 5 6 7 8 9 10 11 12 Species VML DML TL TL w P HL HW ED FuL PW AIIL n 13 Measures For Review Only 14 Illex coindetii Average 1.41 1.52 2.45 2.24 0.52 0.70 0.22 0.73 0.16 0.18 52 15 SD 0.15 0.10 0.21 0.20 0.06 0.06 0.03 0.07 0.02 0.06 16 Range 1.091.82 1.401.71 1.932.93 1.652.65 0.390.67 0.590.87 0.160.28 0.550.91 0.130.19 0.100.30 17 18 Todarodes Average 1.64 1.80 2.63 2.44 0.61 0.79 0.24 0.47 0.16 0.20 29 19 sagittatus SD 0.12 0.14 0.20 0.19 0.08 0.07 0.03 0.04 0.01 0.04 20 Range 1.251.81 1.441.96 2.032.95 1.902.73 0.440.75 0.640.92 0.160.30 0.390.55 0.140.18 0.110.26 21 Todaropsis 22 eblanae Average 2.16 2.19 3.60 3.26 0.83 1.00 0.34 0.98 0.21 0.41 36 23 SD 0.11 0.14 0.24 0.17 0.05 0.06 0.03 0.10 0.01 0.05 24 Range 1.93 2.43 1.92 2.48 3.03 4.12 2.89 3.55 0.71 0.92 0.89 1.10 0.28 0.41 0.74 1.15 0.19 0.25 0.32 0.52 25

26 Ratios DMLI TLI TL w PI HLI HWI EDI FuLI PWI AIILI 27 Illex coindetii Average 107.43 175.53 161.10 37.56 50.46 15.73 53.17 11.75 12.41 42 28 SD 6.69 18.68 16.39 5.88 5.77 2.32 6.61 1.57 3.44 29 95.33 143.86 131.25 27.89 40.36 30 Range 116.43 243.38 219.05 55.22 64.18 11.2921.20 38.0265.82 9.2116.19 6.9822.60 31 Todarodes Average 107.93 162.50 149.93 36.90 49.05 15.16 28.45 9.67 12.38 15 32 sagittatus SD 6.98 8.67 7.01 4.55 4.86 2.12 2.90 0.98 2.52 33 95.94 147.15 136.93 28.82 41.18 34 Range 123.27 182.38 163.52 45.73 61.12 9.4119.50 23.3533.54 8.1411.64 6.4717.12 35 Todaropsis 36 eblanae Average 101.97 168.28 151.72 38.50 46.49 16.08 45.84 9.77 19.00 31 37

38 SD 6.28 10.51 8.40 3.09 3.33 1.42 4.85 0.81 2.50 39 89.78 142.86 128.44 33.78 39.18 40 Range 118.93 183.58 166.67 46.11 52.24 13.78 19.69 35.48 53.33 8.26 11.65 14.75 24.17 41 42 23 43

44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Zoological Journal of the Linnean Society Page 24 of 31

