21.01.2015 Views

The emerging field of electrosensory and semiochemical shark ...

The emerging field of electrosensory and semiochemical shark ...

The emerging field of electrosensory and semiochemical shark ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Ocean & Coastal Management xxx (2012) 1e10<br />

Contents lists available at SciVerse ScienceDirect<br />

Ocean & Coastal Management<br />

journal homepage: www.elsevier.com/locate/ocecoaman<br />

<strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents:<br />

Mechanisms <strong>of</strong> detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions<br />

Craig P. O’Connell a, *, Eric M. Stroud b , Pingguo He a<br />

a School <strong>of</strong> Marine Science <strong>and</strong> Technology, University <strong>of</strong> Massachusetts Dartmouth, New Bedford, MA 02740, USA<br />

b Shark Defense Technologies, LLC, P.O. Box 2593, Oak Ridge, NJ 07438, USA<br />

article<br />

Article history:<br />

Available online xxx<br />

info<br />

abstract<br />

Since the sinking <strong>of</strong> the USS Indianapolis (CA-35) <strong>and</strong> associated <strong>shark</strong> attacks in 1945, the quest to find<br />

an effective <strong>shark</strong> repellent has been endless. Early efforts were focused on finding a <strong>shark</strong> repellent<br />

which would minimize the probability <strong>of</strong> a <strong>shark</strong> attack. However, studies illustrate that <strong>shark</strong> populations<br />

are drastically declining which has led to calls for effective management policies <strong>and</strong> practices to<br />

reduce both directed catch <strong>and</strong> bycatch <strong>of</strong> various <strong>shark</strong> species. With increased need for <strong>shark</strong><br />

conservation, the focus has shifted to protecting <strong>shark</strong>s from harmful anthropogenic pressures, such as<br />

fishing gear <strong>and</strong> beach nets. Current <strong>shark</strong> repellent technologies which aim to minimize elasmobranch<br />

mortality in fishing gears include: permanent magnets, electropositive metal (EPM) alloys, <strong>and</strong> <strong>semiochemical</strong>s.<br />

This paper will review present <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents, the<br />

mechanisms <strong>of</strong> elasmobranch (e.g. <strong>shark</strong>, skate <strong>and</strong> ray) detection <strong>and</strong> repellency, species-specificity in<br />

elasmobranch response to the stimuli, <strong>and</strong> environmental <strong>and</strong> biological conditions which may influence<br />

repellent success. Future research to enhance our knowledge on <strong>electrosensory</strong> repellents <strong>and</strong> to<br />

improve the success <strong>of</strong> repellent implementation <strong>and</strong> application will be discussed.<br />

Ó 2012 Elsevier Ltd. All rights reserved.<br />

1. Introduction<br />

Fisheries <strong>and</strong> beach nets are major contributors to elasmobranch<br />

decline (Bonfil, 1994; Baum et al., 2003; Baum <strong>and</strong> Myers,<br />

2004; Cliff <strong>and</strong> Dudley, 1992; Dudley, 1997; Manire <strong>and</strong> Gruber,<br />

1990; Shepherd <strong>and</strong> Myers, 2005) <strong>and</strong> is an issue that urgently<br />

needs to be addressed. Elasmobranchs are <strong>of</strong>ten inadvertently<br />

caught (bycatch) on fishing gears that results in substantial adverse<br />

economic (Gilman et al., 2007) <strong>and</strong> ecological (Baum et al., 2003)<br />

effects. During fishing operations, fishers rarely retain <strong>shark</strong>s (dead<br />

or alive) due to little market for <strong>shark</strong> meat, trip limits, <strong>and</strong>/or<br />

management plans which prohibit the retention <strong>of</strong> certain species<br />

(Lewison et al., 2004; Harrington et al., 2005). <strong>The</strong>se practices lead<br />

to high incidences <strong>of</strong> discards <strong>and</strong> associated mortality, <strong>and</strong> due to<br />

the K-selected nature <strong>of</strong> elasmobranchs (e.g. slow growth rate, late<br />

maturity, <strong>and</strong> low fecundity), the likelihood <strong>of</strong> <strong>shark</strong> population<br />

rebound is minimal if the present mortality rates continue (Pratt<br />

<strong>and</strong> Casey, 1990). From a fisherman’s st<strong>and</strong>point, interactions<br />

with <strong>shark</strong>s during fishing operations cause considerable issues<br />

* Corresponding author.<br />

E-mail address: coconnell2@umassd.edu (C.P. O’Connell).<br />

including damage <strong>and</strong> loss <strong>of</strong> gear, reduced capture <strong>of</strong> marketable<br />

species, depredation, risk <strong>of</strong> injury to the mates, <strong>and</strong> loss <strong>of</strong> valuable<br />

time untangling <strong>and</strong> reconstructing the fishing gear (Gilman et al.,<br />

2007).<br />

Approaches to reducing anthropogenic-associated elasmobranch<br />

mortality include the use <strong>of</strong> permanent magnets, electropositive<br />

metal (EPM) alloys, <strong>and</strong>/or <strong>semiochemical</strong>s as repellents.<br />

