首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Interactions of two tubicolous, deposit-feeding polychaetes, Clymenella torquata (Leidy) and Polydora ligni Webster, with a filter-feeding bivalve, Gemma gemma (Totten), were examined to test the functional group hypothesis that small, filter-feeding bivalves with large young should attain high densities with tube-builders of any trophic mode because the large young escape predation by tube-builders. The prediction was not supported by Polydora and Gemma, since the two species have not been reported to coexist in high densities in nature. In laboratory experiments, Polydora reduced recruitment and adult survival of Gemma relative to a control. Polydora preyed on juvenile clams and reduced the time which adult clams had their siphons in the feeding position, relative to a control.For Clymenella and Gemma, the functional group hypothesis made the correct prediction (i.e. the two species are reported to coexist in high densities in nature), but for the wrong reason. The size of Gemma's offspring is not related to the intensity of Clymenella predation. Clymenella does not ingest Gemma since the worm does not usually feed from the zone occupied by the clam. In contrast to Polydora, Clymenella did not reduce the time which clams fully extend and open their siphons relative to a control. In laboratory experiments Clymenella increased adult bivalve growth, survival after one year, and recruit size after one month. Clam life history characteristics were similar in the Clymenella and Clymenella plus Polydora treatments because the number of surviving Polydora was reduced by a negative interaction with Clymenella.The tube-builders had different effects on the filter-feeder, Gemma, because these tube-building species differ in feeding position, diet and behavior. Thus, the functional group hypothesis was too general to predict and explain the species interactions and distributions.  相似文献   

2.
Although it is recognized that many species of benthic invertebrates continue to disperse after settlement, particularly in soft-bottom habitats, the scale over which movements of juveniles occur is not well known. This study combined laboratory flume experiments assessing the effects of clam size, species, and water velocity on rates and distances of dispersal of three species of juvenile bivalves with field measurements of loss rates and distances of dispersal of transplanted bivalves in the Navesink River estuary in New Jersey, USA. Dispersal distances measured in the laboratory ranged from an average of 1.6 to 40 cm h− 1 depending on clam size, species, and flow speed. Distances and likelihood of dispersal were generally greater for Mya arenaria than for Mercenaria mercenaria or Gemma gemma, although differences between species were not consistent. As predicted, smaller (1.3 mm) M. arenaria tended to disperse more than larger (3.7 mm) ones, although no significant differences were detected between two sizes (1.8 and 3.4 mm) of M. mercenaria. The similarity of the erosion thresholds of dead clams across sizes and species suggests that burrowing behaviour plays an important role in determining variation in dispersal due to clam size and species. In the field, densities of clams (M.arenaria and M.mercenaria) were reduced to half of that in controls after 3.5-5 h, indicating high levels of dispersal and/or mortality. Some individuals were recovered up to 50 cm away from their initial locations. Overall, our results suggest that dispersal distances of these three species due to bedload transport are likely to be on the order of centimeters per hour. Although these dispersal distances are small, such movements are likely to occur frequently due to tidal currents and, consequently, may have profound impacts on patterns of abundance and distribution.  相似文献   

3.
The present study documents for the first time shell use by juvenile fiddler crabs in the salt marsh. Twenty visits were made to six salt marsh sites at Tybee Island, Georgia between 2007 and 2009. One hundred empty Littorina irrorata shells were collected at each site on each field trip. Juvenile carapace width was measured, crabs sexed, and species identification completed using RFLP analysis. Shell use of up to 79% was observed. Two species of fiddler crabs were found in empty shells, Uca pugnax and U. pugilator. U. pugnax was the dominant species at all sites representing 62-84% of the juvenile fiddler crab population. Juvenile sex ratios were female-biased (1.7:1) at all six sites. Juvenile size did not vary significantly between species but males of both species were significantly larger than females. Size frequency distribution of carapace width revealed that shell use varied with size and sex. In the 3 to 4 mm size class, juvenile females outnumbered juvenile males in empty L. irrorata shells while in the 5 to 6 mm size class and greater, juvenile males outnumbered juvenile females in shells. Significantly more juvenile fiddler crabs were found in empty shells during flood than ebb tide at 3 of the sites. This discovery illuminates the resourcefulness of juvenile fiddler crabs and provides another mechanism that might enhance survival.  相似文献   

