the ecology of octopus aff. o. tetricus and prevalence of

67
1 The ecology of Octopus aff. O. tetricus and prevalence of anisakid nematodes in near-coastal waters of North Fremantle, Western Australia Sleeping Octopus aff. O. tetricus in Coogee, Western Australia. Image by J. Claybrook Submitted by Jorja Claybrook BSc (Marine Science) This Thesis is presented for the degree of Bachelor of Science Honours in Marine Science College of science, health, engineering and education, Murdoch University, 2020.

Upload: others

Post on 29-May-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The ecology of Octopus aff. O. tetricus and prevalence of

1

The ecology of Octopus aff. O. tetricus and prevalence of

anisakid nematodes in near-coastal waters of North Fremantle,

Western Australia

Sleeping Octopus aff. O. tetricus in Coogee, Western Australia. Image by J. Claybrook

Submitted by

Jorja Claybrook BSc (Marine Science)

This Thesis is presented for the degree of Bachelor of Science Honours in Marine Science

College of science, health, engineering and education, Murdoch University, 2020.

Page 2: The ecology of Octopus aff. O. tetricus and prevalence of

2

Declaration

I declare that this Thesis is my own account of my research and contains its main content

work, which has not been previously submitted for a degree at any tertiary education institute

Jorja Lee Claybrook

Page 3: The ecology of Octopus aff. O. tetricus and prevalence of

3

Abstract

In contrast to most fished groups, global landings for cephalopod have been increasing

over the last two decades. This and their fast-growth rate and short lifecycles, responsiveness

to environmental change, and their function as mesopredators within the marine environment

have led to an increased focus on their ecology. Within Western Australia, the common

octopus, known currently as Octopus aff. O. tetricus is the target species of the commercial

fishery which operates in Cockburn Sound. This Thesis focused on the population of O. tetricus

inhabiting near-coastal waters of North Fremantle, ~7 km north of the current commercial

operations and assessed any inter and intra-population variability between the two sites.

Samples of octopus were collected by commercial fishers setting four lines of shelter pots, two

in shallow (≤10 m) and two in deeper water (15-17 m) on seven occasions between March and

June 2020 in North Fremantle. One sample of octopus was collected from Cockburn Sound on

1st April. These data were used to examine if the distribution, abundance, and sex ratio differed

across the sampling period and between depths.

A total of 701 octopus were caught ranging in size from 35 to 218 mm mantle length

(ML) and 5.13 to 691.62 g mantle weight. Over the duration of the study, the mean number of

octopuses caught decreased. Males were more abundant in March and May, while females were

more prevalent in April. The mean size of female octopus fluctuated in the first four hauls the

trap lines and then declined from late-April to June, indicating a possible migration out of the

area by larger females. Females caught in Cockburn Sound over a wider range of habitats were

larger than those from North Fremantle at the same time. The mean size of male octopus

increased gradually over the duration of the study with the highest mean ML recorded in June.

Male octopus matured during the study with the number of immature males decreasing from

March to June, in both depth regions.

Initially, this Thesis aimed to include a detailed analysis of the dietary composition of

Octopus aff. O. tetricus, however due to time constraints only a preliminary analysis is

presented in this dissertation. The contents of the gastric tract (crop and stomach) of 701

octopus were examined. A total of 12 taxa were identified in the gastric contents, with

crustaceans identified as the most common prey, followed by teleosts and non-cephalopod

molluscs. Cannibalism was observed in 51 octopus (10.8% of gastric contents with dietary

items).

Page 4: The ecology of Octopus aff. O. tetricus and prevalence of

4

During the dissection of octopus mantles, parasites, tentatively identified as anisakid

nematodes, were recorded on the internal organs and this became an additional focus of the

research. The incidence of parasites was nearly 60% in North Fremantle and when present,

always occurred in the gastric tract but were also found in the mantle muscles, digestive gland,

posterior salivary glands and gonads. Binomial logistic models with a logit link function

showed that the probability of parasite incidence was nearly ten times higher in the North

Fremantle population compared with Cockburn Sound. The probability of parasite occurrence

was also significantly greater on medium (80-160 mm ML) and large octopus (>160 mm ML)

than small octopus.

The results from this study show that the nearshore waters of North Fremantle are likely

a nursery habitat for juvenile and maturing male octopus’ but females probably migrate to

deeper waters to mate and brood their eggs. The prevalence of parasites and location suggests

that nematodes are being ingested within North Fremantle at a much higher rate than in

Cockburn Sound.

Page 5: The ecology of Octopus aff. O. tetricus and prevalence of

5

Table of Contents Declaration .............................................................................................................................................. 2

Abstract ................................................................................................................................................... 3

Acknowledgments ................................................................................................................................... 7

1. Introduction ..................................................................................................................................... 9

1.1 Cephalopod Fisheries .............................................................................................................. 9

1.2.1 Australian Cephalopod Fisheries ......................................................................................... 11

1.3 Ecology of coastal cephalopods ............................................................................................ 11

1.3.1 Adaptations to the coastal environment ........................................................................ 12

1.3.2 Distribution and Abundance ......................................................................................... 14

1.3.3 Reproduction and Brooding ................................................................................................. 17

1.4 Octopus aff. O. tetricus ......................................................................................................... 19

1.4.1 Western Australian Octopus Fishery ............................................................................ 19

1.5 Rationale and Study Aims ..................................................................................................... 21

2. Methods .................................................................................................................................... 21

2.1 Study site ............................................................................................................................... 21

2.2 Collection of Octopus aff. O. tetricus ................................................................................... 22

2.3 Biological processing of Octopus aff. O. tetricus ................................................................. 24

2.3.1 Maturation ..................................................................................................................... 24

2.3.4 Parasites ........................................................................................................................ 25

2.4 Data analyses......................................................................................................................... 25

2.4.1 Environmental parameters ............................................................................................ 25

2.4.2 Distribution and abundance of octopus ......................................................................... 26

2.4.3 Octopus size structure ................................................................................................... 26

2.4.4 Mantle length distributions ........................................................................................... 26

2.4.5 Length-weight relationships .......................................................................................... 26

2.4.6 Gonadosomatic Index ................................................................................................... 27

2.4.7 Parasites ........................................................................................................................ 27

2.5 Diet Analysis ......................................................................................................................... 28

3. Results ....................................................................................................................................... 28

3.1 Water temperature ................................................................................................................. 28

3.2 Changes in numbers with depth and month .......................................................................... 29

3.3 Sex ratio ............................................................................................................................... 32

3.4 Size distribution .................................................................................................................... 33

3.4.1 Overall size distribution ................................................................................................ 33

3.4.2 Mantle Length-weight relationship .............................................................................. 34

Page 6: The ecology of Octopus aff. O. tetricus and prevalence of

6

3.4.3 Size distribution in deep and shallow water .................................................................. 35

3.4.4 Temporal variation in size ............................................................................................. 37

3.4.5 Cockburn Sound size distribution ................................................................................. 41

3.5 Reproduction and Maturity ................................................................................................... 43

3.5.1. Gonad maturity stages .................................................................................................. 43

3.5.2 Gonadosomatic Indices ................................................................................................. 46

3.6 Preliminary diet results ............................................................................................................... 46

3.7 Parasites ...................................................................................................................................... 47

3.7.1 Parasite presence ........................................................................................................... 47

3.7.2 Parasite presence modelling .......................................................................................... 49

4. Discussion ......................................................................................................................................... 50

4.1 Abundance and sex ratio ............................................................................................................. 51

4.2 Size distribution .......................................................................................................................... 52

4.4 Occurrence of nematodes ............................................................................................................ 56

Conclusions ........................................................................................................................................... 59

References ............................................................................................................................................. 61

Page 7: The ecology of Octopus aff. O. tetricus and prevalence of

7

Acknowledgements

First and foremost, I would like to acknowledge that 2020 has been a tough year and I

do not think there is not a single person who has not been affected by the Covid-19 pandemic

in some way. I know we have all heard the word ‘unprecedented’ more in the past few months

than we probably have in our lifetimes but there is truly no better fitting word. This year has

shown just how resilient we are, and so many people have gone above and beyond to keep us

students going, with that I would like to extend my deepest gratitude for every person who has

helped me complete this thesis during such a challenging time.

To Peter Jasinski and your crew, without your generous support, this project would

have not been possible. Your sense of humour and jovial attitude was always a pleasure to be

around. Without your help, I believe this project would have never gotten off the ground, I am

especially appreciative that you continued to help me even when Covid-19 restrictions stopped

me from participating in the sampling. To Phil, and everyone at Fins Seafood, thank you for

allowing me to use your facilities to freeze and process all my samples. It is no exaggeration

to say that your generosity saved me so much stress during these challenging times, and

everyone was so kind and welcoming every day I was there.

To my supervisors, Neil Loneragan and Anthony Hart, thank you deeply for your

ongoing encouragement and support. Anthony, thank you so much for bringing this project to

life, all your hard work behind the scenes and expertise has been invaluable. I thank you so

much for sharing your knowledge of the octopus’ ecology with me. Neil, where do I even

begin? It seems like so long ago I came to you wanting to do an honours project on octopus,

and despite what felt like a million setbacks, here we are today with a complete thesis! I have

nothing but appreciation and admiration for you Neil. Your unwavering support and kindness

have kept me going even through some of the trickiest times, and without you reminding me

to take a break every now and then I think I would have burnt myself out long ago. I know that

I am so lucky to have had you in my corner throughout this journey, and there are no words to

truly describe how much you have helped this project come to fruition. I thank you from the

bottom of my heart for everything.

Thank you to Claire Greenwell, without you I do not even know if I would have pursued

an Honours at all. From the very get-go, you encouraged me to push for a project I wanted, and

I am truly fortunate to have you as not only a mentor but also a friend. You are incredibly

Page 8: The ecology of Octopus aff. O. tetricus and prevalence of

8

inspirational, with your passion and creativity, if I can become half the scientist you are, I will

consider myself very lucky.

To Charlie Maus, Kurt Krispyn, Thea Linke, James Tweedley and all of the beautiful

people from the fish and fisheries research group, I thank you for your friendship, all of the

laughs and especially all of the cakes. Thank you, Dave Murphy, for all your advice and for

keeping me on track. Thanks to Mikki and Shelly Smith, and especially your Nanna who leant

me her chest freezer and saved me in my moment of need. I would like to especially

acknowledge the beautiful staff members working in the Science Store, on more than one

occasion I had difficulties with accessing my account and I would not have been able to resolve

it without your help and patience.

To all my non-sciencey friends, your support and love have never gone unnoticed.

Thank you for not forgetting about me, putting up with me going MIA and being a much-

needed respite when I am feeling overwhelmed.

A massive thank you to my parents Bob and Jill, who have supported me through each

and every one of my crazy plans throughout the years. There are no words to describe just how

lucky I am to have you. To all of my family and Jake’s family, I love and cherish you all. A

special shoutout to Steve Cossington, as I do not think I would be a marine scientist without

having you as an inspiration.

And to Jake, safe to say that neither of us could have predicted that we would be here

together, all those years ago. Honestly, I do not think I can explain just how much you have

done for me not just throughout this Honours but also through all my years at university. Every

hurdle this year threw at us, your positive, can-do attitude kept me going. Thank you for putting

up with the 200 hundred octopuses in the freezer in our house and all my general clutter when

I had to try and make a home laboratory. For all your efforts, including learning how to dissect

octopus with me, your uncanny knack of being able to remove stylets and just generally keeping

me sane over the past year. Thank you for all the late nights you stayed up with me, for bringing

me dinners and even not complaining all the times you were covered in octopus’ ink and goop.

You did this all while you completed the final year of your degree, you truly are an incredible

person.

Finally, I would like to acknowledge the support afforded by the Myrtle AB Lamb

scholarship and the scholarship provided by the Harry Butler Institute.

Page 9: The ecology of Octopus aff. O. tetricus and prevalence of

9

1. Introduction

Interest in cephalopod ecology has increased exponentially over the past few decades,

with a spotlight being focused on their importance in not only within the marine ecosystem but

also as valuable food resource, especially with growing concerns of the sustainability of

traditionally fished species (O'Brien et al., 2018). Coeloid cephalopods (i.e. those lacking an

external shell) are considered keystone species within many marine ecosystems. In many food

webs, cephalopods are mesopredators which establish the link between apex predators and

lower trophic levels (Boyle & Rodhouse, 2005; Heery et al., 2018; Yick et al., 2012). Being

mesopredators, cephalopods have the potential to influence the structure and dynamics of their

ecosystem (Yick et al., 2012). Due to their sensitivity to environmental change, cephalopods

can act as ecological indicators and exhibit top-down control over their prey (Boyle &

Rodhouse, 2005; Rodhouse et al., 1997). Furthermore, their short-lived, fast-growing life

history means they can rapidly respond to fluctuations within their environment, which can

cause detrimental consequences for upper and lower trophic levels within their ecosystem

(Tian, 2009).

In recent years, considerable focus has been directed to the increasing economic

importance of cephalopods as traditionally targeted finfish species continue to decline due to

anthropogenic stresses such overfishing and climate change (Caddy & Rodhouse, 1998; Rosa

et al., 2019; Watson & Pauly, 2001). This thesis focuses on examining the distribution and

abundance of the target species of the Western Australian Octopus fishery, Octopus aff. O.

tetricus, in near-coastal waters of Western Australia. It aims to build upon current knowledge

of cephalopods and their importance in commercial and ecological settings.

1.1.1 Cephalopod Fisheries

Scientific advances throughout the past 50 years have led to a massive increase in the

comprehension of marine ecosystem functions (FAO, 2020). Commercial fish production from

capture fisheries and aquaculture are estimated to have reached 179 million tonnes in 2018

worldwide, with an estimated value of 401 billion USD (FAO, 2020). Consumption of seafood

has increased by 3.1% annually between 1961 to 2017, which is almost twice the rate of

population increase (1.6%), and higher than other protein foods (2.1% annual increase) (FAO,

2020). Fisheries can, directly and indirectly, affect marine ecosystem predator-prey

relationships, which if overexploited may lead to the shift in trophic relationships and

Page 10: The ecology of Octopus aff. O. tetricus and prevalence of

10

potentially loss biodiversity or collapse of fish stocks (Coleman & Williams, 2002; Thrush et

al., 2016).

Cephalopods have been an increasingly important resource in fisheries worldwide and

contribute 2‒5% of global landings from capture fisheries. Unlike finfish fisheries, which have

plateaued or declined, cephalopod fisheries have gradually increased from 0.5 million tonnes

in 1950 to a peak of 4.85 million tonnes in 2014, this number decreased to 3.5 million in 2018

but has continued to stay at relatively high levels (FAO, 2020).

Worldwide, octopus’ are fished using passive shelter pots, which are low cost, easy to

operate and cause minimal impact on the surrounding environment. However, the pots often

cannot be deployed in deep waters or areas of high energy (Hart et al., 2016; Leporati et al.,

2015; Sauer et al., 2019). Alternative trapping methods used by commercial octopus fisheries

include the frame traps, which utilise a netting that allows the octopus to push in but not back

out (Hernández-Garcı́a et al., 1998). More recently, the use of octopus traps with an active

closing mechanism, triggered by a tripwire, and bait have been trialled in commercial fisheries

(Carreira & Gonçalves, 2009; Sauer et al., 2019). The utilisation of active traps is prohibited

in many octopus' fisheries worldwide. However, the Department of Primary Industries and

Resource Development in Western Australia has recently conducted a trigger trap trial and

allows recreational fisherman to use up to six active traps (Hart et al., 2019; Leporati et al.,

2015; Sauer et al., 2019).

The management and assessment of commercial cephalopod fisheries are challenging

(FAO, 2020) because of the short life-cycle, fast growth rates and the variability in recruitment

and catch rates of cephalopods with environmental fluctuations (Boyle & Boletzky, 1996).

These challenges are highlighted by recent declines in three of the main commercial squid

species; jumbo flying squid (Dosidicus gigas), Argentine shortfin squid (Illex argentinus) and

Japanese flying squid (Todarodes pacificus) (FAO, 2020). In recent years, there has been a

decline in commercial octopus’ suppliers due to an increasing scarcity of octopus, as some

important fisheries have become less productive, requiring urgent and strict management

strategies (FAO, 2020).

Page 11: The ecology of Octopus aff. O. tetricus and prevalence of

11

1.2.1 Australian Cephalopod Fisheries

Commercial cephalopod fisheries are a relatively new development in Australia.

Initially, international chartered vessels discovered significant resources of arrow squid

Nototodarus goudi off the southern coast of Australia in the 1970s (Kailola et al., 1993).

However, it was not until the mid-1980s that cephalopod stocks were exploited commercially

(Kailola et al., 1993; Nottage et al., 2007). Most cephalopod landings in Australia are taken as

a by-product from other fisheries, particularly lobster fisheries. Australian cephalopod

commercial landings primarily consist of squid, followed by octopus. Cuttlefish are caught

only by-catch of other fishing efforts (Kailola et al., 1993; Nottage et al., 2007).

The southern squid jig fishery (SSJF) operates between Fraser Island in Queensland to

the South Australian and Western Australian border and includes waters around Tasmania

(Nowara & Walker, 1998; Trinder & McKinley, 2003). The target species of the fishery is

arrow squid (Nototodarus gouldi) and has incidences of Sepioteuthis australis (southern

calamari), Todarodes filippovae (Southern Ocean arrow squid) and Omnastrephes bartrami

(red ocean squid) as bycatch. Incidental catches of N. gouldi are also caught in the Southern

and Eastern Scalefish and Shark Fishery. SSJF uses up to 10 automatic jig machines on each

vessel, each with 2 spools of heavy line, with ~ 25 jigs attached to each line (Furlani et al.,

2007).