1 2 3 4 5 6 7 8 Table 2. Schematic chromatophore pattern of I. coindetii , T. sagittatus and T. eblanae . Numbers indicate rows of chromatophores in an anteroposterior axis 9 (see Material and Methods for further details). N/A, not applicable. Abbreviations: Ven, Ventral; Dor, Dorsal; Lat, Latera; AI, Arm I; AII, Arm II; Prob, 10 11 Proboscis. 12 13 HeadFor ReviewArm Crown OnlyMantle 14 15 Species Ven Dor Lat AI AII Prob Ven Dor Lat n 16 1 1 2 3 N/A N/A N/A N/A 1 2 3 4 5 6 1 2 3 4 5 1 2 3 17 Illex 18 coindetii Average 1.9 1.0 3.0 N/A 0.2 1.8 0.2 3.4 4.3 2.4 3.7 3.2 1.8 2.0 1.0 2.6 2.4 0.9 1.0 3.1 N/A 3.8 52 19 SD 0.4 0.1 0.3 N/A 0.5 2.5 0.4 3.7 0.8 1.4 1.2 1.3 1.3 0.1 1.0 1.1 0.9 1.3 0.0 0.5 N/A 0.7 20 Range 02 1.0 24 N/A 02 010 01 016 36 06 06 05 04 12 03 05 14 04 1 24 N/A 35 21 Mode 2 1 3 N/A 0 3 0 2 4 2 4 3 3 2 2 3 3 0 1 3 N/A 4 22 Todarodes 23 sagittatus Average 0 1.0 3.0 N/A 0.0 0.0 0.0 0.0 2.6 5.1 3.9 1.8 N/A N/A 3.1 1.0 3.1 0.2 0.0 1.9 N/A 2.5 29 24 SD 0 0.2 0.5 N/A 0.0 0.0 0.0 0.0 0.8 1.1 1.1 0.6 N/A N/A 0.9 0.9 1.0 0.5 0.2 0.2 N/A 0.6 25 Range 0 01 14 N/A 0 0 0 0 25 36 25 03 N/A N/A 25 04 25 02 01 12 N/A 13 26 27 Mode 0 1 3 N/A 0 0 0 0 2 6 5 2 N/A N/A 3 1 4 0 0 2 N/A 3 28 Todaropsis 1.8 1.0 1.9 2.7 0.0 0.0 0.0 0.0 6.6 4.5 1.3 0.8 4.2 0.8 3.6 1.7 2.7 2.3 1.9 3.0 2.3 3.2 43 29 eblanae Average 30 SD 0.5 0.0 0.5 0.7 0.0 0.0 0.0 0.0 0.8 1.4 1.5 1.4 1.4 0.5 1.2 1.5 1.7 1.7 0.6 0.5 0.8 0.7 31 Range 02 1 02 03 0 0 0 0 59 16 05 05 06 02 15 05 05 05 13 24 14 24 32 Mode 2 1 2 3 0 0 0 0 7 5 0 0 5 2 3 0 5 0 2 3 2 3 33 34 35 36 37 38 39 40 41 42 24 43

44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Page 25 of 31 Zoological Journal of the Linnean Society

1 2 3 4 5 6 7 8 Table 3. Rhynchoteuthion paralarvae species according to the ratio of sucker sizes , number and arrangement of pegs of the proboscis and arm suckers. The 9 ratio of sucker sizes is indicated only when greater than 1. 10 11

12 Proboscis suckers Arm suckers Species 13 ParalarvalFor ReviewInternal External FreeOnly Internal External Free 14

15 type Ratio pegs pegs pegs pegs pegs pegs n 16 Illex coindetii Average C 12.9 N/A 6.4 12.1 12.2 0.5 11 17 SD 1.4 N/A 2.8 1.0 1.4 0.5 18 Range 1014 N/A 112 1014 1014 01 19 Mode 14 N/A 8 12 11 0 20 Average B 1.2:1 18.0 17.0 0.0 16.2 14.0 0.0 7 21 Todarodes 22 sagittatus SD 3.0 3.0 0.0 1.7 2.1 0.0 23 Range 1224 1222 0 1419 917 0 24 Mode 18 14 0 16 14 0 25 Todaropsis eblanae Average C 18.2 22.0 0.5 17 20.6 1.6 11 26 SD 1.0 2.2 1.0 0.8 1.5 2.3 27 Range 1624 1727 03 1518 1823 08 28 29 Mode 18 22 0 17 20 0 30 31 32 33 34 35 36 37 38 39 40 41 42 25 43

44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Zoological Journal of the Linnean Society Page 26 of 31