4.
Although originally described as a separate species, Patelloida (Chiazacmea) lampanicola Habe has subsequently been regarded as a “form” of P. (C.) pygmaea (Dunker). The conical shape of the shell in the former was assumed to result from the settlement and growth of some individuals of the population upon shells of intertidal Batillaria (Potamididae).This paper reports upon an analysis of a population of limpets resident on a small sand flat in Hong Kong and it is concluded that Patelloida (C.) lampanicola is distinct from P. (C.) pygmaea. This conclusion is based upon differences in radula teeth structure, shell morphometrics and behaviour. Furthermore it has been demonstrated that the earlier confusion surrounding these species results from a very nice example of selective site segregation. P. (C.) pygmaea inhabits stones and empty shells (in this region of its total range) embedded in the sand but can occasionally does reside, up to a length of 4.2 mm, upon Batillaria shells. Thereafter the flattened form of the limpet shell, on a round substratum, presumably makes such individuals more susceptible to either predation or dislodgement. The high-coned Patelloida (C.) lampanicola, on the other hand, almost exclusively colonizes living Batillaria (particularly B. zonalis and B. multiformis) shells but can, with little or no significant alteration in form, also colonize stones. Patelloida (C.) lampanicola has a positive behavioural response towards Batillaria and is clearly living in symbiotic association with the potamid. The benefits accrued by both are discussed.Thus, the limited hard substratum of a primarily soft shore environment has been segregated by these two limpets; one occupies the stones and oyster shells, the other the dense cover of epifaunal potamid snails.  相似文献   

5.
While piscivory is common in many fishes, there are few accounts for the fundulid, Fundulus heteroclitus (mummichog). We suspected that this species might be involved in several forms of piscivory including predation, cannibalism and scavenging. To evaluate these possibilities we conducted several laboratory experiments and field observations in its primary habitat, salt marshes. We found that digestion of larvae and small juveniles of F. heteroclitus was fast (< 1 h) and this makes detection of small fish prey difficult for any form of piscivory. In addition, laboratory experiments (one-on-one encounters, feeding on dead prey) indicated that both cannibalism and scavenging of conspecific prey were possible. Field observations (n = 2449 stomachs from fish 12-106 mm TL over four years at two salt marshes) also suggested that piscivory occurred frequently (4.3-24.7% of stomachs examined) based on the detection of numerous fish hard parts, especially pharyngeal jaws. These structures allowed us to determine that most prey were F. heteroclitus and to back-calculate the size of the prey and thus the piscivore/prey size relationship. As a result, it appears that both cannibalism (0.2-9.1% of stomachs) and scavenging (0.5-9.9%) are common feeding modes of F. heteroclitus in salt marshes and, subsequently, modes of energy transfer for salt marshes in general. We suggest that future studies of fish feeding consider that cannibalism and scavenging may be frequent possibilities when fish remains are detected in stomach contents.  相似文献   

6.
The new species Polhillia ignota Boatwr. is described. This species is known only from two collections, one between Vredenburg and Saldanha and the other close to Porterville. The new species is most similar to P. obsoleta, which is known only from a few localities around Worcester, in its narrow, sericeous leaflets and flowers of equal size, but differs in its flattened mature leaves with larger leaflets (up to ± 12 mm long), longer pseudo-peduncles (up to ± 15  mm long), denser inflorescences (with up to four flowers), shorter pedicels (1–2 mm long) and non-auriculate wing- and keel petals.  相似文献   