Several studies have been conducted that aim to evaluate the<br />

effectiveness <strong>of</strong> these repellents on different elasmobranch species<br />

(Brill et al., 2009; Rigg et al., 2009; Tallack <strong>and</strong> M<strong>and</strong>elman, 2009;<br />

O’Connell, 2008; O’Connell et al., 2010; Stroud et al. submitted for<br />

publication). This paper will review present <strong>shark</strong> repellents,<br />

elasmobranch (e.g. <strong>shark</strong>, skate <strong>and</strong> ray) detection mechanisms,<br />

species-specificity in elasmobranch response to the stimuli, <strong>and</strong><br />

environmental <strong>and</strong>/or biological conditions which may influence<br />

repellent success. Future research to enhance our knowledge on<br />

<strong>electrosensory</strong> repellents <strong>and</strong> to improve the success <strong>of</strong> repellent<br />

implementation <strong>and</strong> application will also be discussed.<br />

Permanent magnets <strong>and</strong> EPMs are hypothesized to overwhelm<br />

the acute <strong>electrosensory</strong> system, known as the ampullae <strong>of</strong> Lorenzini,<br />

<strong>of</strong> an interacting elasmobranch thus causing aversion<br />

behaviors (WWF, 2006; Rigg et al., 2009; Kaimmer <strong>and</strong> Stoner,<br />

2008). <strong>The</strong> ampullae <strong>of</strong> Lorenzini is a short-range detection<br />

0964-5691/$ e see front matter Ó 2012 Elsevier Ltd. All rights reserved.<br />

http://dx.doi.org/10.1016/j.ocecoaman.2012.11.005<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


2<br />

C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10<br />

system, detecting bioelectric <strong>field</strong>s only centimeters away from<br />

prey to help the elasmobranch better pinpoint that prey prior to<br />

feeding (Kajiura <strong>and</strong> Holl<strong>and</strong>, 2002; Kajiura, 2003). Evidence also<br />

supports that elasmobranchs use this sensory system to detect<br />

geomagnetic <strong>field</strong>s assisting in the determination <strong>of</strong> geolocation<br />

<strong>and</strong> navigation (Klimley, 1993; Klimley et al., 2002). <strong>The</strong> latter,<br />

<strong>semiochemical</strong>s, appear to target the olfactory senses <strong>of</strong> elasmobranchs<br />

<strong>and</strong> appear to function as a Schreckst<strong>of</strong>f, or chemical alarm<br />

signal (von Frisch, 1938; Stroud, 2008b), producing a Schreckreaktion,<br />

or flight reaction, that is highly selective. Each potential<br />

repellent; magnets, EPMs, <strong>and</strong> <strong>semiochemical</strong>s, are hypothesized to<br />

operate under different mechanisms, each <strong>of</strong> which are described<br />

below.<br />

1.1. Ampullae <strong>of</strong> Lorenzini<br />

Elasmobranchs differ from bony fishes (teleosts) in various<br />

ways, but a key difference is that an elasmobranch possesses an<br />

<strong>electrosensory</strong> system known as the ampullae <strong>of</strong> Lorenzini. <strong>The</strong><br />

ampullae <strong>of</strong> Lorenzini were first discovered by Marcello Malpighi in<br />

1663, but described in detail by Stephano Lorenzini in 1678<br />

(Budker, 1971), after whom they were named. However, the function<br />

<strong>of</strong> these receptors remained unknown until Dijkgraaf <strong>and</strong><br />

Kalmijn (1963, 1966; cited by Kalmijn, 1971, 1982) demonstrated<br />

the electrosensitivity <strong>of</strong> these receptors, with additional experimentation<br />

showing these receptors have mechanical <strong>and</strong> thermal<br />

sensitivities (Boord <strong>and</strong> Campbell, 1977; Brown et al., 2003; Murray,<br />

1957, 1960).<br />

<strong>The</strong> ampullae <strong>of</strong> Lorenzini are acute <strong>electrosensory</strong> organs<br />

composed <strong>of</strong> minute jelly-filled pores located around the cephalic<br />

region <strong>of</strong> elasmobranchs (Fig. 1). Each ampullae contains a pore at<br />

the surface <strong>of</strong> the skin, which connects to a subcutaneous canal<br />

filled with conductive jelly (Kalmijn, 1966, 1971, 1974, 2000;<br />

Bastian, 1994). As an elasmobranch encounters a weak electrical<br />

impulse, such as that associated with muscle contraction (e.g.<br />

a fish’s heartbeat) or the associated electrical charges originating<br />

from gill movement, the voltage external to the pores <strong>of</strong> the<br />

ampullae differs from the interior voltage potential. This voltage<br />

gradient is perceived <strong>and</strong> a sensory message is sent to the brain via<br />

an impulse from the afferent neurons. It has been suggested that<br />

elasmobranchs detect magnetic <strong>field</strong>s in a similar manner using<br />

this <strong>electrosensory</strong> system, but through a different mechanism<br />

known as indirect-based magnetoreception via electromagnetic<br />

induction (Kalmijn, 1973, 1974, 1982, 1984; Carey <strong>and</strong> Scharold,<br />