At present, Australia has two dedicated octopus’ fisheries in the Eastern Indian Ocean,

the Octopus aff. O. tetricus fishery in Western Australia (for further detail see 1.4.1 below) and

the O. pallidus fishery in Tasmania (Sauer et al., 2019). The Tasmanian octopus’ fishery was

established in 1980, is operated by a single company with two full-time vessels which fish in

the Bass Strait (Leporati et al., 2009). The fishery uses passive shelter pots on demersal

longlines ~4 km in length, each line contains up to 1000 pots per line and the lines are deployed

in depths from18–85 m (Leporati et al., 2009).

1.3 Ecology of coastal cephalopods

Coleoid cephalopods (i.e. those without an external shell) are present in practically all

types of coastal habitats worldwide (Boyle & Rodhouse, 2005). They are not known to

inhabit marine environments with low salinity, and no cephalopods are endemic to estuarine

environments. Coastal species exhibit demersal lifestyles (i.e. strong association to the

seafloor) (Boyle & Rodhouse, 2005).

Page 12: The ecology of Octopus aff. O. tetricus and prevalence of

12

1.3.1 Adaptations to the coastal environment

Demersal cephalopods utilise sophisticated organs to solve problems they encounter

within the natural environment (Hanlon & Messenger, 2018). Cephalopods have practical

locomotive adaptations (i.e. strong mantle, funnel, and fin muscles) and highly evolved sensory

organs. The sensory organs of cephalopods are the most evolved of all invertebrates and rival

the sophistication of vertebrate systems (Boyle & Rodhouse, 2005). Coleoid cephalopods

possess a ‘camera-like’ eye similar to vertebrates, containing a pupil, lens and rhabdomeric

photoreceptors which form the retina (Serb & Eernisse, 2008). Cephalopods are colour-blind

(Marshall & Messenger, 1996) but are sensitive to polarisation (Moody & Parriss, 1960;

Pignatelli et al., 2011; Shashar et al., 2000).

Most coleoid cephalopods have a complex nervous system, with a multi-lobed brain.

The brain to body ratio of cephalopods is akin to those of lower vertebrates (Mather & Kuba,

2013). The complexity of the nervous systems allows cephalopods to adapt to a wide range of

learning (Mather & Kuba, 2013) including; associative (Cole & Adamo, 2005), spatial (Alves

et al., 2006; Scatà et al., 2016), social (Tricarico et al., 2011) and problem-solving (Richter et

al., 2016). Cephalopods exhibit remarkable behavioural plasticity (Alves et al., 2006; Brown

& Piscopo, 2013; Hanlon & Messenger, 2018).

Coastal species actively use the benthic environment. They are predominantly bottom

feeders, feeding on the rich diversity of invertebrates and small fish available (Boyle &

Rodhouse, 2005). Octopuses have adapted sophisticated hunting and prey-handling (Boyle &

Rodhouse, 2005; Hanlon & Messenger, 2018). Demersal cephalopod eggs are encapsulated in

tough protective sheaths and fixed to the hard substrate, kelp plants and artificial surfaces such

as buoys or fishing traps. Moreover, octopuses are highly dependent on crevices, rock, and

reefs within the coastal environment for protection from predation (Anderson, 1997; Boyle &

Rodhouse, 2005; Mather, 1994; Mather & O'Dor, 1991).

Feeding ecology

Feeding and foraging of coleoid cephalopods have been comprehensively studied and

reviewed (Boyle & Rodhouse, 2005; Villanueva et al., 2017). Coleoid cephalopods have a

voracious, exclusively carnivorous feeding regime, functioning as mesopredators within the

Page 13: The ecology of Octopus aff. O. tetricus and prevalence of

13

marine environment (Boyle & Rodhouse, 2005; Greenwell et al., 2019a; Hanlon & Messenger,

2018). Cephalopods possess a range of physiological, morphological, and behavioural

adaptations that help them be highly successful predators within the marine environment,

which can exhibit significant predatory pressure on benthic communities.

Cephalopods use versatile feeding apparatus (arms and tentacles) to feed on a diverse

group of prey (Boyle & Rodhouse, 2005). All cephalopods exclusively feed by dismembering

prey with their beak, consequently, diet analysis can be exceptionally challenging. Demersal

cephalopods primarily feed on crustaceans, fish and other cephalopods (Boyle & Rodhouse,

2005; Greenwell et al., 2019a; Greenwell et al., 2019b; Villanueva et al., 2017)As cephalopods

are opportunistic, their prey selection varies with prey availability which varies with depth and

habitat type (Smith, 2003). Variations in diets of octopus are also related to size and maturation

stage (Smith, 2003).

The main predators of coastal cephalopods are seabirds, marine mammals, and larger

fishes. Animals which predate on fish will also take cephalopods (Boyle & Rodhouse, 2005).

Except for a few elasmobranch species, no animals are known to exclusively predate on

cephalopods (Boyle & Rodhouse, 2005).

Habitat utilisation

Coeloid cephalopods have lost the external shell to increase mobility, and as a

consequence, they are very vulnerable to predation (Cigliano, 1993). Demersal species often

use dens or naturally occurring crevices for protection (Aronson, 1986; Hanlon & Messenger,

2018). The characteristics of the substrate where benthic species reside are thought to be a key

factor in influencing their distribution (Hanlon & Messenger, 2018). Studies investigating

habitat preference of octopus species found den availability and preference for the reef edges

were key factors affecting octopus’ distribution (Anderson, 1997; Mather, 1994).

Den availability has also been described as a limiting factor for octopus’ density

(Hartwick et al., 1984; Katsanevakis & Verriopoulos, 2004; Mather, 1994). Demersal

cephalopods routinely compete for shelters; with size and sex influencing the success of shelter

occupancy (Mather, 1982; Narvarte et al., 2013).

Sandy substrates often have limited shelter available for cephalopods which can impact

the density of cephalopods and the capacity for females to brood their eggs in these areas.

Molluscan shells are often used by an octopus to build suitable dens on sandy substrate.

Page 14: The ecology of Octopus aff. O. tetricus and prevalence of

14

Narvarte et al. (2013) found that juvenile Patagonian octopus used barnacle shells and adults

utilised larger oyster and clamshells to create dens.

Camouflage

Crypsis and cryptic behaviour is the primary defence mechanism of coastal

cephalopods (Boyle & Rodhouse, 2005; Hanlon et al., 2010; Marshall & Messenger, 1996).

Cephalopod camouflage abilities exploit surrounding environmental features and allow them

to ‘blend in’ to their surroundings (Boyle & Rodhouse, 2005). Hanlon and Messenger (2018)

summarised the different camouflage techniques of cephalopods, including background

resemblance, countershading, disruptive colouration, and deceptive resemblance (Doubleday

et al., 2016; Villanueva et al., 2016).

1.3.2 Distribution and abundance

Cephalopod distribution varies considerably depending on species (Doubleday et al.,

2016). Demersal cephalopods exhibit the lowest dispersal capacity in shelf waters (within 10's

km). Pelagic cephalopods inhabit open oceanic waters and have the highest capacity for

dispersal (thousands of km), due to having a paralarvae stage (duration ~5-115 d) and a highly

motile adult stage

Octopus and cuttlefish tend to be concentrated in demersal habitats in coastal

environments (Boyle & Rodhouse, 2005; Quetglas et al., 2000), while the squids from the order

Myopsida are restricted to the neritic zone, likely due to their dependence on the benthos for

egg-laying and feeding opportunities (Pierce et al., 2008). Within shallow coastal habitats,

benthic octopuses dominate rocky reef habitats whereas, cuttlefish and squid species are more

commonly found in soft substrate habitats. True squids are predominantly pelagic, and their

distribution and depth of occurrence vary widely among species. Most studies examining

distribution and abundance of coastal cephalopods focus on north Atlantic and Mediterranean

populations (Arechavala-Lopez et al., 2019; Belcari et al., 2002; Quetglas et al., 2000; Vargas-

Yáñez et al., 2009) with far less published data on those in Southern Hemisphere (Ramos et

al., 2014).

Impact of urbanisation

Understanding the impact of urbanisation on cephalopod distribution and abundance is

still in its infancy. As marine urbanisation changes natural habitats, this is likely to alter

Page 15: The ecology of Octopus aff. O. tetricus and prevalence of

15

cephalopod interactions with these environments. Artificial structures are known to support

distinct assemblages of marine fauna, especially mobile species such as octopus (Mather,

1994). At present, there is one published study investigating the impact of urbanisation on

distribution and abundance of octopus. Heery et al. (2018) observed urbanisation-related

patterns of Entereoctopus dofleini. Unlike terrestrial animals in urban environments which

benefit directly or indirectly from enhanced food resources from humans, there is no evidence

to suggest that octopus’ diets are subsidised by urbanisation (Heery et al., 2018). However,

octopuses do frequently utilise structures as dens and were more likely to occur in areas with

large amounts of anthropogenic debris (Heery et al., 2018). The use of artificial structures for

dens has been established in several studies (Katsanevakis & Verriopoulos, 2004; Mather,

1994). Urbanisation is likely to affect octopus abundance more in populations which occur on

sandy environments, where dens are scarce (Katsanevakis & Verriopoulos, 2004).

Historically, octopuses were identified as habitually solitary animals, only interacting

with one another in mating periods, with other encounters being hostile. Recently, a few studies

have identified more social behaviours (i.e. aggregation and interaction) in certain species

including O. tetricus (Scheel et al., 2017) and the undescribed Larger Pacific Striped Octopus

(Caldwell et al., 2015). Scheel et al. (2017) identified O. tetricus residing in high densities in

two locations in Jervis Bay, on the east coast of Australia. Typically, O. tetricus has been

identified as a solitary species but within these locations exhibit complex social behaviours. It

has been hypothesised that these large aggregations form in response to limited den availability

but an abundant food supply (Scheel et al., 2017). It has been suggested that clustered shelters

heighten social interactions of octopus (Scheel et al., 2017).

Environmental influences on distribution and abundance

Environmental conditions and the strength of recruitment have a significant impact on

abundance (Sobrino et al., 2002). The size of cephalopod populations is variable at annual

timescales with rapid expansion occurring when conditions are favourable and rapid decline

when conditions become unfavourable (Rodhouse, 2010). In recent years there has been

growing speculation that cephalopod populations may be proliferating in response to changing

conditions, as evidenced by the increasing trends in cephalopod catch rates (Doubleday et al.,

2016). It is well established that cephalopods are highly sensitive to shifts in environmental

conditions. Cephalopods respond to changes in ecological conditions either actively (i.e.

moving with changing conditions to areas more favourable for spawning or feeding) or

Page 16: The ecology of Octopus aff. O. tetricus and prevalence of

16

passively (i.e. effects on growth and paralarvae survival) (Pierce et al., 2008). Due to their

sensitivity to changing conditions, cephalopods could potentially act as environmental

indicators (Pierce et al., 2008).

Octopus distribution shows a significant relationship between size and depth, with an

ontogenetic shift to deeper waters with increasing size (Leite et al., 2009). The ontogenetic

distribution also appears to be linked with environmental conditions such as temperature.

Smaller individuals of O. insularis are known to occupy warmer waters than larger individuals,

likely due to the high productivity attributed to warm waters, which promotes faster growth

and therefore, shortens the period in which they are most vulnerable to predation (Leite et al.,

2009). Migration to brooding habitats has been documented for several species of octopus

including Octopus. aff. O. tetricus (Leporati et al., 2015), O. magnificus (Smith et al., 2006),

O. tetricus (Anderson, 1997), O. insularis (Leite et al., 2009; Lima et al., 2014) and, O. vulgaris

(Belcari et al., 2002; Katsanevakis & Verriopoulos, 2004).

Many coastal octopus species exhibit interannual variability and cyclical patterns of

abundance (Regueira et al., 2014). The relationship of the abundance of O. vulgaris and the

variability of SST were investigated in the Alboran Sea (Vargas-Yáñez et al., 2009). Results

showed that O. vulgaris abundance was adversely affected by warm water anomalies occurring

in the previous year (Vargas-Yáñez et al., 2009). These findings indicated that O. vulgaris

landings could be negatively affected by long-term warming of the Western Mediterranean.

Similarly, Sobrino et al. (2002) identified a beneficial influence from low winter temperature

for the next years landings. They found that cephalopod landings in the Gulf of Cadiz reached

a maximum in winter before decreasing to their minimum in September. Other studies in the

Mediterranean Sea found similar results, with minimum abundance occurring between August

and September (Vargas-Yáñez et al., 2009). The influence of temperature is intrinsically linked

to other environmental and biological variables (i.e. upwelling intensity, enhanced primary

production, embryonic and paralarvae growth) (Sobrino et al., 2002; Vargas-Yáñez et al.,

2009).

Abundance patterns of tropical cephalopod species do not exhibit the same seasonal

variations as those seen in temperate and sub-tropical areas (i.e. O. vulgaris, Katsanevakis and

Verriopoulos, 2004; Entereoctopus dofleini, Hartwick et al., 1984). The abundance of tropical

species appears to be more influenced by habitat characteristics such as depth and substrate

than temperature (Leite et al., 2009).

Page 17: The ecology of Octopus aff. O. tetricus and prevalence of

17

There are growing concerns about the impact of long-term warming and climate change

on cephalopod populations. Warming temperatures are known to increase growth rates which

subsequently accelerates their life cycles (Doubleday et al., 2016) and could cause a faster

population turnover. Warmer temperatures are also known to cause premature hatchings,

leading to decreased chances of survival during the crucial first few days (Uriarte et al., 2016).

The rapid life cycle of cephalopods will likely aid in their adaptations to changing oceanic

conditions, and some cephalopod species may benefit from changing conditions (Doubleday et

al., 2016).

1.3.3 Reproduction and Brooding

Nearly all cephalopods are semelparous (i.e. one reproductive episode) (Boyle &

Boletzky, 1996; Boyle & Rodhouse, 2005). Unlike other molluscs, the sexes are separate, and

there is no evidence of hermaphroditism or sex change within cephalopods.

Assessment of maturity stages is an important aspect of both ecological studies and

fishery management. The temporal and spatial distributions of breeding aggregations, the size

at maturity and the sex ratio all require an assessment to ensure population levels are sustained.

The mating strategy of cephalopods is uncommon among other molluscs, reproduction

always takes place through individual mating. Octopus males possess a hectocotylus (i.e.

modified arm) which is used to transfer spermatophores to females during copulation.

Spermatophores can be stored within the mantle cavity, which means that a significant time

may pass between mating and fertilisation. Males are likely to mate with several females.

Unlike other cephalopods that may spawn in batches over an extended period, octopuses tend

to only lay eggs over a few days (Boyle & Rodhouse, 2005)

Locating a suitable egg-laying habitat is important to demersal cephalopod

reproductive success (Pratasik et al., 2017). Habitat selection for other demersal cephalopods,

especially for mature females, is influenced by suitable environments to brood their eggs

(Guerra et al., 2015; Narvarte et al., 2013; Pratasik et al., 2017). The quality of cephalopod

shelters is known to affect reproductive output (Iribarne, 1990; Narvarte et al., 2013).

Page 18: The ecology of Octopus aff. O. tetricus and prevalence of

18

All female octopuses studied are known to protect and clean their eggs throughout the

entire duration of embryonic development (Robin et al., 2014). Female benthic octopuses brood

their eggs in rocky crevices, coralline overhangs, or confined spaces (Boyle & Rodhouse,

2005). Eggs are deposited into the den using two methods: (i) the eggs are individually

encrusted into the hard substrate or; (ii) cemented into the hard substrate in clusters where

individual eggs are intertwined with one another (Hanlon & Messenger, 2018). Some genera

of octopus that occupy habitats without suitable brooding shelters carry their developing eggs

within the webbing of their arms, e.g. Hapaloclaena (Norman et al., 2000) and Amphioctopus

(Guerrero-Kommritz et al., 2016; Guerrero-Kommritz & Rodriguez-Bermudez, 2018).

Narvarte et al. (2013) investigated the effect of egg predation and competition on

brooding shelter selection of O. tehuelchus in San Antonio Bay, Patagonia Argentina. They

found that shelter availability was a constraint on brooding and affected the level of parental

care for the fertilised eggs. Furthermore, Iribarne (1990) found that the fecundity of

O. tehuelchus was higher in large-volume shelters than smaller shelters. If octopuses of both

sexes at a similar size compete for a single shelter, the female will win occupancy of the shelter

(Narvarte et al., 2013). Contest over shelters between sexes has been observed for several

species including O. tehuelchus (Narvarte et al. 2013), O. dofleini (Hartwick et al. 1984) and

O. briareus (Aronson, 1986). Competition for shelters is more apparent during brooding

periods and a female octopus may inhabit a single shelter for more than two months (Narvarte

et al., 2013).

The final life stage for octopuses is a period of senescence (Anderson et al., 2002).

Usually, males undergo senescence after mating, while females senesce while brooding their

eggs. Typically, symptoms of senescence include lack of feeding, retraction of skin around the

eyes, decreased physical abilities and lesions on the skin (Anderson et al., 2002; Boyle &

Rodhouse, 2005; Wodinsky, 1977). Many aspects of octopus senescence are still poorly

understood (Anderson et al., 2002).

After spawning, female octopus' decrease and then stop feeding during the incubation

of the eggs, focusing all energy entirely on egg care until the hatchlings emerge (Boyle &

Rodhouse, 2005; Hanlon & Messenger, 2018; Robin et al., 2014; Roumbedakis et al., 2018b).