7.
The thermoregulatory behavior of the wavy turban snail Megastrea (Lithopoma) undosa was determined in a horizontal thermal gradient and was 16.31 in day cycle and 14.4 °C in night cycle. Displacement velocity of adults was 29.3±4.2 cm h−1 during the light phase and 26.1±3.2 cm h−1 during the dark phase. The critical thermal maxima of the wavy turban snail were determined. As a measure of thermal tolerance, snails were subjected to increasing water temperatures at a rate of 1 °C every 30 min until they were detached from the substrate. The critical thermal maximum at 50% was 29.7 °C.  相似文献   

8.
Littorina acutispira Smith, a minute gastropod of < 3 mm shell height, lives at great densities in pools and on rock-surfaces at the highest levels of sea-shores in New South Wales. Populations from pools and dry areas were sampled on two shores for 18 months to investigate seasonal changes in density, size-structure, rates of growth and reproductive biology of the snails. Densities of snails increased between February and May, due to an influx of juveniles, and then declined until the following February, when they increased again. The decrease in density was due to the death of the largest snails at the end of summer, and the mortality of medium-sized snails between June and January. Longevity was estimated as 1–2 yr, but most individuals died by ≈ 16 months from settlement on the shore. Newly-settled snails grew to merge in size with those of the previous year's population by winter. During the summer months, the rate of growth of snails from a sheltered shore was greater than that of snails on a shore exposed to wave-action. Laboratory experiments revealed that this could be attributed to the presence of better quality food, or food in greater abundance on the sheltered shore, compared with the exposed shore.During winter months, but not during the summer, snails from dry areas grew more slowly than those from pools. An experiment demonstrated that some snails from dry areas might be able to compensate for reduced periods of feeding by being able to feed faster when submersed. This could not explain the differences in natural rates of growth.L. acutispira bred from October–November to March–April. Spawning in the laboratory was greatest during late summer (January to March). The percentage of mature oocytes in the gonads was small in winter and increased in early summer. Among the largest-sized snails, females outnumbered males. Two experiments, on unsexed and pre-sexed snails, demonstrated that the biased sex-ratio of the largest snails was due to faster growth by females.There was a greater density of snails on the exposed shore, which was correlated with the presence of barnacles. When barnacles were removed from experimental areas, the density of the snails declined within 24 h. This suggested that barnacles provided a refuge from wave-shock, rather than shelter from desiccation or high temperature. In laboratory experiments, snails were exposed to higher temperatures and less humidity than they would normally encounter on the shore. There was negligible mortality of small or large snails after 24 h of these conditions.This minute species grows quickly, recruits annually and has a short life-span. This type of life-history is discussed in comparison with similar small species from other habitats.  相似文献   

9.
The littorinid snail, Cenchritis muricatus, inhabits supralittoral vertical rocky walls along Caribbean shores, at times surpassing 14 meters above mean sea level. As the sole macrofaunal representative of this habitat, this marine gastropod presumably experiences extraordinary conditions of thermal load and desiccation. In order to understand the effect of behavioral choices on periwinkle survivorship and growth, snail distribution, microhabitat utilization, and crawling speed were documented in St John (US Virgin Islands). In general, snails rarely inhabited open surfaces; instead, periwinkles were commonly observed in microhabitats that may reduce water and heat stresses (e.g., > 75% in crevices and depressions). Snails found on occasional buttonwood trees (Conocarpus erectus) were larger than elsewhere. Although typically found in repose, C. muricatus crawling speed on moist, shaded, and smooth substrata averaged more than 3 cm.min- 1, but did not vary with slope. Repeated mark-recapture of tagged periwinkles exhibited high recovery rates (ca. 35% after 4 yr), absence of mortality, and a projected cessation of growth at 16.5 mm (shell height). Nearly 10% of marked individuals were recaptured every year. Dead, tagged snails were never noted; indeed, seven individuals were only recovered once, a full 4 yrs after release. Site-specific growth rates were absent. Projections using von Bertalanffy growth functions (VBGF) suggest that periwinkles will require 15+ years to achieve the maximum shell height. These VBGF models cannot address extraordinary individuals reaching 22 mm. C. muricatus's remarkable supralittoral distribution may be explained by physiological tolerance, selection of microhabitats, lack of predators and long lifespan.  相似文献   