1990; Klimley, 1993; Holl<strong>and</strong> et al., 1999) which will be described<br />

in more detail in the following section.<br />

1.2. Magnetic repellent concept<br />

Magnets vary by composition, strength, axis <strong>of</strong> polarization, size<br />

<strong>and</strong> shape. More specifically, the permanent magnets typically used<br />

in the current <strong>shark</strong> behavioral studies are ceramic magnets that do<br />

not degrade in seawater (e.g. bariumeferrite magnets-BaFe 12 O 19 ;<br />

Rigg et al., 2009; O’Connell et al., 2010, 2011a, 2011b) <strong>and</strong> rare-earth<br />

magnets that degrade in seawater (e.g. neodymiumeironeboron<br />

magnets-Nd 2 Fe 14 B; O’Connell et al., 2011b; Robbins et al., 2011).<br />

<strong>The</strong> magnetic flux associated with ceramic magnets is generally<br />

weaker (e.g. 3850 G) than rare-earth magnets (e.g. 14,800 G),<br />

whereas the axis <strong>of</strong> polarization exists in a variety <strong>of</strong> combinations<br />

regardless <strong>of</strong> magnet-type.<br />

<strong>The</strong> most current explanation as to how elasmobranchs detect<br />

magnetic <strong>field</strong>s is through the process <strong>of</strong> indirect magnetoreception<br />

by electromagnetic induction, or induced electric <strong>field</strong>s that arise<br />

when an object moves through a magnetic <strong>field</strong> (Kalmijn, 1973,<br />

1982, 1984). Through electromagnetic induction there are two<br />

proposed mechanisms on how elasmobranchs can detect magnetic<br />

<strong>field</strong>s: (1) active mode <strong>and</strong> (2) passive mode (Kalmijn, 1981, 1984;<br />

Montgomery <strong>and</strong> Walker, 2001). In active mode, as an elasmobranch<br />

is swimming with a horizontal velocity through the horizontal<br />

component <strong>of</strong> the Earth’s geomagnetic <strong>field</strong>, a vertical<br />

electromotive <strong>field</strong> is induced (Fig. 2a; Kalmijn, 1997; Montgomery<br />

<strong>and</strong> Walker, 2001). Paulin (1995) modified the concept <strong>of</strong> the active<br />

mode <strong>and</strong> stated that directional cues could originate from variations<br />

in induced electroreceptor voltage produced during turns<br />

(Montgomery <strong>and</strong> Walker, 2001). In passive mode, as electrically<br />

charged seawater moves through the vertical component <strong>of</strong> the<br />

Earth’s geomagnetic <strong>field</strong>, an electromotive <strong>field</strong>, which may be<br />

detected <strong>and</strong> utilized by an elasmobranch, is produced in the<br />

horizontal direction (Fig. 2b).<br />

<strong>The</strong> geomagnetic <strong>field</strong> on the Earth’s surface ranges from 0.25 to<br />

0.65 G (G). In comparison, the magnetic <strong>field</strong> associated with<br />

a bariumeferrite magnet (BaFe 12 O 19 ), produces a maximum flux <strong>of</strong><br />

approximately 1000 G. It is hypothesized that a much stronger<br />

electromotive <strong>field</strong> will be induced by a permanent magnet <strong>and</strong><br />

therefore will overstimulate the ampullae <strong>of</strong> Lorenzini <strong>of</strong> an<br />

approaching elasmobranch, thus producing a repellent response<br />

(WWF, 2006).<br />

Although electromagnetic induction is the current explanation<br />

as to how elasmobranchs detect the Earth’s geomagnetic <strong>field</strong>, firm<br />

evidence that this process governs magnetic <strong>field</strong> detection in<br />

elasmobranchs has yet to be presented <strong>and</strong> therefore further<br />

experimentation is needed.<br />

1.3. Electropositive metal concept<br />

Fig. 1. <strong>The</strong> ampullae <strong>of</strong> Lorenzini <strong>of</strong> a tiger <strong>shark</strong> (Galeocerdo cuvier). <strong>The</strong> ampullary<br />

pores are represented by the black dots located around the cephalic region <strong>of</strong> the<br />

<strong>shark</strong>. Ó Craig O’Connell.<br />

Electropositive metals (EPMs) can be found in Group I (e.g.<br />

Lithium), Group II (e.g. Magnesium), Group III (e.g. Yttrium), Group<br />

IIIA (the Lanthanides), <strong>and</strong> Group IIIB (the Actinoids) <strong>of</strong> the Periodic<br />