Maternal care from female octopuses usually involves protecting eggs from predators,

increasing ventilation by flushing water over the eggs, cleaning eggs and removing any dead

embryos (Roumbedakis et al., 2018b). The physiological condition of female brooding octopus

declines over time. The starvation behaviour of female octopuses during the brooding stage

Page 19: The ecology of Octopus aff. O. tetricus and prevalence of

19

is well established throughout literature. In a study of O. hummelincki, Wodinsky (1977)

suggested that the biological trigger for this behaviour could be linked to secretions of the

optical glands. Experiments involving the removal of the optical gland after spawning

revealed that O. hummelincki females returned to feeding, gained weight and the lifespan was

extended by four months or more (Wodinsky, 1977).

1.4 Octopus aff. O. tetricus

Octopus aff. O. tetricus is a merobenthic (i.e. has a paralarval stage) species endemic

to temperate coastal waters (≤ 70 m deep) of Western Australia. This species is relatively short-

lived (≤ 18 months), medium-sized (< 4 kg) and is highly fecund (125 000 – 150 000) (Joll,

1976). Once considered synonymous with the Common Sydney Octopus, Octopus tetricus,

taxonomic studies suggest that Octopus aff. O. tetricus (not yet formally described) is a

separate species, with a geographically distinct population that extends from Shark Bay to the

South Australian Border (Amor et al., 2014). Octopus aff. O. tetricus and O. tetricus are closely

related to the cosmopolitan species O. vulgaris.

Similar to many other octopus’ species, Octopus aff. O. tetricus (hereafter O. tetricus)

is semelparous. Both sexes of O. tetricus are known to have similar life expectancies, with

males reaching maturity at an earlier age than females (Leporati & Hart, 2015). The maximum

size for females (~4 kg) is considerably larger than males (~2.5 kg). This species inhabits rocky

reefs, sandy substrates and seagrass meadows Recent studies have provided evidence of

mature females migrating to deep waters to find appropriate dens to brood their eggs,

followed by mature males seeking mates (Leporati et al., 2015).

1.4.1 Western Australian Octopus Fishery

The first commercial fishing operation for octopus in Western Australia was instigated

by Japanese scientists during 1979 – 1981 as a mitigation strategy for the extensive octopus’

predation and continual bycatch within the Western Australian rock lobster (Panulirus cygnus)

fishery (Joll, 1977). At present, commercial octopus fishing yields approximately 204 tonnes

from three primary fisheries; the Western Australian Developmental Octopus Fishery, the

West Coastal Rock Lobster Fishery and the Cockburn Sound Pot and Line Fishery (Hart et

al., 2016; Sauer et al., 2019) Since the establishment of the Western Australian Octopus

Fishery in 2001, the primary trap used has been passive shelter pots; a light, open-ended

plastic pot which is attached to a

Page 20: The ecology of Octopus aff. O. tetricus and prevalence of

20

demersal long line and deployed in water < 30 m. Due to the tendency for the shelter pots to

be buried or lost in high energy areas, the fishery operations have been limited to shallow areas

around the Perth coastline. The Perth fishery consistently yielded 30 tonnes per year while

exclusively using passive shelter pots (Sauer et al., 2019). In 2010, active trigger traps, consisting

of a rectangular trap baited with a plastic crab attached to a tripwire that activates a trap door when

seized by an octopus (Figure 1.1), were introduced into the fishery (Leporati et al., 2015). These

traps are considerably heavier and require less soak time than passive pots and can be used in

deeper waters around Perth. The introduction of trigger traps in Perth saw the catches increase

substantially to 170 tonnes in the first year of their use. Research conducted by Leporati et al. (2015)

suggested that the catch composition of each trap type also differed significantly, with passive traps

predominantly catching smaller, maturing individuals of both sexes and trigger traps catching

predominantly large (>1 kg), mature males with very few mature females. However, the authors

state that there was no overlap of depths in which each trap type was deployed, therefore

differences in catch compositions between traps cannot be determined conclusively.

The target species in the Western Australian Octopus Fishery is Octopus aff. O. tetricus

although other species are also caught occasionally, such as O. cyanea off the northern coast

and Macroctopus maorum in southern waters (Hart et al., 2016).

Figure 1.1 Cradle of the trigger traps used in the Western Australian Octopus Fishery.

The volume of trap ~15 L, opening area ~12.8 cm2 (Leporati et al., 2015).

Page 21: The ecology of Octopus aff. O. tetricus and prevalence of

21

1.5 Rationale and study aims

The overall aim of this Honours project is to investigate the spatial and temporal

distribution patterns of Octopus aff. O. tetricus in the nearshore, coastal waters in North

Fremantle, Western Australia (Figure 2.1). This region is located ~7 km north of the

traditional commercial fishing grounds. The North Fremantle site was selected following

discussions with commercial octopus fisher (P. Jasinski) about catches of small juvenile

octopuses in the area, indicating that this area is potentially a spawning or nursery area for O.

tetricus. The specific aims of the thesis are to:

(i) Investigate how octopus abundance varies with depth and sex;

(ii) Determine the size distribution and growth of octopus within the area; and

(iii) Compare the size distribution of octopus from this area with that on the fishing

grounds in Cockburn Sound, about 7 km south of the North Fremantle study site.

Detailed studies of octopus diet in shallow and deep waters of the area north of

Fremantle were also planned. Although gastric tracts were removed during dissection and all

dietary items in the gastric mill and stomach identified, it was not possible to analyse these

data comprehensively within the time constraints of Honours. However, preliminary findings

are presented from this component of the research. During the dissection, high numbers of

parasites were identified, and an additional aim was introduced:

(iv) Provide a preliminary description of the diet of octopus in coastal waters, and

(v) Determine the incidence, distribution, and abundance of parasites in octopus in the

coastal waters north of Fremantle and compare this (a) size classes of octopus, (b)

between shallow and deep water and (c) with that in Cockburn Sound.

2. Methods

2.1 Study site

Octopus’ were collected within the near coastal waters between North Fremantle and

Cottesloe, Perth, Western Australia during contracted commercial fishing in the area (31°58 S

115°49 E) (Fig. 2.1). This area is north of the main fishing grounds of the Octopus Fishery,

which is conducted within the Cockburn Sound (DoF, 2015), The study site is located

approximately 1.5 km from the nearest land, where depths range from ~8 to 22 m and the

Page 22: The ecology of Octopus aff. O. tetricus and prevalence of

22

benthic habitat consists of a sandy substrate with seagrass and macroalgal beds dispersed

irregularly throughout (P. Jasinski, Octopus fisher, pers. comm.).

2.2 Collection of Octopus aff. O. tetricus

Octopus were collected using open-ended, weighted, passive shelter pots (Figure. 2.2),

that rely on the octopuses entering the pots for shelter. Shelter pots have a 16.9 cm2 opening,

are 38 cm long and have a volume of 6 L. A commercial fishing vessel, chartered by the

Department of Primary Industries and Regional Development (DPIRD), deployed four

demersal long lines (1 – 1.4 km in length), each comprising 150 pots placed < 10 m apart, and

left in the water for soak times of ~7 to 22 days on each sampling occasions. A total of seven

separate sampling trips were conducted between the first haul on 23rd of March and ending

with the final retrieval of the lines on 15th of June 2020. Two lines (A and B) were set in shallow

water depths of 8 to 10 m and the other two lines (C and D) were set in water > 10 m deep

(Figure. 2.1, see Appendix 1.1 for coordinates of each line). The lines were reset in the same

location each time after being hauled following the normal commercial octopus fishing

operation (Figure. 2.1). The lines were separated by distances of ~500–1500 m and covered

an area of ~7.8 km2. One sample of octopus mantles was obtained from the commercial

octopus’ fishery which operates within Cockburn Sound and, in surrounding waters of Carnac

and Garden Island on 1st of April 2020. Data on size distribution, sex ratios, gonad condition

and parasites from this sample were compared with those from the main study area at the same

time.

After collection, octopus' mantles were severed from their arms, following the

standard commercial practice, and the mantles placed immediately placed on ice. The arms of

the octopus of commercial size i.e. > ~300 g wet weight, were kept by the commercial

fisherman for sale. Whole small octopus’ i.e. < ~300 g were retained on the sampling trips

between 14th April and 15th June and placed on ice. Samples were frozen within 2 h of capture.

Page 23: The ecology of Octopus aff. O. tetricus and prevalence of

23

Figure 2.1 Map showing the location of the four octopus trap lines set between March

and June 2020 in the coastal waters north of Fremantle Harbour, Western

Australia. Each line had 150 traps and was ~1 to 1.5 km in length.

Figure 2.2 Passive shelter pots used in the Western Australian developmental octopus’ fishery (from Leporati et al. 2015). Trap dimensions are 38 x 14 x 17 cm.

Page 24: The ecology of Octopus aff. O. tetricus and prevalence of

24

2.3 Biological processing of Octopus aff. O. tetricus

After defrosting the mantles, the mantle weight (MW; to 0.1g) and mantle length (ML;

mm) were recorded for each octopus. The small octopuses were weighed whole (TW).

The total weight (TW) was estimated from ML using the empirical relationships in Hart et al.

(2019):

Females: TW = 0.00054 ML 2.94 (r2 = 0.90, n = 345)

Males: TW = 0.0005 ML 2.86 (r2 = 0.90, n = 372)

The mantle weight of small juvenile octopus collected was estimated using the empirical

relationships in Hart et al. (2019):

Females: MW = 0.0004 ML 2.94 (r2 = 0.90, n = 10)

Males: MW = 0.0005 ML 2.86 (r2 = 0.90, n = 2)

2.3.1 Maturation

The mantles were then dissected, and the sex of the octopus determined from a

macroscopic inspection of the gonads. The gonads were removed and weighed to the nearest

0.01 g, and the maturity stage was determined by visual examination using the classification

scheme outlined in Leporati et al. (2015) and summarised in Table 2.1. The gastric tract,

including the gastric mill and stomach, was also removed from the mantle for dietary studies.

Although all samples have been analysed to identify and quantify dietary items, these data are

not presented as part of the Thesis due to time constraints. However, details of the processing

of samples are provided below.

Page 25: The ecology of Octopus aff. O. tetricus and prevalence of

25

Table 2.1 Summary of female and male gonad maturation stages of Octopus aff. O. tetricus from

Leporati et al. (2015). Note, Stage IV males were not described by Leporati et al.

Gonad stage Female Male

I Immature Ovary whitish, very small, no signs of granulation. Spermatophoric organ transparent and whitish.

II Developing Ovary yellowish with signs of granulation. Spermatophoric organ with white streaks of sperm

III Mature Ovary very large, yellow/ orange in colour, or clear if in the process of laying eggs. Oviducts and oviducal glands enlarged

Needhams sack full of spermatophores

IV Spent Ovary purple in colour and flaccid, with few to no eggs

2.3.4 Parasites

During preliminary stages of dissecting octopus’ mantles, several parasites were

observed within the viscera. Consequently, the internal organs of the mantle i.e. the crop,

digestive gland, alimentary canal, gonads and tissues, were examined for parasites under a

dissecting microscope. Parasites were removed with tweezers, counted, with their location

recorded and preserved in 70% ethanol.

The parasites found in the mantle and its associated organs were tentatively identified

as anisakid nematodes by Professor Alan Lymbery (Murdoch University). The identification

of nematodes to greater resolutions requires the use of molecular methods which was not

possible within the scope of this Thesis. Therefore, their identification was only possible to

family level (Anisakidae).

2.4 Data analyses

All data exploration and analyses were carried out using R version 4.0.2 (R Core Team,

2020) in the integrated development RStudio, version 1.3.1093 (RStudio Team 2020). All

figures were created using Microsoft ExcelTM.

Prior to all parametric statistical analysis, the assumptions of normality and

homogeneous variance were assessed by using Q-Q plot and Levene’s test, respectively. When

violations of assumptions were detected, the data were transformed.

2.4.1 Environmental parameters

Data on the monthly mean minimum and maximum sea surface temperatures for the

period from July 2019 to June 2020 were obtained from seatemperature.org that summarises

Page 26: The ecology of Octopus aff. O. tetricus and prevalence of

26

NOAA daily satellite readings (http://www. seatemperature.org, 30th October 2020). This covers

the sampling period and the months before sampling commenced.

2.4.2 Distribution and abundance of octopus

A three-way analysis of variance (ANOVA) was used to test whether the numbers of

octopus differed between sexes (Sex), water depths (Depth) and sampling times (Haul). A post

hoc Tukey’s Honest Significance Difference (HSD) was used to investigate any significant

difference between levels of factors. Abundances were compared without adjusting for soak

times.

2.4.3 Octopus size structure

The overall mean (± 1 SE) of the mantle weight (MW) and mantle length (ML) of

Octopus aff. O. tetricus males and females were calculated. Preliminary diagnostic plots

established that no variables violated the assumptions of the T-test and so the data were not

transformed. A two-sample independent T-test was used to test whether mean values for MW

and ML differed significantly between male and female octopuses.

A two-way analysis of variance (ANOVA) was performed to test whether ML differed

significantly between the North Fremantle and Cockburn Sound sites and between sexes.

2.4.4 Mantle length distributions

Octopuses were grouped into 10 mm mantle length (ML) classes and length percentage

frequency histograms were constructed for each sex, both depth regions and each sampling

trip, including octopus collected from the Cockburn Sound site. Descriptive statistics were

used to summarise the size distribution of the mantle length (ML).

The mean size of octopus’ from North Fremantle and Cockburn Sound on 1 April was

compared using an independent T-test.

2.4.5 Length-weight relationships

The relationships between mantle length (ML) and mantle weight (MW) for Octopus

aff. O. tetricus were estimated using a power function in Microsoft Excel for each sex. The

equation for the length-weight relationship was MW = aMLb where a and b are constant.

In R, the ML and MW data were transformed (Log10) and a linear equation was fitted to the

data:

Page 27: The ecology of Octopus aff. O. tetricus and prevalence of

27

𝐿𝑜𝑔10(𝑀𝑊) = 𝑎𝐿𝑜𝑔10(𝑀𝐿) + 𝑏

Analyses of covariance (ANCOVA) was used to determine if the relationship between MW

and ML differed between the sexes (α = 0.05) using ML as the covariate.

2.4.6 Gonadosomatic Index

The gonadosomatic index (GSI) was calculated as a percentage for each octopus as the

gonad weight (GW) divided by the mantle weight (MW) minus the gonad weight (GW):

𝐺𝑊 𝐺𝑆𝐼 = (

𝑀𝑊 − 𝐺𝑊) ∗ 100

The mean GSI (± 1 SE) was calculated for each month for males and females separately, for

both North Fremantle and Cockburn Sound samples. The monthly proportion of O. tetricus at

each stage of gonad development was plotted separately for males and females in North

Fremantle.

2.4.7 Parasites

The intensity (i.e. the number of parasites per octopus) and prevalence (i.e. the number

of octopuses with at least one parasite) was calculated following the definitions given in Bush

et al. (1997). Generalised linear models (GLMs) were used to test whether the presence of

parasites was influenced by depth, month, sex and size of an octopus. Octopuses were assigned

to three size groups based on mantle length classes; i.e., small = < 80 mm, medium = 80-150

mm and large = ≥160 mm, in an attempt to measure the influence of octopus size on parasite

infection.

Three binomial GLMs were used to test the presence of at least one parasite per octopus.

For each GLM the binomial response variable was the presence of nematodes (presence = 1,

absence = 0), which was realised through the logit link function. The binomial model was:

𝑝 𝐺(𝑦) = 𝑙𝑜𝑔𝑖𝑡 (

(1 − 𝑝)

) = 𝑥𝑇𝛽

where, p is the probability of at least one parasite being present, and 𝑥𝑇 denotes the vector of

predictor variables and 𝛽 is the vector of coefficients.

The first model aimed to assess the predictor variables of time (month), size and sex for

the North Fremantle O. tetricus population. The second model used size and sex as predictor

variables for the population from Cockburn Sound and the third model, tested for differences

Page 28: The ecology of Octopus aff. O. tetricus and prevalence of

28

between North Fremantle and Cockburn Sound, using only samples from North Fremantle

collected on the 1st April 2020

2.5 Diet Analysis

During dissection, the gastric tract was removed, with the crop being clamped to avoid

stomach and crop contents mixing. The stomach fullness was estimated by eye on a scale 0 to

10, with 10 being the equivalent to 100% full. The contents from the gastric tract were rinsed

under water through a 500 µm sieve to remove silt to ease in the identification of prey

components. Contents were examined under a dissecting microscope and identified to the

lowest taxonomic level possible, based mainly on an examination of hard parts such as

crustacean exoskeletons, teleost vertebrae, radulae of molluscs and the byssus strands of

bivalves. Cephalopods could be identified by flesh with distinct appearance of chromophores

and suckers, remnants of beaks and presence of black ink throughout the gastric tract. The

contribution of each prey item was estimated by eye and expressed as a percentage. Any prey

items which were heavily masticated or in an advanced stage of digestion was recorded as

‘unidentified’. The percentage frequency of occurrence was calculated and the %volume of

each item was estimated. Only preliminary findings from the dietary studies are presented in

the results.

2.6 Permits and approval of ethics committees

All necessary permits were obtained for this study. Octopus were collected by the

commercial fisher for this research under Murdoch Animal Ethics – Cadaver and/or Tissue

Notification, Permit No. 744 “Cadaver application for octopus’ ecology”.

3. Results

3.1 Water temperature and storms

The mean maximum water temperature derived from NOAA data on SST during the

period from March to June 2020 ranged from 25.1 °C in March to 23 °C in June (Figure 3.1).

In the 12 months before this, the lowest mean maximum water temperature was in September

2019 (19.9 °C) and the highest mean (25.3 °C) was in February 2020.