10.
The suspension-feeding slippersnail Crepidula convexa is commonly associated with hermit crabs (Pagurus longicarpus) living in periwinkle shells (Littorina littorea) at our study site in Nahant, MA, USA. In 15 field surveys conducted at Nahant in 2000, 2001 and 2003, we found that (1) more than 61.8% of individuals of C. convexa resided on shells occupied by hermit crabs, as opposed to the shells of live periwinkles, empty periwinkle shells or other solid substrates; (2) an average of 8.3% of hermit crabs carried at least one individual of C. convexa; and (3) 39.1-75.0% of hermit crabs carrying C. convexa were carrying “large” individuals (snails with wet weight >10% of the weight of the periwinkle shells they occupied). However, it is unlikely that individuals of C. convexa seek out shells occupied by hermit crabs to colonize, and they showed no preference for empty periwinkle shells over other solid substrates in the laboratory. Moreover, in the laboratory the hermit crabs preferentially occupied intact shells bearing individuals of C. convexa only when the alternatives were shells that had been drilled by naticid snails. Thus, neither party preferentially associates with the other: rather, extensive predation by naticid snails on periwinkles at Nahant appears to limit the availability of suitable shells for the hermit crabs, forcing them to inhabit shells bearing “large” individuals of C. convexa. Individuals of C. convexa may benefit from this inadvertent association with hermit crabs: by facilitating snail dispersal, transport by hermit crabs should reduce the potential for inbreeding, an important consideration for a species that lacks free-living larvae in its life history.  相似文献   

11.
Current study determined, in sows, the accuracy of ultrasonography for in vivo (n = 8) and ex vivo (n = 7) evaluation of corpora lutea (CLs) and follicles ≥1.5 mm in size, by comparison with macroscopic findings in sliced ovaries. The accuracy for ex vivo detection of follicles increased with follicle size (P < 0.05), being low for 1.5-1.9 mm follicles (65.9%) and higher for ≥6 mm follicles (93.3%); differences between ultrasonographic and macroscopic observations were significant only for follicles smaller than 3.9 mm (P < 0.05), due to underestimation. Ex vivo observation succeeded to detect presence or absence of CLs in all the ovaries; the efficiency for determining the exact number of CLs being 94.4%. The accuracy for in vivo detection of follicles also increased with follicle size (P < 0.05), dropping to values lower than 40% for 1.5-1.9 mm follicles; therefore, there were significant differences between ultrasonographic and macroscopic observations (P < 0.05). On the other hand, accuracy remained around 92% for ≥6 mm follicles. Ultrasonography was useful again for detecting presence of CLs in all the ovaries; the efficiency for determining CLs number reached 86.7%, due to underestimation in ovaries with higher number of CLs (P < 0.05). Overall, there were no significant differences when comparing the accuracy of ex vivo and in vivo scannings for determination neither of the number of follicles in each size-category larger than 1.9 mm nor of the presence of ovulations or of the CLs number in each ovary. In conclusion, the use of ultrasonography allows an accurate detection of the presence and number of CLs and follicles ≥2 mm of diameter in sows, without significant differences between in vivo and ex vivo observations.  相似文献   

12.
Kuris A. M. 1980. Effect of exposure to Echinostoma liei miracidia on growth and survival of young Biomphalaria glabrata snails. International Journal for Parasitology10: 303–308. Exposure to miracidia of Echinostoma liei resulted in increased mortality and reduced growth of 1–2 mm albino Biomphalaria glabrata snails whether or not the snails became infected. Growth rates for infected and exposed but uninfected snails were significantly more variable than growth rates of unexposed snails. Retarded growth and increased mortality were detected as rapidly as seven to nine days after exposure. Neither growth nor survivorship of 4–6 mm snails was altered upon exposure to or infection by E. liei.  相似文献   