Table <strong>of</strong> the Elements. Magnesium <strong>and</strong> the lanthanide metals find<br />

the most practical application in <strong>shark</strong> bycatch reduction,<br />

producing st<strong>and</strong>ard reduction potentials <strong>of</strong> approximately 1.8 eV<br />

in water, a voltage that is orders <strong>of</strong> magnitude greater in strength<br />

than the bioelectric <strong>field</strong>s (i.e. nanovolts) that are encountered by<br />

elasmobranchs in search <strong>of</strong> prey. As an elasmobranch approaches<br />

the metal through the conductive seawater electrolyte, three<br />

potential scenarios exist: (1) the <strong>shark</strong>, whose skin is much more<br />

electronegative than the metal, completes a galvanic cell, or electrochemical<br />

cell that derives energy from spontaneous reductione<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10 3<br />

Fig. 2. a) ‘Active mode’ e Vertical induced electrical currents are produced as the <strong>shark</strong> swims through the horizontal component <strong>of</strong> the Earth’s magnetic <strong>field</strong>. <strong>The</strong> induced<br />

electrical current moves in a direction which opposes the change in magnetic <strong>field</strong> (Lenz’s Law). Illustration modified from Kalmijn (1974). b)‘Passive mode’ e Horizontal induced<br />

electrical currents are produced by the movement <strong>of</strong> ocean streams (i.e. currents) through the vertical component <strong>of</strong> the Earth’s magnetic <strong>field</strong>. Illustration modified from<br />

Montgomery <strong>and</strong> Walker (2001).<br />

oxidation (redox) reactions, to the metal <strong>and</strong> experiences a net<br />

charge increase that is detected by the ampullae <strong>and</strong> perceived as<br />

repellent, (2) the hydrated metal cation (aquo ion) is a weak Lewis<br />

acid, or electron pair acceptor, <strong>and</strong> therefore repellent to the <strong>shark</strong><br />

purely on low pH/irritancy effects, <strong>and</strong> (3) a <strong>shark</strong> retreats due to<br />

induced currents which are generated perpendicular to the resultant<br />

electric <strong>field</strong> (Rice, 2008; Stroud, 2008a: Fig. 3).<br />

1.4. Semiochemical repellent concept<br />

Elasmobranchs are considered to have a highly developed sense<br />

<strong>of</strong> smell which is evidenced by their large olfactory epithelial<br />

surface area (Schluessel et al., 2008). Due to this, elasmobranchs are<br />

highly responsive to very small concentrations <strong>of</strong> odor molecules,<br />

particularly those <strong>of</strong> amino acids (Tricas et al., 2009). Beginning in<br />

1990, several studies aimed to take advantage <strong>of</strong> an elasmobranchs<br />

acute sense <strong>of</strong> smell <strong>and</strong> demonstrated that elasmobranchs are<br />

chemically aware <strong>of</strong> <strong>semiochemical</strong>s derived from decaying <strong>shark</strong>s<br />

(Stroud, 2008b) <strong>and</strong> predatory heterospecifics (Rasmussen <strong>and</strong><br />

Schmidt, 1992). Semiochemicals are signals which animals may<br />

utilize to assist with mate selection (Nieberding et al., 2008),<br />

foraging (Tumlinson et al., 1993), aggregation (Verheggen et al.,<br />

2010) or navigation within their respected ecosystems, but <strong>semiochemical</strong><br />

derived from decaying <strong>shark</strong> tissue causes a reaction<br />

which is reminiscent <strong>of</strong> a “Schreckreaktion”, originally termed by<br />

von Frisch (1938) <strong>and</strong> is defined by the fright reaction <strong>of</strong> European<br />

minnows (Phoxinus phoxinus) when detecting the chemicals from<br />

an injured conspecific. Thus far, no published <strong>field</strong> work in the<br />

scientific literature exists on the newly discovered elasmobranchderived<br />

<strong>semiochemical</strong>s.<br />

2. Brief overview <strong>of</strong> <strong>electrosensory</strong> repellent studies<br />

To date, a total <strong>of</strong> eleven studies have been conducted which<br />

aimed to determine the effects <strong>of</strong> magnetic (Table 1) <strong>and</strong> EPM alloy<br />

(Table 2) stimuli on elasmobranch behavior. <strong>The</strong>se studies can be<br />

classified into three categories based on the experimental focus<br />

<strong>and</strong>/or potential applications: (1) basic or pro<strong>of</strong>-<strong>of</strong>-concept<br />

research, (2) recreational <strong>and</strong> commercial fishing applications,<br />

<strong>and</strong> (3) beach net applications. <strong>The</strong> findings from each <strong>of</strong> these<br />

eleven studies were highly inconsistent, which may stem from<br />

inexperience with using these materials. As with many new<br />

<strong>emerging</strong> <strong>field</strong>s, there is a learning curve, <strong>and</strong> the <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong><br />

repellents is in its infancy <strong>and</strong> requires substantially more<br />

<strong>field</strong> <strong>and</strong> laboratory investigations prior to the dismissal or<br />

implementation <strong>of</strong> any <strong>of</strong> these materials.<br />