Page 29: The ecology of Octopus aff. O. tetricus and prevalence of

29

During the soaking period of the fourth and fifth haul, considerable storm activity in

Perth caused the loss of several octopus’ pots and moved large amounts of sand into the traps,

resulting in the suffocation of at least one octopus.

27

25

23

21

19 Minimum

17 Maximum

15

Month Study period

Figure 3.1 Mean sea surface temperature in North Fremantle, Western Australia from daily

satellite readings extracted from http://www.seatemperature.org (30th October

2020). Arrow on the X-axis shows the duration of sampling octopus in the region

3.2 Changes in numbers with depth and month

A total of 713 octopuses were collected during the seven sampling trips between 23rd

March and 16th June 2020. These comprised 344 females, 358 males and 12 individuals whose

sex could not be determined. The highest number of octopus were caught from Line B (n =

194), followed by Line A (n = 172), both in waters < 10 m deep. A similar number were caught

on Line C (n = 163) and the lowest numbers from Line D (n =140) in the deepest water (> 10

m). A total of 49 octopuses (30 females, 19 males) were collected from the Cockburn Sound

on 1st April 2020. This compares with the second haul, caught in the waters of North Fremantle

on 1st April 2020 of 90 octopus, with 43 males and 47 females.

The third haul on 14th April 2020 yielded the highest number of octopuses (135) when

54 males and 78 females were caught (Figure 3.2). The last haul yielded the lowest number of

octopus (70), with equal numbers of males and females. The lowest number of males (6) per

line was collected on 2nd June and the lowest number of females (5) per line on 11th May, both

Tem

per

atu

re °

C

July

Au

gust

Sep

tem

ber

Oct

ob

er

No

vem

ber

Dec

emb

er

Jan

uar

y

Feb

ruar

y

Mar

ch

Ap

ril

May

Jun

e

Page 30: The ecology of Octopus aff. O. tetricus and prevalence of

30

on Line D. The biggest discrepancies between male and female catches were recorded was on

the 23rd March on Line B (9 females, and 23 males) and on 14th April on Line D (21 females,

and 7 males).

A two-way analysis of variance (ANOVA) showed that the mean number of octopuses

collected per haul differed significantly (P = 0.01) among hauls but not between depths or sexes

and that none of the interaction terms were significant (P >0.05, Table 3.1). The Tukey’s post

hoc test showed that only the third and seventh haul differed significantly (P = <0.05), with a

significantly higher mean catch on the third haul (14th April, 16.5±0.49) than on the 7th haul

(15th June, 8.4±0.20).

Page 31: The ecology of Octopus aff. O. tetricus and prevalence of

31

Nu

mb

er o

f O

cto

pu

ses

Nu

mb

er o

f O

cto

pu

ses

a) c)

25 25

20 Females 20 Males

15 15

10 10

5 5

0 0

Haul Haul

b) d)

25 25

20 20

15 15

10 10

5 5

0 0

Haul Haul

Figure 3.2 The number of female (black) and male (grey) Octopus aff, O. tetricus collected on each line from North Fremantle, Western Australia for each sampling trip between 23rd March and 15th June 2020. a) Line A, b) Line B, c) Line C and d) Line D. Mean depths of lines were: A = 9.29 m; B = 9.64 m; C = 16.79 m; D = 15.34 m.

Nu

mb

er o

f O

cto

pu

ses

Nu

mb

er o

f O

cto

pu

ses

23

-Mar

2

3-M

ar

1-A

pr

1-A

pr

14

-Ap

r 1

4-A

pr

24

-Ap

r 2

4-A

pr

11

-May

1

1-M

ay

2-J

un

2

-Ju

n

15

-Ju

n

15

-Ju

n

23

-Mar

2

3-M

ar

1-A

pr

1-A

pr

14

-Ap

r 1

4-A

pr

24

-Ap

r 2

4-A

pr

11

-May

1

1-M

ay

2-J

un

2

-Ju

n

15

-Ju

n

15

-Ju

n

Page 32: The ecology of Octopus aff. O. tetricus and prevalence of

32

1

1

Table 3.1 Summary of results for the three-way analysis of variance (ANOVA) to test differences in the abundance of Octopus aff. O. tetricus between depth regions, sex, and sampling haul. Significant factors are in bold. df = degrees of freedom, MS = mean squares.

Source of variation df MS % MS P-value

Haul 6 49.03 26.94 0.01

Depth 1 27.16 14.92 0.18

Sex 1 17.16 9.43 0.29

Haul*Depth 6 27.41 15.06 0.12

Haul*Sex 6 32.99 18.13 0.07

Depth*Sex 1 0.16 0.09 0.92

Haul*Depth*Sex 6 13.49 7.41 0.49

Residuals 28 14.59 8.02

3.3 Sex ratio

Overall, the sex ratio for female and male Octopus aff. O. tetricus sampled from the

North Fremantle site was 1.08:1 (339 males, 313 females), which does not differ significantly

from 1:1 (χ21 = 0.959, P = 0.328) (Table 3.2). Neither did the overall sex ratio differ

significantly between shallow and deep locations (χ21 = 0.041, P = 0.8). Comparisons of the

sex ratios across the study region (i.e. pooled for shallow and deep water) for each haul showed

a significant dominance of males in the catches in May (1.89:1) (χ21 = 7.440, P <0.01) but that

sex ratios did not differ significantly from 1:1 for any other haul. (Table 3.2).

Within the shallow region, the only haul in which the sex ratio differed significantly

from 1:1 was the first haul (23 March, χ21 = 4.063, P = 0.04), with a male to female ratio of

1.73:1. In the deeper region, sex ratios differed from 1:1 in two hauls: on the third haul (14

April) significantly more females than males were caught (0.58:1, χ2 = 3.750, P = 0.05), and

on the fifth haul, significantly more males than females were present (1.89:1, χ2 = 5.357,

P = 0.02) (Table 3.2).

The sex ratio for octopuses collected from Cockburn Sound (0.60:1) did not differ significantly

from that for octopuses sampled from North Fremantle on 1st April 2020 (0.91:1) (χ21 = 0.956,

P = 0.328).

Page 33: The ecology of Octopus aff. O. tetricus and prevalence of

33

Table 3.2 The observed sex ratio (male: female), sample sizes (n), P-values (P) and chi-squared values (χ2), presented for overall ratio, by month and the two depth regions for Octopus aff. O. tetricus collected from North Fremantle near-coastal waters between March and June 2020. Bold values = Significant differences from 1:1

Haul <10 m n χ2 P >10 m n χ2 P Overall n χ2 P

23-Mar 1.73:1 63 4.063 0.04 1.05:1 37 ~0 ~1 1.43:1 100 2.890 0.09

1-Apr 0.72:1 55 1.163 0.28 1.34:1 35 0.457 0.50 0.91:1 90 0.100 0.75

14-Apr 0.80:1 72 0.681 0.41 0.58:1 60 3.750 0.05 0.69:1 132 4.008 0.05

24-Apr 1.35:1 56 0.446 0.50 0.89:1 34 0.029 0.86 1.09:1 90 0.100 0.75

11-May 1.63:1 42 1.929 0.17 2.23:1 42 5.357 0.02 1.89:1 84 7.440 <0.01

2-Jun 1.05:1 39 ~0 ~1 1.32:1 51 0.706 0.40 1.20:1 90 0.544 0.45

15-Jun 1.12:1 36 0.028 0.868 31 0.129 0.72 0.97:1 67 ~0 ~1

Overall 1.10:1 362 0.798 0.371 1.06:1 290 0.169 0.681 1.08:1 652 0.959 0.327

3.4 Size distribution

3.4.1 Overall size distribution

A total of 709 octopuses were measured and weighed. The smallest female was

measured at 35 mm ML, 31.94 g total weight (TW) and as the octopus was taken whole the

mantle weight was estimated at 5.13 g. The largest female was 211 mm ML, 687.72 MW and

an estimated TW of 2397.96 g (Table 3.3). The smallest male was 14 mm ML larger than the

smallest female (mm), but the largest male was a similar size to the largest female (Table

3.3). The mean estimated TW for females was

882.38 g which was significantly lighter than that of males (1253.01 g) (T = -7.926, df =

633.56, P <0.001). The mean ML and MW did not differ significantly between females (141.49

mm, 262.52 g) and males (144.83 mm, 264.44 g) (P > 0.5, Table 3.3).

Eleven of the small juvenile O. tetricus were too small to determine their sex. The

smallest of which was 14 mm ML and 2.75 g TW. The average TW for undetermined

individuals was 3.70±1.70 g and average ML was 35.3±6.82 mm.

Females collected from the Cockburn Sound site ranged from 102 to 217 mm ML and

71.3 to 588.8 g MW, with an estimated TW range of 299.9 to 2598.2 g. Males collected from

Cockburn Sound ranged from 99 to 191 mm ML, 80.5 to 441.74 g MW and an estimated total

weight range of 368.2 to 2542.2 g.

Page 34: The ecology of Octopus aff. O. tetricus and prevalence of

34

Table 3.3: Summary of the mean sizes (± 1 SE) and range of mantle length (ML), mantle weight (MW) and total weight (TW), for Octopus aff. O. tetricus caught in waters north of Fremantle, Western Australia. The t-statistic (T) and significance (P) of the Welch two-sample t-test between sexes also shown. df = degrees of freedom

Sex n Mean±SE Range T P

Mantle length (mm) (df = 611.05)

F 313 141.49±1.95 35-211

M 339 144.83±1.57 49-218 -1.333 0.18

Mantle weight (g) (df = 604.99)

F 313 262.52±8.35 5.13-687.72

M 339 266.86±6.59 15.22-691.62 -1.333 0.18

Estimated total weight (g) (df = 633.56)

F 313 882.38±28.59 31.94-2598.18

M 329 1253.01±34.75 29.95-3750.01 -7.926 <0.001

The total weight (TW) estimated from the equation in Hart et al. 2019

The mean MLs in the shallow and deep waters was almost identical (< 10: 143.21 mm,

> 10: 143.25 mm, P ~ 1). The mean MW for the shallower region was 268.52 g, slightly, but

not differ significantly greater than that for the deeper region of 260.11 g.

3.4.2 Mantle Length-weight relationship

An analysis of covariance (ANCOVA) found that the slopes of the relationship between

Log mantle length (ML) and the Log mantle weight (MW) (Figure 3.3) did not differ

significantly between males and females, therefore the data were pooled. The equations for the

relationship was:

Log(MW) = 2.483Log(ML) – 6.849 (n = 696, r2 = 0.92)

Table 3.4 Summary of results from the analysis of covariance (ANCOVA) to test whether the

relationship between mantle weight (MW) and mantle length (ML) differed

significantly between sexes in Octopus aff. O. tetricus. Both MW and ML were log-

transformed.

Variable df MS F P

LogML 1 259.38 5137.46 < 0.001

Sex 1 0.002 0.03 0.853

LogML*Sex 1 0.184 3.65 0.057

Residuals 646 0.050

Page 35: The ecology of Octopus aff. O. tetricus and prevalence of

35

a)

b)

Figure. 3.3 The relationship between mantle length (ML) and mantle weight (MW) and mantle

length (ML) for Octopus aff. O. tetricus for (a) untransformed and (b) log-

transformed (Log) data sampled from North Fremantle, Western Australia collected

between March and June 2020. (n = 655).

3.4.3 Size distribution in deep and shallow water

The mantle length (ML) size distribution pooled over all sampling times and depths

showed that most male and female octopus were between 120 and 170 mm long with modal

sizes of 150 mm ML for both males and females (Figure 3.4a). The number of females in each

size class declined more sharply than the number of males between 160 and 190 mm and few

male or female octopus >200 mm ML were caught.

The modal ML class for females in the < 10 m depth region was 160 mm and 150 mm

for the >10 m region. The modal ML class for males was the same for both depth regions (150

mm). The largest individuals for both sexes were collected from the shallow region. The

smallest males were found in the shallow waters (Figure 3.4b) in comparison, the smallest

females were found in the deeper region (Figure 3.4c). There is a sharp decline in the number

of females after the 150 to 160 mm ML classes for both shallow and deep regions.

Page 36: The ecology of Octopus aff. O. tetricus and prevalence of

36

a)

20 Females (n = 313)

15 Males (n = 340)

10

5

0

Mantle Length Classes (mm)

b)

20

15

Females (n = 191)

Males (n = 172)

10

5

0

Mantle Length Class (mm)

c) 20

18 16 14 12 10

8 6 4 2 0

Females (n = 141)

Mantle Length Class (mm)

Figure 3.4 Mantle length-frequency distributions for female and male Octopus aff. O. tetricus for pooled (a), ≤10 m (b) and > 10 m (c) depth areas in North Fremantle, Western Australia. Collected between March and June 2020

Fre

qu

ency

%

Fre

qu

ency

%

Fre

qu

ency

%

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

Page 37: The ecology of Octopus aff. O. tetricus and prevalence of

37

3.4.4 Temporal variation in size

One small juvenile male was collected during the first sampling trip on 23rd March (0-

9.9 mm ML class). Except for this individual, small octopus (< 80 mm ML class) were not seen

until the third sampling trip, hauled on 1st April. Individuals in the smallest length classes (<

80 ML mm) were the least common across the sampling period, with small females (1.84%)

being more common than small males (0.61%). Small juveniles of indeterminate sex were

collected on the 14th April (n = 6, mean ML 48.6 mm), 11th May (n = 2, mean ML 22.5 mm)

and 2nd June (n = 3, mean ML 18.34) trip but are not shown on the ML frequency figure

(Figure 3.5).

Female length distribution varied across the sampling period but tended to decrease

from March to June, which contrasts with males, who increased in size across the sampling

periods (Figure 3.5). Throughout the majority of the sampling period, the modal ML classes

for females stayed between 150 to 160 mm but was more variable in the third sampling trip

(Figure 3.5a). From the fifth sampling trip, a second modal ML class can be identified for

females at 100 mm and remains consistent between 100 to 130 mm for the until the conclusion

of the hauls. The ML size distribution for males varied over the four months of sampling.

Initially, the ML modal length for males was 130 mm, over the duration of the sampling period

the modal length for males increased to 160 – 170 mm. From 14th April onwards most males

were above 140 mm.

The mean ML for females above 80 mm ML was highest in March (158.3 mm) and

lowest in May (133.7 mm). The mean ML for males gradually increased during the study, with

the lowest mean recorded on 1st April (140.7 mm) and the highest mean on 15th June (153.6

mm) (Figure 3.6). The mean ML of male and female octopus >80 mm ML was almost identical

on April 1 and 24 (~140‒142 mm, Figure 3.6).

A two-way analysis of variance (ANOVA) showed that the mean ML differed

significantly among hauls (P < 0.001) and that the haul x sex interaction was significant (P =

0.01) (Table 3.5). The factor Haul accounted for 46.1% of the total mean squares and the

interaction for 30.5%. A Tukey’s post hoc test showed that the mean ML for females collected

on the 23rd March (156.44 mm) was significantly longer than that on 14th April (131.95 mm)

(P = 0.003). The mean ML for female octopuses collected on 11th May (125.72 mm) were

significantly smaller than those collected on March 23rd (Tukey post hoc P = 0.004). In contrast

Page 38: The ecology of Octopus aff. O. tetricus and prevalence of

38

to females, the mean ML of males did not differ significantly between any of the sampling trips

(P >0.05), which explains the significant sex x haul interaction.

Page 39: The ecology of Octopus aff. O. tetricus and prevalence of

210

Frequency % Frequency % Frequency %

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

Frequency % Frequency % Frequency %

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

30

40

50

60

70

80

90

100

110

120

130

140

150

160

170

180

190

200

210

a) d

)

30

3

0

Females (n

= 43

)

Males (n

= 47

)

Females (n

= 41

)

Males (n

= 59

) 2

5

25

20

2

0

15

1

5

10

1

0

5

5

0

0

Man

tle Length

Class (m

m)

Man

tle Length

Class (m

m)

b)

e)

30

3

0

Females (n

= 47

)

Males (n

= 43

) 2

5

25

Fem

ales (n = 2

9)

Males (n

= 55

) 2

0

20

15

1

5

10

1

0

5

5

0

0

Man

tle Length

Class (m

m)

Man

tle Length

Class (m

m)

c) f)

30

3

0

Females (n

= 78

)

Males (n

= 54

) 2

5

25

Fem

ales (n = 4

1)

Males (n

= 49

) 2

0

20

15

1

5

10

1

0

5

5

0

0

Man

te Length

Class (m

m)

Man

tle Length

Class (m

m)

39

Page 40: The ecology of Octopus aff. O. tetricus and prevalence of

40

g) 30

25

20

15

10

5

0

Mantle Length Class (mm)

Figure 3.5 Mantle length frequencies for female and male Octopus aff. O. tetricus for each

sampling trip collected from North Fremantle, Western Australia between March

and June 2020. a) = 23rd March, b) = 1st April, c) = 14th April, d) = 24th April, e) = 11th

May, f) = 2nd June, g) = 15th June.

180

160

Females

Males

140

120

100

80

Haul

Figure 3.6 Mean (± 1SE) mantle length (ML) of Octopus aff. O. tetricus larger than 80 mm for each haul in North Fremantle, collected between March and June 2020.