13.
To document the relative importance of meiofauna as prey for juvenile Crangon crangon and Carcinus maenus, short interval (1.5-2 h) collections were made in the muddy Lynher Estuary (Plymouth, Great Britain) and in the sandy-bottom Ythan Estuary (Aberdeenshire, Scotland) in 1990. Gut passage times of Crangon fed flaked fish food and fluorescent tracer in the laboratory at 13 °C ranged from 4 to 20 h. Wild shrimp exhibited feeding periodicity, with guts fullest during high tide in both locations. Visual and immunological gut contents analyses revealed that meiofaunal nematodes and harpacticoid copepods were present only in recently settled shrimp from 8 to 12 mm total length on muddy bottoms. Larger shrimp collectively consumed up to 33 different macrobenthic prey types. Shrimp were fullest at night (mean gut contents weight = 8% wet body weight, Lynher) or at dawn (6%, Ythan). The Lynher Carcinus gut contents—from animals 8 to 30 mm carapace width, examined visually only—contained mostly fluids, green benthic algae, sediment particles, and masses of unidentifiable prey remains plus digestion-resistant hard parts visually identifiable as macrobenthic in origin. None of the 203 crabs examined from the 24-h collection contained meiofaunal prey. Crangon shrimp probably eat meiofaunal prey for only a brief period of time after their initial settlement to the bottom. Evidence for significant top-down impacts on meiofauna from these two abundant shallow-water predators was weak. More trophic studies are needed on newly settled epibenthic predators to test the hypothesis that biological control of shallow-water meiofauna is important.  相似文献   

14.
The intertidal gastropod, Tegula funebralis (A. Adams) exhibits a shore-level size gradient with mean shell size increasing in a down-shore direction. Snails transferred to zones where they do not usually occur migrated back towards their original zone, thus re-establishing a size gradient and implying differential movement among size classes. Both large (≥2.1 cm shell width) and small (≤ 1.77 cm) snails were photonegative on a horizontal surface and geonegative in the laboratory; there were no statistical differences between size classes. Light, however, inhibited upward, or caused downward, movement of large snails on vertical surfaces. Small snails were unaffected, ranging higher on illuminated vertical surfaces than large snails. Both sizes exhibited similar distributions in the dark. In an experimental chamber providing both emersed and immersed surfaces, T. funebralis established vertical size gradients when the chamber was illuminated from above. It is suggested that light is an important factor in the formation and maintenance of Tegula's shore-level size gradient.In response to water-borne chemicals derived from the sea star Pisaster ochraceus (Brandt), large snails moved up vertical surfaces in greater proportion than small. In response to contact with the predator, large snails moved away faster than small and individuals collected from crevices in the field moved away slower than those collected from open rock faces. Although predation may select for a size gradient in Tegulafunebralis, it is unlikely that responses to predatory sea stars directly and proximally cause or maintain them over the short term.  相似文献   

15.
The feeding rate and behaviour of whelks (Buccinum undatum)offered cockles (Cerastoderma edule) in laboratory experimentswere examined. When presented with cockles in a range of sizes(10–40 mm), 14 B. undatum (34.6–88.3 mm),held individually in aquaria, consumed a wide size range ofcockles. Small whelks (<40 mm) consumed cockles (<23 mm),whereas large whelks, (>60 mm) ate a greater numberof larger cockles (>30 mm) and a wider size range ofcockles (12–40 mm) than smaller whelks. The majority(90%) of the shells of the predated cockles were undamaged andthe few (<10%) that were damaged showed only slight abrasionsto the anterior and posterior shell margin. Filmed observationsof B. undatum feeding on C. edule showed a method of attackthat has not previously been reported and involved the use ofthe whelk's foot to asphyxiate the cockle or to pull the shellvalves apart. No filmed evidence was found for the previouslyreported shell ‘wedging’ technique for prising openthe closed shell valves of C. edule, although 10% of the shellsof consumed cockles in feeding experiments had damaged shellmargins. (Received 4 April 2007; accepted 30 June 2007)  相似文献   