2.1. Basic or pro<strong>of</strong>-<strong>of</strong>-concept research<br />

Fig. 3. This figure represents the mechanism behind electropositive metal detection in<br />

elasmobranchs. When an electropositive metal is placed in seawater, it undergoes<br />

spontaneous hydrolysis generating hydrogen gas <strong>and</strong> metal hydroxide which precipitates<br />

out <strong>of</strong> solution. <strong>The</strong> hypothesized mechanism behind the repellent capabilities <strong>of</strong><br />

the EPM alloys occurs as the cations transfer to the approaching electronegative <strong>shark</strong><br />

(i.e. a galvanic cell). This build up <strong>of</strong> positively charged ions overwhelms the ampullae<br />

<strong>of</strong> Lorenzini <strong>and</strong> is the most current explanation for the observed deterrent capabilities<br />

<strong>of</strong> the repellents; however, another hypothesized explanation is that a <strong>shark</strong> may<br />

retreat due to the associated induced current which is produced from the electric <strong>field</strong>/<br />

magnetic <strong>field</strong>. Illustration modified from Rice (2008).<br />

Four studies have been conducted which serve as pro<strong>of</strong>-<strong>of</strong>concept<br />

experiments, or basic experiments exploring the efficacy<br />

<strong>of</strong> <strong>electrosensory</strong> repellents, on several elasmobranch species. In<br />

2008, a study was conducted which examined the behavior <strong>of</strong><br />

Galapagos (Carcharhinus galapagensis) <strong>and</strong> s<strong>and</strong>bar (Carcharhinus<br />

plumbeus) <strong>shark</strong>s to NdePr (neodymiumepraseodymium) metal,<br />

a type <strong>of</strong> EPM (Wang et al., 2008). Results demonstrate that <strong>shark</strong>s<br />

fed more frequently from the lead-weight controls with a significantly<br />

greater quantity <strong>of</strong> aversive behaviors (e.g. sharp turns <strong>and</strong><br />

cease feeding) toward the NdePr metal illustrating the potential<br />

utility <strong>of</strong> these <strong>electrosensory</strong> stimuli for recreational <strong>and</strong><br />

commercial fishing applications (Fig. 4a).<br />

In a laboratory study, Rigg et al. (2009) evaluated the effects <strong>of</strong><br />

ferrite magnets on five elasmobranch bycatch species; scalloped<br />

hammerhead <strong>shark</strong> (Sphyrna lewini), Australian blacktip <strong>shark</strong><br />

(Carcharhinus tilstoni), gray reef <strong>shark</strong> (Carcharhinus amblyrhynchos),<br />

milk <strong>shark</strong> (Rhizoprionodon acutus), <strong>and</strong> speartooth <strong>shark</strong><br />

(Glyphis glyphis), as well as the teleost species, the seabass (Lates<br />

calcarifer). Results from this study demonstrated that all elasmobranch<br />

species responded at distances ranging from 0.26 to 0.58 m<br />

from the magnets <strong>and</strong> no signs <strong>of</strong> habituation were observed, with<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


4<br />

C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10<br />

Table 1<br />

Description <strong>of</strong> all studies conducted from 2008epresent using magnets (NdeFeeB ¼ neodymiumeironeboron magnets; BaeFeeO ¼ barium ferrite magnets; Ferrite<br />

magnets ¼ used either strontium ferrite or barium ferrite magnets e study did not specify) as elasmobranch repellents. Elasmobranch species which were studied are listed in<br />

the appropriate column, along with the purpose <strong>of</strong> the research (i.e. application). If the researchers were just evaluating the effects <strong>of</strong> repellents, but not for a particular<br />

fisheries application, the term “pro<strong>of</strong> <strong>of</strong> concept” was assigned. Additionally, success was ranked on a scale <strong>of</strong> “None” e no observed effects, “Yes” e induced some effect on all<br />

species tested, “Partial” e induced some effect on some, but not all species studied. Although Stoner <strong>and</strong> Kaimmer (2008) observed flinches by Squalus acanthias towards the<br />

magnets initially, when bait was present no deterrent effects were observed <strong>and</strong> therefore this study is categorized as “None” in the success category.<br />

Study Species studied Repellent tested Application Success<br />

Stoner <strong>and</strong> Kaimmer (2008) Pacific Squalus acanthias NdeFeeB Magnets Captive experiment None<br />

Rigg et al. (2009)<br />

Sphyrna lewini<br />

Ferrite Magnets Pro<strong>of</strong> <strong>of</strong> concept Yes<br />

Carcharhinus tilstoni<br />

Carcharhinus amblyrhynchos<br />

Rhizoprionodon acutus<br />

Glyphis glyphis<br />

O’Connell et al. (2010)<br />

Ginglymostoma cirratum<br />

BaeFeeO Magnets Pro<strong>of</strong> <strong>of</strong> concept Yes<br />

Dasyatis americana<br />

O’Connell et al. (2011a) Negaprion brevirostris BaeFeeO Magnets Captive experiment Yes<br />