Females (n = 34)

Males (n = 33)

Man

tle

Len

gth

(m

m)

Freq

uen

cy %

23

-Mar

3

0

40

50

60

70

80

90

10

0

11

0

12

0

13

0

14

0

15

0

16

0

17

0

18

0

19

0

20

0

21

0

1-A

pr

14

-Ap

r

24

-Ap

r

11

-May

2-J

un

15

-Ju

n

Page 41: The ecology of Octopus aff. O. tetricus and prevalence of

41

Table 3.5 Summary of the results from the two-way analysis of variance (ANOVA) to test if the

mean mantle length (ML) of Octopus aff. O. tetricus differed significantly between

sexes and among each haul from March to June, in North Fremantle, Western

Australia. Bold = significant terms

Variable df MS MS% P

Haul 6 4014 46.1 < 0.001

Sex 1 1067 12.3 0.293

Haul*Sex 6 2654 30.5 0.01

Residuals 636 965 11.1

3.4.5 Cockburn Sound size distribution

Octopus collected from the Cockburn Sound in April 2020 had a modal ML of 120 mm

for males and 160 mm for females (Figure 3.7a). This is smaller than the modal sizes for males

(140 to 150 mm) and larger for females (150 mm) caught in North Fremantle (Figure 3.7b).

A two-way analysis of variance (ANOVA) showed the mean ML for octopuses

collected on April 1 differed significantly between North Fremantle and Cockburn Sound

(P = 0.003) and sexes (P = 0.02) and that the site and sex interaction was significant (P = 0.03)

(Table 3.6). The mean ML for male (139.78 mm) and females (162.27 mm) collected from the

Cockburn Sound differed significantly (P = 0.02). The mean ML of female octopus sampled

from Cockburn Sound was 162.27 mm which was significantly longer than that from North

Fremantle (140.36 mm). Moreover, the males collected from North Fremantle were

significantly smaller than the females collected from Cockburn Sound on the same date (Tukey

post hoc test P = < 0.05). The mean ML for males did not differ significantly between the sites

(Table 3.6).

Page 42: The ecology of Octopus aff. O. tetricus and prevalence of

42

a)

30

25 Females (n = 30)

Males (n = 18) 20

15

10

5

0

Mantle Length Class (mm)

b)

30

25

20

15

10

5

0

Females (n = 47)

Males (n = 43)

Mantle Length Class (mm)

Figure 3.7 Mantle length frequencies for male and female Octopus aff. O. tetricus for samples

collected from the Cockburn Sound (a) and North Fremantle (b), collected from 1st

April 2020.

Table 3.6 Summary of results for the two-way analysis of variance (ANOVA) to test differences

in mean mantle length (ML) of Octopus aff. O. tetricus collected between the two sites (North Fremantle and Cockburn Sound) and sex on 1st April 2020, df = degrees of freedom, MS = mean squares.

Source of variation df MS % MS P-value

Sex 1 2788 21.22 0.04

Site 1 6683 50.86 0.002

Sex*Site 1 3029 23.01 0.03

Residuals 134 640 4.87

Fre

qu

ency

%

Fre

qu

ency

%

30

40

50

60

70

80

90

10

0

11

0

12

0

13

0

14

0

15

0

16

0

17

0

18

0

19

0

20

0

21

0

30

40

50

60

70

80

90

10

0

11

0

12

0

13

0

14

0

15

0

16

0

17

0

18

0

19

0

20

0

21

0

Page 43: The ecology of Octopus aff. O. tetricus and prevalence of

43

3.5 Reproduction and Maturity

3.5.1. Gonad maturity stages

The majority (74.4%) of females sampled from North Fremantle between March and

June 2020 were immature, 23.72% were maturing, 1.60% mature and 0.32% spent (Figure 3.8).

Catches of immature females were lowest in March and peaked in May before declining again

in June in both depth regions. Mature females were only caught during April in both depth

regions. The only spent female was collected from the deep region in April (Figure 3.8).

Maturing males dominated catches at 46.38% overall, with immature males at 26.25%

and mature 25.39%. All male maturity stages were found each month. The rate of immature

males being caught decreased from March to June, with a more notable decline in shallower

waters. The lowest proportion of mature males was recorded in March in shallow waters in

April for the deep.

More immature females (232) were caught between March and June than males (89).

Most males (165) collected from North Fremantle were in the second maturity stage (maturing)

(Table 3.5). The mean mantle length and mantle weight for females were larger than males in

each maturity stage except for Stage III (mature) ML, in which the mean ML was ~171 mm

for both sexes. The mean gonad weight (GW) was considerably heavier for mature female

octopus (41.1 g) than that for mature males (17.8 g) (Table 3.5). One spent (Stage IV) female

was collected during this study on 14th April, an individual with a ML of 132 mm, weighing

233.4 g MW and 23 g GW (GSI = 10.92%) (not included in Table 3.5).

On 1st April, most females were classed as immature in samples from both North

Fremantle and Cockburn Sound (Figure 3.9). Cockburn Sound had considerably more maturing

females (36.67%) than North Fremantle (17.02%), but no mature females were recorded in the

Sound at this time. Males collected from both sites on the 1st April were predominantly at the

maturing stage, with 48.4% maturing males at North Fremantle and 72.2% from Cockburn

Sound. North Fremantle had a higher frequency for immature and mature males when

compared with Cockburn Sound (Figure 3.9).

Page 44: The ecology of Octopus aff. O. tetricus and prevalence of

44

Freq

uen

cy %

Fr

equ

ency

%

a) b)

n = 172 n = 141 100%

80%

60%

40%

20%

0%

March April May June

Month

Stage III

Stage II

Stage I

100%

80%

60%

40%

20%

0%

March April May June

Month

Stage IV

Stage III

Stage II

Stage I

c) d)

n = 191 n = 149

100%

80%

100%

80%

60%

40%

20%

0%

March April May June

Month

Stage III

Stage II

Stage I

60%

40%

20%

0%

March April May June

Month

Stage III

Stage II

Stage I

Figure 3.8 Percentage frequency of gonad stages for females (n = 318) (a) and males (n = 331) (b) Octopus aff. O. tetricus collected between 23rd March and 15th June 2020 from North Fremantle, Western Australia.

Freq

uen

cy %

Fr

equ

ency

%

Page 45: The ecology of Octopus aff. O. tetricus and prevalence of

45

Table 3.7 Summary of the mean sizes (± 1SE) for mantle length (ML), mantle weight (MW) and gonad weight (GW) for each maturity stage of Octopus aff. O. tetricus sampled from North Fremantle, collected between March to June 2020.

Sex n ML (mm) MW (g) GW (g)

Stage I (Immature)

F 232 131.23(±2.09) 209.59(±7.62) 2.58(±0.16)

M 89 110.24(±2.11) 124.60(±6.31) 3.23(±0.28)

Stage II (Maturing)

F 74 171.49(±2.56) 415.45(±13.4) 11.81(±0.85)

M 165 149.54(±1.45) 275.99(±5.82) 10.96(±0.26)

Stage III (Mature)

F 5 171.6(±13.3) 460.17(±86.1) 49.14(±9.49)

M 86 171.71(±1.72) 396.68(±8.20) 17.80(±0.31)

a)

100

80

60

North Fremantle (n = 48)

Cockburn Sound (n = 30)

40

20

0

b) 100

Immature Maturing Mature

Maturity Stage

North Fremantle (n = 51) 80

Cockburn Sound (n = 18)

60

40

20

0

Immature Maturing Mature

Maturity Stage

Figure 3.8 Percentage frequency of gonad stages for Octopus aff. O. tetricus collected from North Fremantle and Cockburn Sound on 1st April 2020, a) Females, b) Males.

Freq

uen

cy %

Fr

equ

ency

%

Page 46: The ecology of Octopus aff. O. tetricus and prevalence of

46

1

% ex d n I c i at m so Female o ad Male n Go

0

0 0 0 0

Month

1

% ex d n I c i at m so

Female o ad Male n Go

0

0 0 0 0

Month

3.5.2 Gonadosomatic Indices

The maximum GSI recorded for females from North Fremantle was 20.33% and 7.64%

for males. The monthly mean gonadosomatic index (GSI) for males increased from 3.35 % in

March to 4.17 % in June (Figure 3.7). In contrast, the mean GSI for females, which was always

lower than that for males, ranged from only 2.08% in March to 1.48% in June and did not show

an increasing trend (Figure 3.7).

The mean GSI for female octopuses collected from Cockburn Sound was marginally

smaller (1.30%) than those collected from North Fremantle on 1st April (1.4%). Similarly,

males collected from Cockburn Sound had a smaller mean GSI (3.40%) than those from North

Fremantle (3.78%) at this time.

Figure 3.9 Monthly mean gonadosomatic index (±1SE) for male (n = 312) and female (n = 291)

Octopus aff. O. tetricus for specimens collected in North Fremantle, Western

Australia between March and June 2020.

3.6 Preliminary diet results

In total, 701 octopus’ gastric tracts were examined and 67% contained dietary items.

Preliminary analysis of the dietary contents found that the dominant categories in both male

and female gastric tracts by frequency of occurrence were brachyurans (>60), teleosts (30 and

56) and non-cephalopod molluscs (~20) (Figure 3.10). The frequency of brachyurans and

5

4

3 Female

2 Male

1

0

March April May June

Month

Go

nad

oso

mat

ic In

dex

%

Page 47: The ecology of Octopus aff. O. tetricus and prevalence of

47

teleosts was higher for males than females. A total of 50 octopus contained cephalopods in

their stomachs (Figure 3.10).

80

Females (n = 205)

60 Males (n = 264)

40

20

0

Prey Species

Figure 3.10 The frequency of each prey item of Octopus aff. O. tetricus collected from North Fremantle and Cockburn Sound between March and June 2020.

3.7 Parasites

3.7.1 Parasite presence

Nearly 60% of all octopus in North Fremantle (58.19%) had anisakid nematodes

(hereafter nematodes) present in at least one internal organ. Only one parasite taxon was found

during the dissection of all octopus. In octopus with at least one parasite, it was always found

in the crop (100 % prevalence), 58.5% in the mantle muscles, 9.56% in the digestive gland,

2.82% in posterior salivary glands, 0.25 % in the gonads (Figure 3.11). Across all samples, 7

octopuses (1.53 %) had eggs of parasites present in the crop.

The mean intensity (± 1 SE) of nematode infection in North Fremantle was 18.31 ± 0.68

per octopus and ranged from 1 to 75 nematodes per octopus. Of the infected octopus, 59.9%

had nematodes in one internal organ, 32.2% in two and 8.0% in three or more organs. The

prevalence of nematodes was markedly higher in octopus from North Fremantle on April 1

(75.6%) than those from Cockburn Sound (25%) on the same date (Figure 3.11). The GLM of

nematode presence found that the odds of octopus collected from North Fremantle having at

Occ

urr

ence

Bra

chyu

ran

Her

mit

Cra

b

Iso

po

d

Po

lych

aete

Ost

raco

d

Dec

apo

d

Tele

ost

Cep

hal

op

od

Biv

alve

Am

ph

ipo

d

No

n-c

eph

alo

po

d M

ollu

sc

Gas

tro

po

d

Page 48: The ecology of Octopus aff. O. tetricus and prevalence of

48

least one nematode present were 9.27 (= e2.227) times greater (P < 0.001) than those from

Cockburn Sound (Table 3.8c).

Figure 3.11 Examples of Anisakidae infestation in Octopus aff. O. tetricus in crop (a) and

digestive gland (b) from samples collected in Perth, Western Australia between

March and June 2020 (Photo: Jorja Claybrook).

Page 49: The ecology of Octopus aff. O. tetricus and prevalence of

49

80

60

40

20

0

North Fremantle Cockburn Sound

Site

Figure 3.11 The prevalence of infections of anisakid nematode in Octopus aff. O. tetricus collected from North Fremantle and Cockburn Sound on 1st April 2020.

3.7.2 Parasite presence modelling

Binomial logistic regression with a logit link function was used to examining the

relationship between on the presence of at least one nematode and depth, haul and size of

octopus, using March and the size group 'Small' as the intercept. The odds of nematode

infestation for octopus were significantly greater for both larger classes of octopus than small

octopus: by 2.35 times (= e0.855) for medium (80-150 ML), and 2.69 times (= e0.991) for large

octopus (≥160 mm ML) than the small individuals (<80 mm ML) (Table 3.8a). The odds of

an octopus having a nematode present were significantly lower in June (0.58) than March (1.0)

(= e-0.549; Table 3.8a) The presence of nematodes did not differ significantly between sexes (P

= 0.9).

Nematode presence also differed significantly with octopus size in Cockburn Sound,

with the odds of nematode infestation being significantly higher for the medium (2.09, = e0.738)

and large (2.16, = e0.771) than small octopus (Table 3.6b). Like North Fremantle, the presence

of nematodes did not differ significantly between sexes in Cockburn Sound (P = 0.65, Table

3.8b).

n = 90

n = 49

Freq

uen

cy %

Page 50: The ecology of Octopus aff. O. tetricus and prevalence of

50

Table 3.8 Coefficients of binomial Generalised Linear Models (GLMs) analysis with logit link function for the presence of anisakid nematodes in Octopus aff. O. tetricus caught in North Fremantle and Cockburn Sound and the influence of octopus size, month and sex.

Variable Estimate SE Z -value P-value

a) North Fremantle (size + month + sex)

Intercept (Size: Small, March, Females)

-0.253

0.077

-0.776

0.438

Size: Medium 0.855 0.065 3.314 0.002 Size: Large 0.991 0.068 3.426 <0.001 April 0.066 0.243 -0.271 0.787 May -0.505 0.306 -1.650 0.099 June -0.549 0.265 -2.070 0.038 Males 0.022 0.163 0.135 0.893

b) Cockburn Sound (size + sex)

Intercept (Size: Small, Females) -0.471 0.255 -1.847 0.064 Medium 0.738 0.267 2.767 0.006 Large 0.771 0.280 2.753 0.006 Males 0.070 0.154 0.453 0.650

c) North Fremantle vs. Cockburn Sound

Intercept (Site: Cockburn Sound) -1.099 0.333 -3.296 <0.001 Site: North Fremantle 2.227 0.412 5.381 <0.001

4. Discussion

This study evaluated the spatial and temporal variability of the North Fremantle population of

Octopus aff. O. tetricus. The data used for these analyses were collected by commercial fishing

operation setting four lines of traps, two in shallow (≤ 10 m) and two in deeper water (>15-

17 m) on seven occasions between March and June 2020. Differences in size distribution,

gonad stage and body size were examined to test whether they differed between the sexes,

depths of water and sampling time. Preliminary investigations of octopus’ diet in this region

have also been made. During the study, an anisakid parasite was found infesting octopus’ and

data on their incidence and intensity of infection were collected to investigate the site of

infection. Comparisons of the biology and parasite infestation have been made between

octopus in North Fremantle and those collected at one time from octopus fishing grounds in

Cockburn Sound, ~7 km south of the main study.

Page 51: The ecology of Octopus aff. O. tetricus and prevalence of

51

4.1 Abundance and sex ratio

The abundance of octopus varied between sampling times but not between depths or

sexes of octopus. The major difference between sampling times was the significantly greater

numbers caught in early April compared with those in mid-June. The overall decrease in the

abundance of octopus in the North Fremantle site in June coincided with the decrease in SST.

Octopuses exhibit large fluctuations in abundance, mainly due to their susceptibility to

environmental factors during their paralarval stage (Pierce et al., 2008).

Significant storm activity in the region during the fifth and sixth hauls in late April and

May could have contributed to the lower catches at this time. The increased wave action

associated with the storms resulted in several octopus’ pots becoming submerged in sand and

at least one octopus was suffocated by sand (P. Jasinski, pers. comm.). Potentially, any small

juvenile octopus’, which are at mercy to the currents, may have swept out to other areas during

these weather events. Decreases in octopuses per haul from within the North Fremantle study

may also be the result of the targeted removal of octopuses during the study period. Decreases

in abundance due to targeted fishing pressure were examined in San Matas Gulf, northern

Patagonia. The relative abundance of O. tehuelchus decreased during the fishing season on

commercial fishing grounds compared with no change in abundance within a Marine Protected

Area (Narvarte et al., 2013)

The abundance of octopus was not significantly affected by depth (<10 m and ≥10 m)

within the waters of North Fremantle. While there was not any statistically significant

difference, Line C and D continually yielded the least number of octopus per haul. The soak

time of shelter pots was considered an important variable for the number of octopuses collected

per haul; it was expected that catches would increase with increasing soak time. However, the

number of octopus did not appear to be affected by the soak time of traps, as the hauls that

were submerged for the shortest amount of time and the haul which was submerged for longest

had very similar catches.

Sex ratio

The sex ratio did not differ significantly from 1:1, except on one occasion in the shallow

waters and twice in the deeper waters. On the 23rd March, nearly twice as many males were

caught in the shallow lines than females, while in the deep, females were significantly more

prevalent than males on the 14th April, but this was reversed in early May when more than

double the number of males to females were caught. The overall sex ratio did not differ for

Page 52: The ecology of Octopus aff. O. tetricus and prevalence of

52

either depth region or the Cockburn Sound population and did not differ between Cockburn

Sound and North Fremantle in early April.

The sex ratio for April is consistent with the known composition of O. tetricus in the

Western Australian fishery, where males tend to dominate the catches but have the lowest

frequency of capture in April and the highest in October (Hart et al., 2016; Hart et al., 2019;

Leporati et al., 2015). Until senescence, males will actively hunt and move around their

environment seeking potential mates, which explains why males typically dominate catches

(Leporati et al., 2015). Aside from the octopus’ collected in April, males consistently

dominated the catches over females. This is quite unusual to other octopus’ population in

which fluctuation between male and female abundance occur throughout the year, usually

coinciding with peaks in the GSI (Alonso-Fernández et al., 2017; Guard & Mgaya, 2002).