1 2 3 4 5 6 Table 4. Wild rhynchoteuthion from the NE Atlantic (most from the NW Mediterranean) identified to the species level using the identification key described 7 here. The collecting information and the specific diagnostic characters used in the identification are shown. The measurements were performed on preserved 8 9 paralarvae, so an undetermined shrinkage is expected in relation to live paralarvae. * Measurements taken from a defrosted specimen. 10 11 Cruise, station Seafloor and geographic Depth depth Date and DML VML 12 Species coordinates (m) (m) hour n (mm) (mm) Diagnostic characters 13 Illex coindetii CACO 1, st. 9, 081 For88 7192003, Review 1 No 0.97 Type COnly rhynchoteuthion, no photophores, only one row of pegs in proboscis 14 40.42 º N 1.05 E 12:31 data suckers. 15 Illex coindetii CACO 2, st. 5, 0120 125 912 2003, 1 2.43 2.18 Type C rhynchoteuthion, no photophores, only one row of pegs in proboscis 16 40.23º N 1.16º E 2:35 suckers (Fig. 4a) . 17 Illex coindetii CACO 2, st. 9, 080 88 9122003, 1 2.15 2.13 Type C rhynchoteuthion, no photophores, only one row of pegs in proboscis 18 40.43º N 1.06º E 13:10 suckers. 19 Illex coindetii CACO 3, st. 4, 085 95 624 2004, 1 1.29 1.21 Type C rhynchoteuthion, no photophores, only one row of pegs in proboscis 20 40.27º N 1.02º E 0:47 suckers. 21 Illex coindetii CACO 3, st. 8, 0100 107 6242004, 2 1.06, 1.02, no Type C rhynchoteuthion, no photophores, only one row of pegs in proboscis 22 40.38º N 1.21º E 12:03 1.16 data suckers. 23 Illex coindetii CACO 3, st. 23, 065 75 6262004, 1 2.02 2.06 Type C rhynchoteuthion, no photophores, only one row of pegs in proboscis 24 40.95º N, 1.12º E 00:48 suckers. Illex coindetii CACO 3, st. 24, 0100 110 6262004, 1 1.33 1.24 Type C rhynchoteuthion, no photophores, only one row of pegs in proboscis 25 40.90º N 1.26º E 05:38 suckers 26 Todarodes CACO 3, st. 43, 0125 135 6282004, 2 1.68, 1.60, Type B rhynchoteuthion, no photophores, two rows of pegs in proboscis 27 sagittatus 41.49º N, 2.70º E 03:40 2.20 2.09 suckers (Fig. 4d) . 28 Illex coindetii CACO 3, st. 55, 0200 765 6292004, 1 2.84 3.06 Type C rhynchoteuthion, no photophores, only one row of pegs in proboscis 29 41.75º N 3.50º E 13:47 suckers. 30 Todarodes CACO 4, st. 34, 0200 625 725 2004 2 0.77, 0.94, Type B rhynchoteuthion, no photophores, two rows of pegs in proboscis 31 sagittatus 41.11º N 2.20º E 1.03 0.93 suckers. 32 Illex coindetii or 33 Todaropsis LLUÇ3, 4.3, 41.00 0186 No data 4061999, 1 No 1.97 Type C rhynchoteuthion, no photophores, proboscis suckers not visible. 34 eblanae °N 1.42° E 21:00 data 35 Illex coindetii FishJelly, 41.38´ N 010 No data 10212014, 1 1.03 0.94 Type C rhynchoteuthion, no photophores. Chromatophore pattern visible in this 36 2.21´ E 11:00 specimen and compatible with I. coindetii and not with T. eblanae (Fig. 4bc) . 37 Stenoteuthis MAFIA, PEL07, 0100 4245 417 2015, 1 7.71* 7.11* Type C rhynchoteuthion, two equal sized intestinal photophores (Fig. 4g ) and pteropus 7.15 ºN 23.97 ºW 23:28 an ocular photophore in each eye (Fig. 4f) . 38 39 40 41 42 26 43

44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Page 27 of 31 Zoological Journal of the Linnean Society

1 2 3 4 5 6 7 8 9 10 11 12 13 For Review Only 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 27 43