O’Connell et al. (2011b)<br />

Carcharhinus limbatus<br />

BaeFeeO Magnets<br />

Recreational fishing<br />

Partial<br />

Carcharhinus plumbeus<br />

Dasyatis americana<br />

Mustelus canis<br />

Raja eglanteria<br />

Rhizoprionodon terraenovae<br />

NdeFeeB Magnets<br />

Longline fishing<br />

Robbins et al. (2011) Carcharhinus galapagensis Ferrite Magnets NdeFeeB Magnets Pro<strong>of</strong> <strong>of</strong> concept Partial<br />

the teleost L. calcarifer, not showing any aversive reactions to the<br />

magnets illustrating the selectivity <strong>of</strong> this technology. Additionally,<br />

the distances with which the elasmobranchs responded to the<br />

magnets differed, with C. tilstoni, R. acutus <strong>and</strong> S. lewini reacting at<br />

similar distances from the treatments <strong>and</strong> C. amblyrhynchos <strong>and</strong><br />

G. glyphis reacting at much closer distances from the magnets<br />

(Fig. 4b). <strong>The</strong>se species-specific responses to the magnets were<br />

suggested to be due to differences in habitat <strong>and</strong> feeding ecology,<br />

with certain habitats (e.g. coastal/turbid environments) <strong>and</strong> associated<br />

feeding ecologies requiring greater electroreception capabilities<br />

leading to greater sensitivity to magnetic <strong>field</strong>s.<br />

O’Connell et al., 2010 conducted a baseline <strong>field</strong> experiment<br />

which examined the behavioral responses <strong>of</strong> wild southern stingrays<br />

(Dasyatis americana) <strong>and</strong> nurse <strong>shark</strong>s (Ginglymostoma cirratum)<br />

to grade C8 bariumeferrite (BaFe 12 O 19 ) magnets. D. americana<br />

<strong>and</strong> G. cirratum exhibited a significantly greater quantity <strong>of</strong> avoidance<br />

behaviors <strong>and</strong> significantly fewer feeding behaviors toward<br />

magnetic treatments. <strong>The</strong> results from this study demonstrated the<br />

potential utility <strong>of</strong> bariumeferrite magnets as benthic<br />

elasmobranch-deterrent devices <strong>and</strong> in conjunction with Rigg et al.<br />

(2009), suggested that futures studies should test these repellents<br />

in a more applied sense.<br />

2.2. Recreational <strong>and</strong> longlining applications<br />

2.2.1. Focal order: Squaliformes<br />

<strong>The</strong> spiny dogfish (Squalus acanthias), has been subjected to<br />

severe overharvesting causing great concern to many fisheries<br />

managers (MAFMC, 1998). In response to this overharvesting, Fisheries<br />

Management Plans have been implemented by the Atlantic<br />

States Marine Fisheries Commission, the Mid-Atlantic Fisheries<br />

Management Council <strong>and</strong> the New Engl<strong>and</strong> Fishery Management<br />

Council, leading to the rebuilt status <strong>of</strong> the northwest Atlantic<br />

S. acanthias stock (Rago <strong>and</strong> Sosobee, 2009). Although population<br />

rejuvenation <strong>of</strong> S. acanthias has been successful in the northwest<br />

Atlantic, the development <strong>of</strong> bycatch reduction technology is <strong>of</strong><br />

upmost importance to minimize future S. acanthias bycatch on<br />

a variety <strong>of</strong> commercial gears. In response to the need for effective<br />

bycatch technology, a variety <strong>of</strong> studies have been conducted which<br />

examine the potential use <strong>of</strong> <strong>electrosensory</strong> materials on S. acanthias<br />

behavior to determine their utility on longline gear.<br />

Stoner <strong>and</strong> Kaimmer (2008) conducted laboratory investigations<br />

on the effects <strong>of</strong> an EPM alloy (64.02% cerium, 34.22% lanthanum,<br />

0.55% neodymium, 0.11% praseodymium <strong>and</strong> impurities) <strong>and</strong><br />

neodymiumeironeboron magnets on Pacific spiny dogfish<br />

Table 2<br />

Description <strong>of</strong> all studies conducted from 2008epresent using electropositive metal (EPM) alloys (CeeLaeNdePr ¼ alloy primarily composed <strong>of</strong> cerium, lanthanum,<br />

neodymium, <strong>and</strong> praseodymium; NdePr ¼ alloy primarily composed <strong>of</strong> neodymium <strong>and</strong> praseodymium; CeeLa ¼ alloy primarily composed <strong>of</strong> cerium <strong>and</strong> lanthium) or a pure<br />

EPM (Nd ¼ neodymium) as elasmobranch repellents. Elasmobranch species which were studied are listed in the appropriate column, along with the purpose <strong>of</strong> the research<br />

(i.e. application). If the researchers were just evaluating the effects <strong>of</strong> repellents, but not for a particular fisheries application, the term “pro<strong>of</strong> <strong>of</strong> concept” was assigned.<br />

Additionally, success was ranked on a scale <strong>of</strong> “None” e no observed effects, “Yes” e induced some effect on all species tested, “Partial” e induced some effect on several, but not<br />

all species studied.<br />

Study Species studied Repellent tested Application Success<br />

Brill et al. (2009) Carcharhinus plumbeus CeeLaeNdePr Alloy Captive experiments Yes<br />