16.
Chiu J.-K., Ong S.-J., Yu J.-C., Kao C.-Y. and Iuima T. 1981. Susceptibility of Oncomelania hupensis formosana recombinants and hybrids with Oncomelania hupensis nosophora to infection with Schistosoma japonicum. International Journal for Parasitology11: 391–397. Three generations of Oncomelania hupensis formosana recombinants were produced by mating the Kaohsiung race with the Ilan and Changhua races of the snails. F2 and F3 recombinants were produced by back-crossing F1 and F2 with Kaohsiung O. h. formosana. Subsequently, susceptibility of recombinants to infection with the original strain of Schistosoma japonicum, Ilan or Changhua strain, was investigated. Results indicated that susceptibility of recombinants declined steadily generation by generation. Marked decline of infectivity was observed for Kaohsiung-Ilan recombinants as compared with Kaohsiung-Changhua recombinants. For example, the overall infection rate of Kaohsiung-Ilan F1 recombinants was 7.3 % with a 51.4 % of control snails. The same figures for F2 were 4.2 and 52.6%, and 1 and 40.3 % for F3. On the other hand, the overall infection rate of Kaohsiung-Changhua F1 recombinants was 21.9% with a 46.9% of control snails; and 11.9 and 50.3 % for F2; and 7.6 and 33.2% for F3. The F3 hybrids of Oncomelania hupensis nosophora from Japan and O. h. formosana from Kaohsiung and Ilan were also produced, and susceptibility with the Japanese strain of S. japonicum was studied. A highly significant decline of susceptibility was observed among hybrids (4.4%) in contrast with control snails (85.6 %).Feasibility of applying O. h. formosana in biological control of S. japonicum was discussed. One must determine in the laboratory, prior to application, which race of O. h. formosana should be used depending on the Oncomelania snails of an endemic area. For S. japonicum prevalent in Yamanashi, Japan, the Ilan race of O. h. formosana was found to be better choice than the Kaohsiung race of the snails.  相似文献   

17.
In shallow coastal habitats scavenging netted whelks Nassarius reticulatus attached egg capsules to the stipes of red algae Chondrus crispus and occasionally on Furcellaria lumbricalis and Plumaria plumose. In the laboratory egg capsules were laid on aquaria sides and lids by individuals ≥ 21 mm shell length. Larger size classes produced more egg capsules and spawned over a longer period and in doing so partitioned less energy into shell growth. Large netted whelks (25-28.9 mm) produced larger capsules which contained significantly more and larger eggs than those produced by smaller individuals (21-24.9 mm). Egg capsule production continued throughout the year by regularly fed N. reticulatus held at ambient seawater temperatures. Egg production increased in the spring and summer with peak production during June (15 °C), decreased between August and October and resumed again during the winter (November to February at ∼ 7 °C). During the summer (15-16 °C) egg capsules were smaller and contained smaller eggs than those deposited during the winter (7-10 °C), although the number of eggs · capsule1 was similar. Enforced food limitation reduced the number and size of the egg capsules, the number and size of eggs produced · female1 and the duration of the breeding period. Hatching success of N. reticulatus egg capsules was high (95%) even at winter seawater temperatures (11-8.5 °C) and the duration of embryonic development was fastest between 15 and 17.5 °C.  相似文献   