Similar patterns of male dominance have been observed in the Mexican O. maya fishery

(Duarte et al., 2018). Currently, female coastal octopuses are hypothesised to cease feeding

around the time they lay their eggs, seeking refuge in dens to care for their eggs until they

hatch (Hart et al., 2019). During this time the females are unlikely to be actively hunting and

would be less likely to enter traps (Leporati et al., 2015). It is not known precisely why the

Western Australian octopus’ fishery has mostly continuous male dominance, but it is likely

attributed to the use of active trigger trap use in the deep regions of the fishery (>30 m) which

is considerably more likely to capture actively hunting octopus’ (Hart et al., 2019; Leporati et

al., 2015).

4.2 Size distribution and growth

The mean mantle length (ML) of octopus from North Fremantle did not differ

significantly between the sexes. Although the overall modal ML over the four months was the

same for both males and females, the frequency of males in the larger size classes (>160 mm)

was higher than females. The largest individual collected over the sampling period was a male

caught in June, with a ML of 218 mm and MW of 691.62 g from the shallow region. This

octopus was larger than those in previous studies of O. tetricus from within the commercial

fishery where the largest males was 210 mm (Larsen, 2008). The maximum ML recorded for

females in Larson (2008) was 250 mm, substantially longer than at the maximum of 211 mm

recorded in the current study.

Small juveniles were first observed in the third haul in mid-April, with the smallest

individual (sex undetermined) being observed at the end of April. As Octopus aff. O. tetricus

has a two-month paralarval stage, these individuals were potentially hatched between January

Page 53: The ecology of Octopus aff. O. tetricus and prevalence of

53

and February. Octopus aff. O. tetricus have been identified as recruitment pulses every six

months, with a peak spawning event occurring in September (Joll, 1978; Larsen, 2008; Leporati

et al., 2015) and a secondary one in the summer months (Larsen, 2008; Leporati et al., 2015).

The two modal length classes that become prominent from the mid-April to the end of June

indicate that two cohorts are present in the population.

The monthly mean ML of male octopus increased steadily between March and June,

with the largest mean ML (152.27 mm) recorded in June. In contrast to males, the mean ML

of Females fluctuated between months, with octopus captured in March being significantly

larger than those in April and May. This coincides with the influx of juveniles on the third haul

in mid-April. Male and female octopuses typically exhibit similar growth rates until the onset

of maturation, however males reach maturity earlier, and females reach a significantly larger

maximum size (Leporati et al., 2015). Leporati et al. (2015) estimated the ML at 50% maturity

for male Octopus aff. O. terticus in the WA commercial fishery was 128 mm and reached at

243 d of age, 150 d faster and 54 mm smaller than the size at maturity for females (182 mm

ML at 379 d), (Leporati et al., 2015). These findings are consistent with octopus’ collected in

North Fremantle and demonstrate that two cohorts are likely to be present, one that spawned

the previous year in August or September of 2019, and the second from spawning in January

or February 2020.

During the current study, growth from the North Fremantle region did not appear to be

very rapid for either sex. This may be due to two main reasons: (1) larger individuals migrate

from the shallow waters and move outside the study region where the deepest line samples

were 16.79 m, and move into waters deeper than 20 m, as was found previously for octopus in

the WA fishery (Larsen, 2008; Leporati et al., 2015), and (2) growth slowed with the decreasing

SST from March to June, which is consistent with the Larson (2008), who examined samples

from the developing octopus fishery in the Perth region between June and August in water

<20 m deep and in deeper water (30‒40 m) from May to September. She found slower growth

between April and August. Temperature is a major influence on the growth of octopus,

especially in the early life stages. Modelled projections of temperature-dependent growth of O.

ocellatus and O. pallidus suggests that an increase of 1 °C (from 22 to 23°C) could lead to a

62.6% greater octopus body mass after 100 days at the higher temperature (André et al., 2009).

Temperature related growth is also intrinsically linked with an increase in metabolism, food

intake and food availability (Rigby & Sakurai, 2004). The food intake of O. briareus doubled

when water temperatures increased from 20 to 30 °C (Borer & Lane, 1971).

Page 54: The ecology of Octopus aff. O. tetricus and prevalence of

54

4.3 Maturity and reproduction

Most females collected during the sampling period were immature. The percentage of

immature females increased over the sampling period reaching a maximum in May, with more

than 90% of the females being immature in both depth regions. This is consistent with the

findings of Larson (2008), who found that virtually all females in waters < 20 m deep were

immature.

The data on gonad stages and gonadosomatic indices (GSIs) collected during this study

are not sufficient to define the spawning season for octopus in this region. However, in a

longer-term study of octopus based on samples from the commercial fishery in Cockburn

Sound, Larsen (2008) suggested that some spawning may occur in late autumn (April-May)

and early spring (September-October). Moreover, based on laboratory studies, Joll (1978)

proposed that the peak spawning for O. tetricus wild populations may occur in spring.

The maturity stages recorded for O. tetricus in Cockburn Sound were very similar to

those collected in North Fremantle on the same day in early April. North Fremantle had a higher

frequency of immature females than Cockburn Sound, however, North Fremantle also recorded

at least one female in each maturity stage, and no mature females were collected from the

Cockburn Sound site. Both sites had a high proportion of maturing males and recorded at least

one male in each maturity stage. Despite Cockburn Sound females having significantly larger

body sizes, the mean GSI was slightly lower than that for North Fremantle. The mean GSI

differed significantly for males and females in both North Fremantle and Cockburn Sound sites,

with males having a higher GSI for both areas. Results from this study provide strong evidence

of males maturing earlier than females in both sites, which is consistent with the findings from

previous studies (Joll, 1976; Leporati et al., 2015).

Only five mature females were recorded during the four months of this study which

may have been due to a migration out of the study area by maturing females or a change in

female catchability as they mature; mature females actively brood eggs, do not feed and are

not actively hunting at this time, and are thus not adventuring into traps (Leporati et al., 2015).

The migration of female octopuses to find mates and suitable dens for brooding eggs during

the breeding season has been suggested for several species, including the east coast O. tetricus

(Anderson, 1997), O. insularis in Brazil (Lima et al., 2014) and O. vulgaris in South Africa

(Oosthuizen & Smale, 2003).

Page 55: The ecology of Octopus aff. O. tetricus and prevalence of

55

Both sexes exhibited very similar monthly maturity proportions in the shallow and

deeper water of North Fremantle. The remarkable similarity of the maturity stages indicates

that the difference between the depth region is not large enough to demonstrate the migration

of mature individuals. The study design divided the two depth areas into ≤10 and >10 m to a

maximum of 20 m. In Larson (2008) the two depth regions examined were a shallow inshore

habitat (<20 m) and an offshore deeper habitat (30-40 m). The findings in this study are

consistent with Larson (2008) results for the inshore habitat, and the distinct lack of mature

females within the North Fremantle site demonstrates that the depth range is not adequate to

observe the size-related offshore migration that females undertake while maturing.

Mature males were observed in every haul in North Fremantle, with the highest number

being recorded in May. The presence of mature males for much of the year suggests that they

have adapted an opportunistic reproductive strategy. Octopus aff. O. tetricus mates all year

round, with two distinct pulses every six months (Larsen, 2008; Leporati et al., 2015) similar

reproductive patterns are seen with O. vulgaris (Katsanevakis & Verriopoulos, 2004).

Moreover, the storm incidents in May and June 2020 in the North Fremantle study area

may have accelerated the maturation of O. tetricus. Turbidity caused by high wind speeds and

wave action reduces light penetration within the water. In studies of the sexual maturation of

O. vulgaris it was found that light inhibits the release of the gonadotrophin hormone from the

optic gland (Mangold, 1983) which delays the onset of maturing gonads.

Males reach maturity earlier than females and are typically present within the

population throughout most of the year (Leporati et al., 2015). Female O. tetricus reach

maturity at ~12 months, as they have a semelparous life history and longevity of ~ 18 months

there is a ~6-month spawning window (Joll, 1976; Leporati et al., 2015). However, females

may mate at an earlier maturity stage and are able to store the spermatophores in the oviducts

for up to 16 weeks (Joll, 1976). Laboratory mating trials at the University of Western Australia

of O. tetricus revealed that multiple paternity is common among broods of O. tetricus (Morse,

2008). Theories behind female promiscuity are that either, females mate with multiple partners

to ensure phenotypic diversity among their offspring, or females may be sperm limited, and as

they only have a singular opportunity to reproduce, mating with multiple partners ensures the

fertilisation of larger broods (Joll, 1976; Morse, 2008). There is strong evidence of the latter,

with observations of female octopods (O. oliveri) initiating copulation with the first male they

are introduced to and become more selective as more mates are presented (Ylitalo et al., 2019).

Page 56: The ecology of Octopus aff. O. tetricus and prevalence of

56

Previous studies revealed that in captivity, female O. tetricus brood their eggs for ~1

month and then survive for 1–3 weeks after their eggs hatch (Joll, 1976). The spent individual

caught in April thus indicates a potential spawning event around February or March. The exact

length of the paralarval stage for O. tetricus has not been determined however it is likely to be

similar to O. vulgaris, which has a duration of 47-54 days at 21.1 °C and 30-35 days at 23 °C

(Iglesias et al., 2007). Furthermore, the number of very small juvenile (<80 mm ML) octopuses

observed was notably lower than in previous years (P. Jasinski, pers. comm.). The lower

number of small octopuses may be attributed to different water temperature conditions in the

previous year. Preliminary investigations of the interaction of SST and successful recruitment

were conducted for O. vulgaris in the Strait of Sicily, showed a significant correlation of

recruitment success and higher SST (Garofalo et al., 2010).

Due to the large proportion of small, immature octopus caught within the waters of

North Fremantle, it is likely that this site acts as a nursery area for juvenile octopuses. The

significant proportion of small octopus’ within the North Fremantle study site indicates an

ontogenetic migration of juveniles occurs to the shallower regions of North Fremantle,

potentially due to the differences in prey availability. It is well established that juvenile

octopuses frequently prey on crustaceans, and shift to teleosts and other cephalopods as body

size increases (Smith, 2003; Villanueva et al., 2017; Vincent et al., 1998). This is consistent

with the preliminary analysis of diets conducted in this study which shows that >40% of

octopuses had consumed at least one crustacean. These preliminary findings contrast with those

on O. tetricus diet from an abalone sea ranch in Flinders Bay, approximately 300 km south of

Cockburn Sound. Octopus consumed much greater amounts abalone, non-cephalopod

molluscs, on the sea ranch (Greenwell et al. 2019) than in North Fremantle. Future studies of

O. tetricus should examine the differences in the dietary composition of juvenile and adults,

including samples from deeper depth regions than those evaluated in this research.

4.4 Occurrence of nematodes

This study was the first to examine the presence of anisakid nematodes in O. tetricus.

Little information and very few publications are available on infestations of nematodes in

cephalopods, especially octopuses. Nearly 60% of octopuses sampled from North Fremantle

site were infested with nematodes. Modelling on nematode prevalence showed that the

probability of infection doubled for medium and larger octopus and that the incidence of

Page 57: The ecology of Octopus aff. O. tetricus and prevalence of

57

nematodes was much greater in March than June (about double). Temporal variation in

nematode prevalence does not appear to be related to octopus’ size, as the smallest mean sizes

was recorded in May. Potentially, the difference in prevalence due to size could be attributed

to changes in feeding regimes between the warmer and cooler months for O. tetricus, as

seasonal variation in diets has been identified for O. bimaculatus (Villegas et al., 2014) and E.

cirrhosa (Regueira et al., 2017). Additionally, the nematode reproductive cycle is known to

slow at lower temperatures, which may reduce the incidence of infection during the cooler

months (Fazio et al., 2012; Nagasawa et al., 1994).

The prevalence of nematodes in early April was almost three times in North Fremantle

than those from Cockburn Sound. The probability of infection by at least one nematode for

North Fremantle octopus was nearly ten times that of octopus in Cockburn Sound. The regional

differences of nematode infection between Cockburn Sound and North Fremantle may

demonstrate the existence of two separate populations of O. tetricus. Anisakis infections have

been applied as biological tags to assess sub-populations of pelagic and demersal octopus

species from the Mediterranean Sea (Mattiucci et al., 2015; Rello et al., 2009; Roumbedakis et

al., 2018a).

There is a high probability that O. tetricus ingested the nematodes from predating on

crustaceans, which are usually the first intermediate host of anisakid nematodes (Mattiucci et

al., 2018; Roumbedakis et al., 2018a). This coincides with the preliminary dietary analysis

performed on O. tetricus which demonstrated that a major component of the stomach contents

of the North Fremantle population was crustaceans. Predation on other cephalopods could also

result in the spread of nematodes throughout the population, as the incidence of cannibalism

detected in the diet analysis was quite high.

Infested crustaceans are likely to inhabit the same areas as O. tetricus due to the limited

migration of the species. At present, the extent of O. tetricus broad-scale movement is

unknown, however demersal octopuses tend to have restricted migrations (tens of kilometres)

(Semmens et al., 2007), and mark-recapture trials on similar species such as O. vulgaris have

demonstrated movement was limited to a home range of <1 km (Arechavala-Lopez et al., 2019;

Mereu et al., 2015). Minimal broad-scale movement accounts for the vast difference in

infection levels between the two sites, as they are unlikely to be sharing feeding grounds. Other

fish and cephalopods, particularly squid, can exhibit large broad-scale movement from feeding

Page 58: The ecology of Octopus aff. O. tetricus and prevalence of

58

and mating ground, therefore there is a higher likelihood that they will intermix with other

populations (Nadolna & Podolska, 2014; Semmens et al., 2007).

Octopus aff. O. tetricus is a suitable host due to their trophic position and morphological

adaptation. Coleoid cephalopods are susceptible to parasites due to the loss of an external shell,

which has allowed high-speed locomotion but increased their vulnerability, and the complex

structure of their skin, which is prone to lesioning (Kinne, 1990). Moreover, their function as

mesopredators and voracious feeding behaviours mean they easily accumulate multi-host

parasites through predation (Boyle & Rodhouse, 2005; Kinne, 1990; Roumbedakis et al.,

2018a).

Although it was not possible to identify the species of anisakid nematode, this would

be possible with molecular sequencing techniques. Nematodes in the genus Anisakis

(Anisakidae) are some of the most common and abundant parasites known to cause

pathological effects on fish and cephalopods, including ulceration and potential castration

(Abollo et al., 1998; Abollo et al., 2001). Examination of anisakid infection on a squid species

Illex coindetii, Todaropsis eblanae and Totarodes sagittatus found nematodes within the

gonads had caused partial destruction and alteration of the tissue, with partial inhibition of the

formation of reproductive cells. Within the North Fremantle site, only one octopus, a maturing

female, was found to have had a nematode infestation in the gonads. A small proportion of

octopus’ (3%) were found with nematodes encysted in the salivary glands. It not currently

known what impact a nematode infestation of the salivary glands has on the health of the

octopus; however, other parasitic infections have been known to cause severe damage to

salivary glands. Cestode (Prochristianella sp.) infection on the salivary glands of O. maya

caused deterioration of the tissues and the glands were dysfunctional, producing little to no

secretions (Guillén-Hernández et al., 2018). Consequently, the damage caused to organs by

parasites hinders the hosts capacity to feed and reproduce (Guillén-Hernández et al., 2018).

Multi-host parasites have a complex life history and are transmitted through the food

web, with cephalopods and fish acting as secondary intermediate hosts, and marine mammals

as definite hosts. A recent review of parasitism in cephalopods outlined several cuttlefish and

squid species known to have been infected by nematodes but does not identify any known

octopus species (Mattiucci et al., 2018; Roumbedakis et al., 2018a). Currently, there have only

been three other published incidents of larval nematode infection within octopuses, O. vulgaris

Page 59: The ecology of Octopus aff. O. tetricus and prevalence of

59

(Angelucci et al., 2011; Gestal et al., 1999), Eledone cirrhosa and E. moschata (Guardone et

al., 2020).

The presence of anisakid nematodes is also relevant to human health as the third stage

anisakids are aetiological agents of painful gastrointestinal diseases in humans if raw or

undercooked fish or cephalopods are consumed (Abollo et al., 2001). Anisakidosis, a parasitic

infection in humans, occurs when larvae penetrate the gastrointestinal mucosa, which can

trigger anaphylactic shock, and may require an endoscopic procedure or surgery (Abollo et al.,

2001; Shamsi, 2014). Additionally, anaphylactic reactions can still occur even a dead nematode

is ingested (Audicana et al., 2002). The first recorded incident of anisakidosis in Australia was

in 2011 and was attributed to the consumption of locally caught mackerel (Shamsi & Butcher,

2011). Due to the popularity of seafood consumption in Australia, in particular sushi and

sashimi, more undocumented incidents of anisakidosis in Australia are likely (Shamsi, 2014).

The extent of nematode infection on the arms of O. tetricus is not known and therefore the risk

of infection from consuming octopus from Perth waters cannot be assessed.

Due to time constraints of this thesis, and the challenges of nematode taxonomy, the

species of nematodes could not be determined. Future studies should perform molecular

identification on the nematodes and collect further samples from Cockburn Sound to identify

the true extent of the infestation within the commercial fishery.

Conclusions and Recommendations

The examination of the spatial and temporal distribution of O. tetricus in North

Fremantle, Western Australia, based on data collected between March and June 2020

demonstrates that the nearshore area of North Fremantle acts as a nursery area for juvenile and

maturing octopus. The changes in sex ratio between March to June were consistent with the

known composition of O. tetricus within the commercial fishery. Due to their early maturation,

mature males were observed throughout the entirety of the study. The distinct lack of mature

females within either depth region is consistent with the current biological hypothesis that O.

tetricus females have a size-related migration during maturation to find mates and suitable dens

for brooding.