Brill et al. (2009) Carcharhinus plumbeus NdePr Alloy Longline fishing Yes<br />

Jordan et al. (2011)<br />

Squalus acanthias<br />

Nd Captive experiment Partial<br />

Mustelus canis<br />

Kaimmer <strong>and</strong> Stoner (2008)<br />

Squalus acanthias<br />

CeeLaeNdePr Alloy Longline fishing Yes<br />

Raja rhina<br />

Robbins et al. (2011) Carcharhinus galapagensis NdePr Alloy Pro<strong>of</strong> <strong>of</strong> concept No<br />

Stoner <strong>and</strong> Kaimmer (2008) Squalus acanthias CeeLaeNdePr Alloy Captive experiments Yes<br />

Tallack <strong>and</strong> M<strong>and</strong>elman (2009) Squalus acanthias CeeLa Alloy Captive experiments No<br />

Tallack <strong>and</strong> M<strong>and</strong>elman (2009) Squalus acanthias CeeLa Alloy Longline fishing<br />

No<br />

Jigging<br />

Wang et al. (2008)<br />

Carcharhinus galapagensis<br />

Carcharhinus plumbeus<br />

NdePr alloy Pro<strong>of</strong> <strong>of</strong> Concept Yes<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10 5<br />

Fig. 4. a) Total aversion responses (p < 0.01) <strong>of</strong> Galapagos (Carcharhinus galapagensis) <strong>and</strong> s<strong>and</strong>bar <strong>shark</strong>s (Carcharhinus plumbeus) towards the lead weight control <strong>and</strong> electropositive<br />

metal (NdePr metal alloy). Modified from Wang et al. (2008). b) <strong>The</strong> mean reaction distance to turn from ferrite magnets for five elasmobranch species: gray reef <strong>shark</strong><br />

(Carcharhinus amblyrhynchos), Australian blacktip <strong>shark</strong> (Carcharhinus tilstoni), milk <strong>shark</strong> (Rhizoprionodon acutus), scalloped hammerhead (Sphyrna lewini), <strong>and</strong> speartooth <strong>shark</strong><br />

(Glyphis glyphis). <strong>The</strong> means for C. tilstoni, R. acutus, <strong>and</strong> S. lewini were not significantly different; however, both C. amblyrhynchos <strong>and</strong> G. glyphis were. Error bars represent 95%<br />

confidence limits. Modified from Rigg et al. (2009).<br />

(S. acanthias) <strong>and</strong> Pacific halibut (Hippoglossus stenolepis). Results<br />

demonstrated that metals significantly altered the feeding behavior<br />

<strong>of</strong> interacting S. acanthias <strong>and</strong> not H. stenolepis, whereas magnets<br />

were observed to have little effect during baited experimentation.<br />

Encouraged by the results <strong>of</strong> the laboratory studies, Kaimmer <strong>and</strong><br />

Stoner (2008) conducted <strong>field</strong> investigations using an EPM alloy <strong>of</strong><br />

same elemental composition as used in Stoner <strong>and</strong> Kaimmer (2008)<br />

to examine its deterrent capabilities in a commercial fishery targeting<br />

Pacific halibut (H. stenolepis), sablefish (Anoplopoma fimbria),<br />

<strong>and</strong> Pacific cod (Gadus macrocephalus) near Homer, Alaska. EPM<br />

alloy-associated hooks yielded a 19% reduction in S. acanthias<br />

bycatch <strong>and</strong> a 46% reduction <strong>of</strong> longnose skate (Raja rhina) bycatch in<br />

comparison to controls. Besides the catch <strong>of</strong> both S. acanthias <strong>and</strong><br />

R. rhina, several teleosts species were captured in high frequency,<br />

including H. stenolepis <strong>and</strong> sculpins (unclassified); however, neither<br />

showed catch trends on EPM alloy-associated hooks demonstrating<br />

the selective characteristics <strong>of</strong> the metal.<br />

In contrast to the positive findings reported by Kaimmer <strong>and</strong><br />

Stoner (2008) <strong>and</strong> Stoner <strong>and</strong> Kaimmer (2008), Tallack <strong>and</strong><br />

M<strong>and</strong>elman (2009) conducted a very similar study with contradicting<br />

results. Tallack <strong>and</strong> M<strong>and</strong>elman analyzed the effect <strong>of</strong> an<br />

EPM alloy, composed mainly <strong>of</strong> cerium <strong>and</strong> lanthanide, on the<br />

Atlantic spiny dogfish (S. acanthias) in both laboratory <strong>and</strong> <strong>field</strong><br />

studies. Laboratory results provided minimal evidence <strong>of</strong> deterrent<br />

success, with signs <strong>of</strong> deterrence being highly correlated with food<br />

deprivation level. Additionally, in both longline <strong>and</strong> jigging experiments,<br />

no observed reduction in S. acanthias capture was observed,<br />

<strong>and</strong> too few haddock (Melanogrammus aeglefinus), redfish (Sebastes<br />

fasciatus) <strong>and</strong> sea raven (Hemitripterus americanus) were captured<br />

to conduct any statistical analyses.<br />

Lastly, Jordan et al. (2011) conducted a comparative laboratory<br />

experiment examining the electro-sensitivities <strong>of</strong> two species <strong>of</strong><br />