18.
The purpose of this study was to determine a practical method in Wapiti (Cervus elaphus) of using predetermined sexed Sika (Cervus nippon) semen. Semen was collected by electro-ejaculation from one stag of proven fertility and transported to the laboratory where it was retained as unsorted (control) or was separated into X- and Y-chromosome-bearing sperm using a modified high-speed cell sorter. Wapiti hinds (n = 81) were inseminated into the uterus by rectum manipulation with 1 × 106 (X1 and Y1 group, respectively) or 2 × 106 (X2 and Y2 group, respectively) of sorted frozen-thawed and 1 × 107 non-sorted frozen-thawed (a commercial dose control) Sika motile sperm 60–66 h after removal of intra-vaginal progesterone-impregnated CIDR devices and administration of 700 IU of PMSG at the time of CIDR removal. The percentage of hinds calving after insemination was similar for X1 (38.5%), X2 (41.7%), Y1 (44.4%), Y2 (38.9%) groups (P > 0.05), but higher for control (75%) treatment (P < 0.05). Ultimately 15 out of the 16 Sika and Wapiti-hybrid calves produced by Wapiti hinds inseminated with Y-sorted sperm were male (93.7%) and 10/10 (100%) Sika and Wapiti-hybrid calves from hinds inseminated with X-sorted sperm were female. The sex ratio of the Sika and Wapiti-hybrid calves born to hinds inseminated with sex-sorted sperm deviated significantly (P < 0.05) from 50% and 50.0% in the control group. All Sika and Wapiti-hybrid calves were born between 237 and 250 d of gestation. Male and female calves in the control group had similar birth weights and weaning weights as calves from hinds inseminated with X- or Y-sorted sperm. In conclusion it can be said that normal Sika and Wapiti-hybrid calves of predicted sex can be produced after artificial insemination of Wapiti does with low numbers of sex-sorted cryopreserved Sika sperm.  相似文献   

19.
The symbiotic lifestyle is widespread among porcellanid crabs, which maintain ecological and co-evolutionary associations with annelid polychaetes, poriferans, cnidarians, echinoderms, gastropod mollusks, and other crustaceans such as shrimps and hermit crabs, among others. We investigated the ecological association between the hermit crab Dardanus insignis and the porcellanid Porcellana sayana, in southeastern Brazil. Porcellanid crabs, hermit crabs, and available shells were collected monthly from July 2001 to June 2003, with a shrimp boat equipped with two double-rig trawl nets. The majority of P. sayana specimens were collected in shells occupied by D. insignis (96.6%); a few were found in empty shells (3.4%). The catch of both symbionts and hosts increased with increasing depth, with the highest occurrence at 35 m. The P. sayana crabs of various sizes could be found solitary or forming aggregations of up to 14 individuals per host, showing no sex or size segregation. In spite of the high diversity of shell species occupied by the hermit crabs and also available in the field, only a few of them were also utilized by P. sayana. The majority (93%) of shells utilized by P. sayana also hosted other symbiont species, constituting the basis of extensive symbiotic complexes. Thus, the ecological relationship between D. insignis and P. sayana may be classified as a non-obligate and non-specific symbiosis that may also involve other facultative organisms such as sea anemones.  相似文献   

20.
Early life history patterns were studied in the dominant euphausiids from the northern Gulf of Alaska (GOA) in 2001-2004. Gravid females of Thysanoessa inermis were observed in April and May. Brood size varied from 10 to 1021 eggs with an average of 138 ± 19 (95% CI) eggs female− 1. Most gravid females started to release eggs within the first 2 days of incubation. The average number of eggs released per female was similar in incubation Day 1 and 2, but significantly smaller on Day 3 and 4. About 25% of the females were continuously releasing eggs over 3 days rather than producing a single distinctive brood. In contrast, gravid females of Euphausia pacifica were observed from early July through October. Most gravid females released eggs on the first day of observation, while only 2% of females produced eggs repeatedly. Brood size varied from 20 to 246 eggs with an average of 102 ± 12 (95% CI) eggs female− 1. The relationship between E. pacifica brood size and ambient chlorophyll-a concentration was sigmoidal (r2 = 0.73), with food saturated brood size of 144 ± 14(SE, P < 0.001) eggs, and half-saturation occurring at 0.46 ± 0.02(SE, P < 0.001) mg chlorophyll-a m− 3. The average interbrood interval of E. pacifica reared at 12 °C and satiated food conditions in the laboratory was ∼ 8 days, suggesting their potential individual fecundity in the GOA was 1148-1530 eggs per spawning season. Hatching and early development (from egg to furcilia stage) was studied under 5 °C, 8 °C and 12 °C. Hatching was nearly synchronous and lasted 3-6 h, depending on incubation temperature. Development times from egg to the first furcilia stage ranged between 20 and 33 days for T. inermis, and 15 and 45 days for E. pacifica at 12 °C and 5 °C, respectively.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号