The data on the mantle length distribution suggest a slower growth period for O. tetricus

between autumn and winter as SST is declining, and the presence of small juvenile octopus

Page 60: The ecology of Octopus aff. O. tetricus and prevalence of

60

from April onwards indicates a spawning event in January or February. This slower growth

during the time of lower SSTs is consistent with a previous study of the abundance and growth

of O. tetricus (Larsen, 2008), and with the findings for other octopus species. However, a more

detailed examination of the effect of environmental conditions on the population carried out

over an 18 to 24-month period is required more accurately patterns of abundance and growth

rates in the population.

Female mantle length was significantly longer in Cockburn Sound than North

Fremantle, and the overall proportion of maturing individuals was greater, which suggests that

Cockburn Sound may have spawned earlier or had a faster growth rate, potentially due to food

availability, differing environmental conditions. Moreover, some of the samples from

Cockburn Sound could have come from areas where females were maturing or brooding their

eggs. It may be worthwhile for future studies evaluate the potentially distinguishing growth

rates between the two sites and identifying where reproduction and egg brooding takes place,

with the utilisation of samples throughout the year to enable identification of seasonal

variances.

This study provides the first record of anisakid nematodes in the viscera of O. tetricus.

Nematodes occurred in the viscera of 60% of the octopus’ from North Fremantle, and

modelling indicated that the probability of North Fremantle octopus having at least one

nematode was almost ten times higher than the Cockburn Sound population. The full extent of

infection was not determined as the octopus’ arms were not examined during this study. While

the ingestion of seafood infected with anisakids by humans is known to cause anisakidosis, the

incidences of this in Australia are very rare, and it is likely to be low risk to the public at this

time. Molecular identification of the nematode species may help to determine what, if any,

potential risks may be associated with the consumption of O. tetricus by people. Examination

of any potential pathological effects on the octopus infested with nematodes would be

beneficial to monitor the health of the O. tetricus population.

The priority for future research is to perform a detailed analysis of the dietary

composition of O. tetricus and to assess any differences that may occur between different size

classes, sexes or depth regions. For a better understanding of the ingestion of nematodes by

octopuses, comparative assessment on diets between North Fremantle and Cockburn Sound

populations would be very valuable Furthermore, taking invertebrates from identified feeding

grounds of octopus may help identify the first intermediate nematode hosts.

Page 61: The ecology of Octopus aff. O. tetricus and prevalence of

61

References

Abollo, E., Gestal, C., López, A., González, A. F., Guerra, A., & Pascual, S. (1998). Squid as trophic

bridges for parasite flow within marine ecosystems: the case of Anisakis simplex (Nematoda:

Anisakidae), or when the wrong way can be right. South African journal of marine science,

20(1), 223-232. doi:10.2989/025776198784126575

Abollo, E., Gestal, C., & Pascual, S. (2001). Anisakis infestation in marine fish and cephalopods from

Galician waters: an updated perspective. Parasitology Research, 87(6), 492-499.

doi:10.1007/s004360100389

Alonso-Fernández, A., Otero, J., Bañón, R., Campelos, J. M., Santos, J., & Mucientes, G. (2017). Sex

ratio variation in an exploited population of common octopus: ontogenic shifts and spatio-

temporal dynamics. Hydrobiologia, 794(1), 1-16. doi:10.1007/s10750-016-3065-3

Alves, C., Modéran, J., Chichery, R., & Dickel, L. (2006). Plasticity of spatial learning strategies in

the common cuttlefish. Cognitive Processing, 7(S1), 111-111. doi:10.1007/s10339-006-0090-

6

Amor, M. D., Norman, M. D., Cameron, H. E., & Strugnell, J. M. (2014). Allopatric speciation within

a cryptic species complex of Australasian octopuses. PloS one, 9(6), e98982.

doi:10.1371/journal.pone.0098982

Anderson, R. C., Wood, J. B., & Byrne, R. A. (2002). Octopus Senescence: The Beginning of the

End. Journal of applied animal welfare science, 5(4), 275-283.

doi:10.1207/S15327604JAWS0504_02

Anderson, T. J. (1997). Habitat selection and shelter use by Octopus tetricus. Marine Ecology

Progress Series, 150(1/3), 137-148. doi:10.3354/meps150137

André, J., Eric, P. M. G., Semmens, J. M., Pecl, G. T., & Segawa, S. (2009). Effects of temperature on

energetics and the growth pattern of benthic octopuses. Marine ecology. Progress series

(Halstenbek), 374, 167-179. doi:10.3354/meps07736

Angelucci, G., Meloni, M., Merella, P., Sardu, F., Madeddu, S., Marrosu, R., . . . Salati, F. (2011).

Prevalence of Anisakis spp. and Hysterothylacium spp. larvae in teleosts and cephalopods

sampled from waters off Sardinia. Journal of food protection, 74(10), 1769-1775.

doi:10.4315/0362-028X.JFP-10-482

Arechavala-Lopez, P., Minguito-Frutos, M., Follana-Berná, G., & Palmer, M. (2019). Common

octopus settled in human-altered Mediterranean coastal waters: from individual home range to

population dynamics. ICES Journal of Marine Science, 76(2), 585-597.

doi:10.1093/icesjms/fsy014

Aronson, R. B. (1986). Life history and den ecology of Octopus briareus Robson in a marine lake.

Journal of Experimental Marine Biology and Ecology, 95(1), 37-56. doi:https://doi.org/10.1016/0022-0981(86)90086-9

Audicana, M. a. T., Ansotegui, I. J., de Corres, L. F., & Kennedy, M. W. (2002). Anisakis simplex:

dangerous — dead and alive? Trends in parasitology, 18(1), 20-25. doi:10.1016/s1471-

4922(01)02152-3

Belcari, P., Cuccu, D., González, M., Srairi, A., & Vidoris, P. (2002). Distribution and abundance of

Octopus vulgaris Cuvier, 1797 (Cephalopoda: Octopoda) in the Mediterranean Sea. In (Vol.

66, pp. 157-166): Consejo Superior de Investigaciones Científicas.

Borer, K. T., & Lane, C. E. (1971). Oxygen requirements of Octopus briareus Robson at different

temperatures and oxygen concentrations. Journal of Experimental Marine Biology and

Ecology, 7(3), 263-269. doi:https://doi.org/10.1016/0022-0981(71)90009-8 Boyle, P. R., & Boletzky, S. v. (1996). Cephalopod Populations: Definition and Dynamics.

Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences,

351(1343), 985-1002. doi:10.1098/rstb.1996.0089

Boyle, P. R., & Rodhouse, P. (2005). Cephalopods: ecology and fisheries (1st ed.). Oxford: Blackwell Science.

Page 62: The ecology of Octopus aff. O. tetricus and prevalence of

62

Brown, E. R., & Piscopo, S. (2013). Synaptic plasticity in cephalopods; more than just learning and memory? Invertebrate neuroscience : IN, 13(1), 35-44. doi:10.1007/s10158-013-0150-4

Bush, A. O., Lafferty, K. D., Lotz, J. M., & Shostak, A. W. (1997). Parasitology Meets Ecology on Its

Own Terms: Margolis et al. Revisited. The Journal of parasitology, 83(4), 575-583.

doi:10.2307/3284227

Caddy, J. F., & Rodhouse, P. G. (1998). Cephalopod and Groundfish Landings: Evidence for

Ecological Change in Global Fisheries? Reviews in Fish Biology and Fisheries, 8(4), 431-

444. doi:10.1023/A:1008807129366

Caldwell, R. L., Ross, R., Rodaniche, A., & Huffard, C. L. (2015). Behavior and Body Patterns of the

Larger Pacific Striped Octopus. PloS one, 10(8), e0134152.

doi:10.1371/journal.pone.0134152

Carreira, G. P., & Gonçalves, J. M. (2009). Catching Octopus vulgaris with traps in the Azores: first

trials employing Japanese baited pots in the Atlantic. Marine biodiversity records, 2.

doi:10.1017/S1755267209000499

Cigliano, J. A. (1993). Dominance and den use in Octopus bimaculoides. Animal Behaviour, 46(4), 677-684. doi:10.1006/anbe.1993.1244

Cole, P. D., & Adamo, S. A. (2005). Cuttlefish (Sepia officinalis: Cephalopoda) hunting behavior and

associative learning. Animal Cognition, 8(1), 27-30. doi:10.1007/s10071-004-0228-9

Coleman, F., & Williams, S. (2002). Overexploiting marine ecosystem engineers: Potential

consequences for biodiversity. Trends in Ecology & Evolution, 17, 40-44. doi:10.1016/S0169-

5347(01)02330-8

DoF. (2015). Western Australian Octopus Fishery Interim Management Plan. Retrieved from

https://www.slp.wa.gov.au/statutes/subsiduary.nsf/FisheriesT?openpage

Doubleday, Z. A., Prowse, T. A. A., Arkhipkin, A., Pierce, G. J., Semmens, J., Steer, M., . . . Gillanders, B. M. (2016). Global proliferation of cephalopods. Current Biology, 26(10), R406-R407. doi:https://doi.org/10.1016/j.cub.2016.04.002

Duarte, J. A., Hernández-Flores, A., Salas, S., & Seijo, J. C. (2018). Is it sustainable fishing for

Octopus maya Voss and Solis, 1966, during the breeding season using a bait-based fishing

technique? Fisheries Research, 199, 119-126. doi:10.1016/j.fishres.2017.11.020

FAO. (2020). The State of World Fisheries and Aquaculture 2020. Sustainability in action.

doi:https://doi.org/10.4060/ca9229en

Fazio, G., Sasal, P., Mouahid, G., Lecomte-Finiger, R., & Moné, H. (2012). Swim Bladder

Nematodes (Anguillicoloides crassus) Disturb Silvering In European Eels (Anguilla anguilla).

The Journal of parasitology, 98(4), 695-705. doi:10.1645/GE-2700.1

Furlani, D., Hobday, A. J., Ling, S., Dowdney, J., Bulman, C., Sporcic, M., & Fuller, M. (2007).

Ecological Risk Assessment for the Effects of Fishing: Southern Squid Jig Sub-fishery.

Canberra

Garofalo, G., Ceriola, L., Gristina, M., Fiorentino, F., & Pace, R. (2010). Nurseries, spawning

grounds and recruitment of Octopus vulgaris in the Strait of Sicily, central Mediterranean

Sea. ICES Journal of Marine Science, 67(7), 1363-1371. doi:10.1093/icesjms/fsq101

Gestal, C., Abollo, E., Arias, C., & Pascual, S. (1999). Larval Nematodes (Spiruroidea:

Cystidicolidae) in Octopus vulgaris (Mollusca: Cephalopoda: Octopodidae) from the

Northeastern Atlantic Ocean. The Journal of parasitology, 85(3), 508-511.

doi:10.2307/3285787

Greenwell, C. N., Loneragan, N. R., Admiraal, R., Tweedley, J. R., & Wall, M. (2019a). Octopus as

predators of abalone on a sea ranch. In (Vol. 26, pp. 108-118): Wiley Subscription Services,

Inc.

Greenwell, C. N., Loneragan, N. R., Tweedley, J. R., & Wall, M. (2019b). Diet and trophic role of

octopus on an abalone sea ranch. Fisheries management and ecology, 26(6), 638-649.

doi:10.1111/fme.12381 Guard, M., & Mgaya, Y. D. (2002). The Artisanal Fishery for Octopus cyanea Gray in Tanzania.

AMBIO, 31(7), 528-536. doi:10.1579/0044-7447-31.7.528

Guardone, L., Bilska-Zając, E., Giusti, A., Malandra, R., Cencek, T., & Armani, A. (2020). Larval

ascaridoid nematodes in horned and musky octopus (Eledone cirrhosa and E. moschata) and

longfin inshore squid (Doryteuthis pealeii): Safety and quality implications for cephalopod

Page 63: The ecology of Octopus aff. O. tetricus and prevalence of

63

products sold as fresh on the Italian market. International journal of food microbiology, 333, 108812-108812. doi:10.1016/j.ijfoodmicro.2020.108812

Guerra, Á., Hernández-Urcera, J., Garci, M. E., Sestelo, M., Regueira, M., González, Á. F., . . .

Morales-Nin, B. (2015). Spawning habitat selection by Octopus vulgaris: New insights for a

more effective management of this resource. Fisheries Research, 167, 313-322.

doi:10.1016/j.fishres.2015.03.011

Guerrero-Kommritz, J., Guerrero-Kommritz, J., Camelo-Guarin, S., & Camelo-Guarin, S. (2016).

Two new octopod species (Mollusca: Cephalopoda) from the southern Caribbean. Marine

biodiversity, 46(3), 589-602. doi:10.1007/s12526-015-0406-9

Guerrero-Kommritz, J., & Rodriguez-Bermudez, A. (2018). Soft-bottom octopods (Cephalopoda:

Octopodidae) of the southern Caribbean with the description of a new species of

Macrotritopus. Marine biodiversity, 49(3), 1197-1215. doi:10.1007/s12526-018-0903-8

Guillén-Hernández, S., González-Salas, C., Pech-Puch, D., & Villegas-Hernández, H. (2018).

Octopus maya parasites off the Yucatán Peninsula, Mexico. II. Salivary gland damage by

cestodes. Diseases of aquatic organisms, 130(1), 45-50. doi:10.3354/dao03252

Hanlon, R. T., Watson, A. C., & Barbosa, A. (2010). A "mimic octopus" in the Atlantic: flatfish

mimicry and camouflage by Macrotritopus defilippi. The Biological Bulletin, 218. Retrieved

from

https://link.gale.com/apps/doc/A222229086/ITOF?u=murdoch&sid=ITOF&xid=f8c6b231

Hanlon, R. T. a., & Messenger, J. B. a. (2018). Cephalopod behaviour (Second ed.). New

York;Cambridge, United Kingdom;: Cambridge University Press.

Hart, A. M., Leporati, S. C., Marriott, R. J., & Murphy, D. (2016). Innovative Development of the

Octopus (cf) tetricus fishery in Western Australia FRDC Project No 2010/200. Western

Australia

Hart, A. M., Murphy, D., Hesp, S. A., & Leporati, S. (2019). Biomass estimates and harvest strategies

for the Western Australian Octopus aff. tetricus fishery. ICES Journal of Marine Science,

76(7), 2205-2217. doi:10.1093/icesjms/fsz146

Hartwick, E. B., Ambrose, R. F., & Robinson, S. M. C. (1984). Den utilization and the movements of

tagged Octopus dofleini. Marine Behaviour and Physiology, 11(2), 95-110.

doi:10.1080/10236248409387038

Heery, E. C., Olsen, A. Y., Feist, B. E., & Sebens, K. P. (2018). Urbanization-related distribution

patterns and habitat-use by the marine mesopredator, giant Pacific octopus (Enteroctopus

dofleini). Urban Ecosystems, 21(4), 707-719. doi:10.1007/s11252-018-0742-1

Hernández-Garcı́a, V., Hernández-López, J. L., & Castro, J. J. (1998). The octopus (Octopus vulgaris)

in the small-scale trap fishery off the Canary Islands (Central-East Atlantic). Fisheries Research, 35(3), 183-189. doi:10.1016/S0165-7836(98)00080-0

Iglesias, J., Sánchez, F. J., Bersano, J. G. F., Carrasco, J. F., Dhont, J., Fuentes, L., . . . Villanueva, R.

(2007). Rearing of Octopus vulgaris paralarvae: Present status, bottlenecks and trends. Aquaculture, 266(1-4), 1-15. doi:10.1016/j.aquaculture.2007.02.019

Iribarne, O. O. (1990). Use of shelter by the small Patagonian octopus Octopus tehuelchus:

availability, selection and effects on fecundity. Marine Ecology Progress Series, 66(3), 251-

258. doi:10.3354/meps066251

Joll, L. M. (1976). Mating, egg-laying and hatching of Octopus tetricus (Mollusca: Cephalopoda) in

the laboratory. Marine Biology, 36(4), 327-333. doi:10.1007/BF00389194

Joll, L. M. (1977). The predation of pot-caught western rock lobster (Panulirus longipescygnus) by octopus Perth [W.A.]: Dept. of Fisheries and Wildlife.

Joll, L. M. (1978). Observations on the Embryonic Development of Octopus tetricus (Mollusca :

Cephalopoda). Marine and freshwater research, 29(1), 19. doi:10.1071/MF9780019

Kailola, P., Williams, M., P.Stewart, R.Reichelt, A.McNee, & Grieve, C. (1993). Australian Fisheries

Resources.

Katsanevakis, S., & Verriopoulos, G. (2004). Abundance of Octopus vulgaris on soft sediment. In

(Vol. 68, pp. 553-560): Consejo Superior de Investigaciones Científicas.

Kinne, O. (1990). Diseases of Marine Animals. Vol. III: Introduction, Cephalopoda, Annelida,

Crustacea, Chaetognatha, Echinodermata, Urochordata. In: Biologische Anstalt Helgoland,

Hamburg.