44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Zoological Journal of the Linnean Society Page 28 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 For Review Only 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 Figure 1. (a-d) Illex coindetii. a) Ventral view, aged 354 h and incubated at 17 ºC. b) Dorsal view, aged 427 46 h and incubated at 17 ºC. c) Lateral view, aged 262 h and incubated at 21 ºC. d) Ventral view of an 47 individual with expanded chromatophores, aged 236 h and incubated at 17 ºC. (e-h) Todarodes sagittatus. e) Ventral view, aged 364 h and incubated at 15 ºC. f) Dorsal view, aged 358 h and incubated at 17 ºC. g) 48 Lateral view, aged 360 h and incubated at 17 ºC. h) Ventral view of an individual with expanded 49 chromatophores, aged 336 h and incubated at 17 ºC. (i-l) Todaropsis eblanae. i) Ventral view, aged 475 h 50 and incubated at 17 ºC. j) Dorsal view, aged 500 h and incubated at 15 ºC. k) Lateral view, aged 498 h and 51 incubated at 15 ºC. l) Ventral view of an individual with expanded chromatophores, aged 649 h and 52 incubated at 15 ºC. a-c, e-g, i-k, specimens anaesthetized with ethanol, which potentially causes 53 chromatophore contraction. d, h, l, individuals without anaesthesia. Scale bars: 1 mm. 54 55 173x197mm (300 x 300 DPI) 56 57 58 59 60 Page 29 of 31 Zoological Journal of the Linnean Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 For Review Only 19 20 21 22 23 24 25 26 27 28 29 30 31 32 Figure 2. (ad) Illex coindetii. a) SEM image of the ventral view of the head, aged 270 h and incubated at 17 33 ºC. b) SEM image of the arm I sucker, aged 270 h and incubated at 17 ºC. c) SEM image of a proboscis 34 sucker, aged 454 h and incuba ted at 17 ºC. d) Detail of the ventral skin of an anaesthetized specimen, aged 329 h and incubated at 21 ºC. (eh) Todarodes sagittatus. e) SEM image of proboscis tip, showing the 35 differences between the lateral and medial suckers, aged 361 h and incubated at 15 ºC. f) SEM image of the 36 left arm I sucker, aged 361 h and incubated at 15 ºC. g) SEM image of a proboscis sucker, aged 361 h and 37 incubated at 15 ºC. h) Detail of the ventral skin of an anaesthetized specimen, aged 383 h and incubated at 38 15 ºC. (il) Todaropsis eblanae. i) SEM image of the ventrolateral view of the head of a paralarva, aged 477 39 h and incubated at 15 ºC, the III and IV pairs of arm stumps are visible. j) SEM image of the sucker of the 40 left arm I, aged 475 h and incubated at 15 ºC. k) SEM image of a proboscis sucker, aged 477 h and 41 incubated at 15 ºC. l) Detail of the ventral skin of an anaesthetized specimen, aged 498 h and incubated at 42 15 ºC. Scale bars: a, e, i, 100 µm; b, c, f, g, k, 20 µm; j, 50 µm; d, h, l, 0.5 mm.

43 193x143mm (300 x 300 DPI) 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Zoological Journal of the Linnean Society Page 30 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 For Review Only 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 Figure 3. Schematic drawing of the chromatophore and photophore pattern of the seven North-eastern 35 Atlantic rhynchoteuthions. (a-c) Illex coindetii. a) Ventral view. b) Dorsal view. c) Lateral view. (d-f) 36 Todarodes sagittatus. d) Ventral view. e) Dorsal view. f) Lateral view. (g-i) Todaropsis eblanae. g) Ventral 37 view. h) Dorsal view. i) Lateral view. (j-k) Ommastrephes bartramii. j) Ventral view. k) Dorsal view. (l-m) 38 Hyaloteuthis pelagica. l) Ventral view. m) Dorsal view. n) Sthenoteuthis pteropus, ventral view. o) 39 Ornithoteuthis antillarum, ventral view. Grey chromatophores of a-f depict those seen in both dorsal and ventral views. Concentric black and white circles on l-n depict ocular and intestinal photophores. 40 Chromatophore pattern of a-f is based on the mode of the chromatophore pattern (see table 2); j-k based 41 on Sweeney et al. (1992), Young & Hirota, (1990), Sakurai et al. (1995) and Vijai et al. (2015). Photophore 42 and chromatophore pattern of l-m based on Harman & Young (1985) and Sweeney et al. (1992). 43 Photophore pattern of n based on Sweeney et al. (1992), of o based on Sweeney et al (1992) and Diekmann 44 et al. (2002). The chromatophore pattern of S. pteropus (n) and O. antillarum (o) are not known. 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Page 31 of 31 Zoological Journal of the Linnean Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 For Review Only 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 Figure 4. Main taxonomic characters used to identify wild-collected rhynchoteuthions by the dichotomous 46 key provided here. (a-c) lllex coindetii. a) 2.4 mm DML. SEM image of the proboscis suckers, showing a 47 single row of pegs. b-c) 1.03 mm DML. b) Dorsal view of the head showing the chromatophore pattern 1 + 48 3. c) Ventral view of the head showing one row of two chromatophores. (d) Todarodes sagittatus, 2.20 mm 49 DML. SEM image of the proboscis suckers showing lateral suckers larger than the medial sucker. (e-g) 50 Sthenoteuthis pteropus, 7.71 mm DML. e) Dorsal view of the specimen. f) Ventral view of the head showing 51 the ocular photophores. g) Ventral view of the specimen with the mantle opened to show the two equally- sized intestinal photophores. Scale bars: a, d: 0.1 mm; b-c: 0.5 mm; e-g: 1 mm. 52

53 190x220mm (300 x 300 DPI) 54 55 56 57 58 59 60