Page 64: The ecology of Octopus aff. O. tetricus and prevalence of

64

Larsen, R. J. (2008). The Western Australian developmental octopus fishery: species composition and

aspects of morphology and biology. (Unpublished Honours thesis). Murdoch University,

Western Australia,

Leite, T. S., Haimovici, M., Mather, J., & Oliveira, J. E. L. (2009). Habitat, distribution, and

abundance of the commercial octopus ( Octopus insularis) in a tropical oceanic island, Brazil:

Information for management of an artisanal fishery inside a marine protected area. Fisheries

Research, 98(1), 85-91. doi:10.1016/j.fishres.2009.04.001

Leporati, S. C., & Hart, A. M. (2015). Stylet weight as a proxy for age in a merobenthic octopus population. Fisheries Research, 161, 235-243. doi:10.1016/j.fishres.2014.08.001

Leporati, S. C., Hart, A. M., Larsen, R., Franken, L. E., & De Graaf, M. (2015). Octopus life history

relative to age, in a multi-geared developmental fishery. Fisheries Research, 165, 28-41.

doi:https://doi.org/10.1016/j.fishres.2014.12.017

Leporati, S. C., Ziegler, P. E., & Semmens, J. M. (2009). Assessing the stock status of holobenthic

octopus fisheries: is catch per unit effort sufficient? ICES Journal of Marine Science, 66(3),

478-487. doi:10.1093/icesjms/fsn224

Lima, F. D., Leite, T. S., Haimovici, M., Nóbrega, M. F., & Oliveira, J. E. L. (2014). Population

structure and reproductive dynamics of Octopus insularis (Cephalopoda: Octopodidae) in a

coastal reef environment along northeastern Brazil. Fisheries Research, 152(1), 86-92.

doi:10.1016/j.fishres.2013.08.009

Mangold, K. (1983). Octopus vulgaris. 335-364 in PR Boyle, ed Cephalopod life cycles, Vol. In: I

Academic press, London.

Marshall, N. J., & Messenger, J. B. (1996). Colour-blind camouflage. Nature, 382(6590), 408-409.

doi:10.1038/382408b0

Mather, J. A. (1982). Choice and competition: Their effects on occupancy of Shell Homes by Octopus

joubini. Marine Behaviour and Physiology, 8(4), 285-293. doi:10.1080/10236248209387025

Mather, J. A. (1994). ‘Home’ choice and modification by juvenile Octopus vulgaris (Mollusca:

Cephalopoda): specialized intelligence and tool use? Journal of Zoology, 233(3), 359-368.

doi:10.1111/j.1469-7998.1994.tb05270.x

Mather, J. A., & Kuba, M. J. (2013). The cephalopod specialties: complex nervous system, learning,

and cognition. Canadian Journal of Zoology, 91(6), 431-449.

Mather, J. A., & O'Dor, R. K. (1991). Foraging Strategies and Predation Risk Shape the Natural

History of Juvenile Octopus vulgaris. Bulletin of Marine Science, 49(1), 256-269. Mattiucci, S., Cimmaruta, R., Cipriani, P., Abaunza, P., Bellisario, B., & Nascetti, G. (2015).

Integrating Anisakis spp. parasites data and host genetic structure in the frame of a holistic

approach for stock identification of selected Mediterranean Sea fish species. Parasitology,

142(1), 90-108. doi:10.1017/S0031182014001103

Mattiucci, S., Cipriani, P., Levsen, A., Paoletti, M., & Nascetti, G. (2018). Molecular Epidemiology of

Anisakis and Anisakiasis: An Ecological and Evolutionary Road Map (Vol. 99). England:

Elsevier.

Mereu, M., Agus, B., Cannas, R., Cau, A., Coluccia, E., & Cuccu, D. (2015). Mark-recapture

investigation on Octopus vulgaris specimens in an area of the central western Mediterranean

Sea. Journal of the Marine Biological Association of the United Kingdom, 95(1), 131-138.

doi:http://dx.doi.org/10.1017/S002531541400112X

Moody, M. F., & Parriss, J. R. (1960). The Visual System of Octopus: (2) Discrimination of Polarized

Light by Octopus. Nature, 186(4728), 839-840. doi:10.1038/186839a0 Morse, P. (2008). Female Mating Preference Polyandry and Paternity Bias in Octopus tetricus.

Nadolna, K., & Podolska, M. (2014). Anisakid larvae in the liver of cod (Gadus morhua) L. from the

southern Baltic Sea. Journal of Helminthology, 88(2), 237-246.

doi:10.1017/S0022149X13000096

Nagasawa, K., Kim, Y., & Hirose, H. (1994). Anguillicola crassus and A. globiceps (Nematoda:

Dracunculoidea) parasitic in the swimbladder of eels (Anguilla japonica and A. anguilla) in

East Asia: a review. Folia Parasitologica, 41, 127-137.

Narvarte, M., González, R. A., Storero, L., & Fernández, M. (2013). Effects of competition and egg

predation on shelter use by Octopus tehuelchus females. In (Vol. 482, pp. 141-151): Inter-

Research.

Page 65: The ecology of Octopus aff. O. tetricus and prevalence of

65

Norman, M. D., Reid, A. L., & ProQuest. (2000). A guide to squid, cuttlefish and octopuses of Australasia. Collingwood, Vic;Moorabbin, Vic;: CSIRO Pub.

Nottage, J. D., Nottage, J. D., West, R. J., West, R. J., Montgomery, S. S., Montgomery, S. S., . . .

Graham, K. (2007). Cephalopod diversity in commercial fisheries landings of New South

Wales, Australia. Reviews in Fish Biology and Fisheries, 17(2), 271-281.

doi:10.1007/s11160-006-9032-8

Nowara, G. B., & Walker, T. I. (1998). Effects of time of solar day, jigging method and jigging depth

on catch rates and size of Gould's squid, Nototodarus gouldi (McCoy), in southeastern

Australian waters. Fisheries Research, 34(3), 279-288. doi:10.1016/S0165-7836(97)00095-7

O'Brien, C. E., Roumbedakis, K., & Winkelmann, I. E. (2018). The Current State of Cephalopod

Science and Perspectives on the Most Critical Challenges Ahead From Three Early-Career

Researchers. Frontiers in Physiology, 9, 700-700. doi:10.3389/fphys.2018.00700

Oosthuizen, A., & Smale, M. J. (2003). Population biology of Octopus vulgaris on the temperate

south-eastern coast of South Africa. In (Vol. 83, pp. 535-541). Cambridge: Cambridge

University Press.

Pierce, G. J., Valavanis, V. D., Guerra, A., Jereb, P., Orsi-Relini, L., Bellido, J. M., . . . Zuur, A. F.

(2008). A review of cephalopod–environment interactions in European Seas. In (Vol. 612, pp.

49-70). Dordrecht: Springer Netherlands.

Pignatelli, V., Temple, S. E., Chiou, T.-H., Roberts, N. W., Collin, S. P., & Marshall, N. J. (2011).

Behavioural relevance of polarization sensitivity as a target detection mechanism in

cephalopods and fishes. Philosophical Transactions: Biological Sciences, 366(1565), 734-

741. doi:10.1098/rstb.2010.0204

Pratasik, S. B., Marsoedi, Arfiati, D., & Setyohadi, D. (2017). Egg placement habitat selection of

cuttlefish, Sepia latimanus (Sepiidae, Cephalopoda, Mollusca) in North Sulawesi waters,

Indonesia. In (Vol. 10, pp. 1514-1523). Cluj-Napoca: Bioflux SRL.

Quetglas, A., Carbonell, A., & Sánchez, P. (2000). Demersal Continental Shelf and Upper Slope

Cephalopod Assemblages from the Balearic Sea (North-Western Mediterranean). Biological

Aspects of Some Deep-Sea Species. Estuarine, Coastal and Shelf Science, 50(6), 739-749.

doi:10.1006/ecss.1999.0603

R Core Team. (2020). R: A Language and Environment for Statistical Computing. Vienna, Austria: R

Foundation for Statistical Computing.

Ramos, J. E., Pecl, G. T., Moltschaniwskyj, N. A., Strugnell, J. M., León, R. I., & Semmens, J. M.

(2014). Body size, growth and life span: implications for the polewards range shift of Octopus

tetricus in south-eastern Australia. PloS one, 9(8), e103480.

doi:10.1371/journal.pone.0103480

Regueira, M., González, A. F., & Guerra, A. (2014). Habitat selection and population spreading of the

horned octopus Eledone cirrhosa (Lamarck, 1798) in Galician waters (NW Atlantic). Fisheries Research, 152, 66-73. doi:https://doi.org/10.1016/j.fishres.2013.09.003

Regueira, M., Regueira, M., Guerra, Á., Guerra, Á., Fernández-Jardón, C. M., Fernández-Jardón, C.

M., . . . González, Á. F. (2017). Diet of the horned octopus Eledone cirrhosa in Atlantic

Iberian waters: ontogenetic and environmental factors affecting prey ingestion. Hydrobiologia, 785(1), 159-171. doi:10.1007/s10750-016-2916-2

Rello, F. J., Adroher, F. J., Benítez, R., & Valero, A. (2009). The fishing area as a possible indicator

of the infection by anisakids in anchovies (Engraulis encrasicolus) from southwestern

Europe. International journal of food microbiology, 129(3), 277-281. Richter, J. N., Hochner, B., & Kuba, M. J. (2016). Pull or Push? Octopuses Solve a Puzzle Problem.

PloS one, 11(3), e0152048. doi:10.1371/journal.pone.0152048

Rigby, R. P., & Sakurai, Y. (2004). Temperature and Feeding Related Growth Efficiency of Immature

Octopuses Enteroctopus dofleini. Suisanzoshoku, 52, 29-36.

doi:10.11233/aquaculturesci1953.52.29

Robin, J.-P., Roberts, M., Zeidberg, L., Bloor, I., Rodriguez, A., Briceño, F., . . . Mather, J. (2014).

Transitions During Cephalopod Life History: The Role of Habitat, Environment, Functional

Morphology and Behaviour. In E. A. G. Vidal (Ed.), Advances in marine biology (Vol. 67,

pp. 361-437): Academic Press.

Page 66: The ecology of Octopus aff. O. tetricus and prevalence of

66

Rodhouse, P. G. (2010). Effects of environmental variability and change on cephalopod populations:

an Introduction to the CIAC '09 Symposium special issue. ICES Journal of Marine Science,

67(7), 1311-1313. doi:10.1093/icesjms/fsq096 Rodhouse, P. G., Nigmatullin, C. M., & unav. (1997). Role as consumers. Philosophical transactions.

Biological sciences, 351(1343), 1003-1022. doi:10.1098/rstb.1996.0090

Rosa, R., Pissarra, V., Borges, F. O., Xavier, J., Gleadall, I. G., Golikov, A., . . . Villanueva, R.

(2019). Global Patterns of Species Richness in Coastal Cephalopods. Frontiers in Marine

Science, 6. doi:10.3389/fmars.2019.00469

Roumbedakis, K., Drábková, M., Tyml, T., Di Cristo, C., & Lawrence Berkeley National Lab, B. C.

A. (2018a). A perspective around cephalopods and their parasites, and suggestions on how to

increase knowledge in the field. Frontiers in Physiology, 9, 1573.

doi:10.3389/fphys.2018.01573

Roumbedakis, K., Roumbedakis, K., Mascaró, M., Mascaró, M., Martins, M. L., Martins, M. L., . . .

Pascual, C. (2018b). Health status of post-spawning Octopus maya (Cephalopoda:

Octopodidae) females from Yucatan Peninsula, Mexico. Hydrobiologia, 808(1), 23-34.

doi:10.1007/s10750-017-3340-y

Sauer, W. H. H., Gleadall, I. G., Downey-Breedt, N., Doubleday, Z., Gillespie, G., Haimovici, M., . . .

Pecl, G. (2019). World Octopus Fisheries. Reviews in fisheries science & aquaculture, 1-151.

doi:10.1080/23308249.2019.1680603

Scatà, G., Jozet-Alves, C., Thomasse, C., Josef, N., & Shashar, N. (2016). Spatial learning in the

cuttlefish Sepia officinalis: preference for vertical over horizontal information. The Journal of

experimental biology, 219(Pt 18), 2928-2933. doi:10.1242/jeb.129080

Scheel, D., Chancellor, S., Hing, M., Lawrence, M., Linquist, S., & Godfrey-Smith, P. (2017). A

second site occupied by Octopus tetricus at high densities, with notes on their ecology and

behavior. Marine and Freshwater Behaviour and Physiology, 50(4), 285-291.

doi:10.1080/10236244.2017.1369851

Semmens, J. M., Pecl, G. T., Gillanders, B. M., Waluda, C. M., Shea, E. K., Jouffre, D., . . . Shaw, P.

W. (2007). Approaches to resolving cephalopod movement and migration patterns. Reviews in Fish Biology and Fisheries, 17(2), 401-423. doi:10.1007/s11160-007-9048-8

Serb, J. M., & Eernisse, D. J. (2008). Charting Evolution's Trajectory: Using Molluscan Eye Diversity

to Understand Parallel and Convergent Evolution. Evolution: Education and Outreach, 1(4),

439. doi:10.1007/s12052-008-0084-1

Shamsi, S. (2014). Recent advances in our knowledge of Australian anisakid nematodes. International

journal for parasitology. Parasites and wildlife, 3(2), 178-187.

doi:10.1016/j.ijppaw.2014.04.001

Shamsi, S., & Butcher, A. R. (2011). First report of human anisakidosis in Australia. Medical journal

of Australia, 194(4), 199-200. doi:10.5694/j.1326-5377.2011.tb03772.x

Shashar, N., Hagan, R., Boal, J. G., & Hanlon, R. T. (2000). Cuttlefish use polarization sensitivity in

predation on silvery fish. Vision Research, 40(1), 71-75. doi:https://doi.org/10.1016/S0042-

6989(99)00158-3

Smith, C. D. (2003). Diet of Octopus vulgaris in False Bay, South Africa. Marine Biology, 143(6),

1127-1133. doi:10.1007/s00227-003-1144-2

Smith, C. D., Groeneveld, J. C., & Maharaj, G. (2006). The life history of the giant octopus Octopus

magnificus in South African waters. African Journal of Marine Science, 28(3-4), 561-568.

doi:10.2989/18142320609504206

Sobrino, I., Silva, L., Bellido, J. M., & Ramos, F. (2002). Rainfall, river discharges and sea

temperature as factors affecting abundance of two coastal benthic cephalopod species in the

Gulf of Cadiz (SW Spain). Bulletin of Marine Science, 71(2), 851-865.

Thrush, S. F., Ellingsen, K. E., & Davis, K. (2016). Implications of fisheries impacts to seabed

biodiversity and ecosystem-based management. ICES Journal of Marine Science,

73(suppl_1), i44-i50. doi:10.1093/icesjms/fsv114

Tian, Y. (2009). Interannual–interdecadal variations of spear squid Loligo bleekeri abundance in the

southwestern Japan Sea during 1975–2006: Impact of the trawl fishing and recommendations

for management under the different climate regimes. Fisheries Research, 100(1), 78-85.

doi:10.1016/j.fishres.2009.06.005

Page 67: The ecology of Octopus aff. O. tetricus and prevalence of

67

Tricarico, E., Borrelli, L., Gherardi, F., & Fiorito, G. (2011). I know my neighbour: individual recognition in Octopus vulgaris. PloS one, 6(4), e18710. doi:10.1371/journal.pone.0018710

Trinder, D. M., & McKinley, C. P. (2003). Interactions between fur seals and a squid jig fishery in

southern Australia. Marine & Freshwater Research, 54(8), 979-984. doi:10.1071/MF03030

Uriarte, I., Martínez‐Montaño, E., Espinoza, V., Rosas, C., Hernández, J., & Farías, A. (2016). Effect

of temperature increase on the embryonic development of Patagonian red octopus

Enteroctopus megalocyathus in controlled culture. Aquaculture Research, 47(8), 2582-2593.

doi:10.1111/are.12707 Vargas-Yáñez, M., Moya, F., García-Martínez, M., Rey, J., González, M., & Zunino, P. (2009).

Relationships between Octopus vulgaris landings and environmental factors in the northern

Alboran Sea (Southwestern Mediterranean). Fisheries Research, 99(3), 159-167.

doi:https://doi.org/10.1016/j.fishres.2009.05.013

Villanueva, R., Perricone, V., & Fiorito, G. (2017). Cephalopods as Predators: A Short Journey

among Behavioral Flexibilities, Adaptions, and Feeding Habits. Frontiers in Physiology, 8,

598. doi:10.3389/fphys.2017.00598

Villanueva, R., Vidal, E. A. G., Fernández-Álvarez, F. Á., & Nabhitabhata, J. (2016). Early Mode of

Life and Hatchling Size in Cephalopod Molluscs: Influence on the Species Distributional

Ranges. PloS one, 11(11), e0165334-e0165334. doi:10.1371/journal.pone.0165334

Villegas, E. J. A., Ceballos-Vazquez, B. P., Markaida, U., Abitia-Cardenas, A., Medina-Lopez, M. A.,

& Arellano-Martinez, M. (2014). Diet of Octopus bimaculatus verril, 1883 (Cephalopoda:

Octopodidae) in Bahia de Los Angeles, Gulf of California. Journal of Shellfish Research,

33(1), 305-314.

Vincent, T. L. S., Scheel, D., & Hough, K. R. (1998). Some Aspects of Diet and Foraging Behavior of

Octopus dofleini Wülker, 1910 in its Northernmost Range. In (Vol. 19, pp. 13-29). Oxford,

UK: Blackwell Publishing Ltd.

Watson, R., & Pauly, D. (2001). Systematic distortions in world fisheries catch trends. Nature

(London), 414(6863), 534-536. doi:10.1038/35107050

Wodinsky, J. (1977). Hormonal Inhibition of Feeding and Death in Octopus: Control by Optic Gland

Secretion. Science (American Association for the Advancement of Science), 198(4320), 948-

951. doi:10.1126/science.198.4320.948

Yick, J. L., Barnett, A., & Tracey, S. R. (2012). The trophic ecology of two abundant mesopredators

in south-east coastal waters of Tasmania, Australia. Marine Biology, 159(6), 1183-1196.

doi:10.1007/s00227-012-1899-4

Ylitalo, H., Oliver, T. A., Fernandez-Silva, I., Wood, J. B., & Toonen, R. J. (2019). A behavioral and

genetic study of multiple paternity in a polygamous marine invertebrate, Octopus oliveri.

PeerJ (San Francisco, CA), 7, e6927-e6927. doi:10.7717/peerj.6927