<strong>shark</strong>, S. acanthias <strong>and</strong> the smooth dogfish (Mustelus canis);<br />

however, the results pertaining to M. canis will be discussed in<br />

a later section. For the first part <strong>of</strong> this study, S. acanthias exhibited<br />

a sensitivity to weak electric <strong>field</strong>s (


6<br />

C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10<br />

Fig. 5. <strong>The</strong> positioning at 1 s intervals (black dots) <strong>of</strong> one juvenile s<strong>and</strong>bar <strong>shark</strong> (Carcharhinus plumbeus) measured over a 2 h duration (1 h lead weight control <strong>and</strong> 1 h EPM). a) <strong>The</strong><br />

positioning at 1 s intervals (black dots) <strong>of</strong> one juvenile C. plumbeus to three lead ingots (gray triangle) suspended in the tank for a duration <strong>of</strong> 1 h b) Experimental treatment<br />

examining the swimming pattern <strong>of</strong> one C. plumbeus at 1 s intervals in response to three EPM ingots (gray triangle) suspended in the tank for 1 h. Significantly fewer <strong>shark</strong><br />

interactions occurred within 100 cm <strong>of</strong> the electropositive metal bars (b), in comparison to the lead ingots (a). Illustration taken from Brill et al. (2009).<br />

neodymiumeironeboron magnets significantly reduced total elasmobranch<br />

<strong>and</strong> capture <strong>of</strong> several elasmobranch species: Atlantic<br />

sharpnose <strong>shark</strong> (Rhizoprionodon terraenovae) <strong>and</strong> smooth dogfish<br />

(M. canis). For the inshore longline experiment, results demonstrated<br />

that neodymiumeironeboron magnets had no impact on elasmobranch<br />

capture whereas the bariumeferrite magnet significantly<br />

reduced total elasmobranch capture, in addition to the capture <strong>of</strong> the<br />

blacktip <strong>shark</strong> (Carcharhinus limbatus) <strong>and</strong>thesouthernstingray(D.<br />

americana). Teleosts, such as red drum (Sciaenops ocellatus), Atlantic<br />

croaker (Micropogonias undulatus), oyster toadfish (Opsanus tau),<br />

black sea bass (Centropristis striata), <strong>and</strong> the bluefish (Pomatomus<br />

saltatrix), showed no hook preference in either hook-<strong>and</strong>-line or<br />

longline studies. This study concluded that magnet-induced repellent<br />

behaviors may be a species-specific phenomenon <strong>and</strong> the magnetic<br />

characteristic known as the axis <strong>of</strong> polarization, similar to findings in<br />

Robbins et al. (2011), may be the key component to finding an effective<br />

magnetic <strong>shark</strong> repellent (Fig. 6b).<br />

2.3. Beach net applications<br />

Fig. 6. a) This illustration represents the 50 mm magnetic rare-earth magnetic discs<br />

which reduced Galapagos <strong>shark</strong> (Carcharhinus galapagensis) depredation rate by 50%.<br />

<strong>The</strong> magnetic flux along the bait ranged from 372 to 1474 G. Illustration taken from<br />

Robbins et al. (2011). b) This illustration represents the grade C8 bariumeferrite<br />

permanent magnet which significantly reduced overall elasmobranch capture in<br />

inshore longline trials. <strong>The</strong> axis <strong>of</strong> polarization extended vertically, over the entire bait<br />

<strong>and</strong> had a maximum flux <strong>of</strong> 3850 G. Illustration modified from O’Connell et al. (2011b).<br />

Neither image is drawn to scale.<br />

Beach nets are devices used to minimize the potential interaction<br />

between predatory <strong>shark</strong> species (e.g. great white <strong>shark</strong>-<br />

Carcharodon carcharias, tiger <strong>shark</strong>-Galeocerdo cuvier, <strong>and</strong> bull<br />

<strong>shark</strong>-Carcharhinus leucas) <strong>and</strong> beachgoers (Cliff <strong>and</strong> Dudley, 1992;<br />

Dudley, 1997). Although these nets are successful at minimizing<br />

this interaction, experimental evidence demonstrates that local<br />

elasmobranch populations have plummeted (Dudley, 1997; Stevens<br />

et al., 2000; Dudley <strong>and</strong> Cliff, 1993).<br />

As a means to examine the potential utility <strong>of</strong> ceramic magnets<br />

to exclude <strong>shark</strong>s from netted areas, O’Connell et al. (2011a) conducted<br />

a captive study where lemon <strong>shark</strong>s (Negaprion brevirostris)<br />

were individually subjected to a fence-like apparatus containing<br />

both a control (clay bricks) <strong>and</strong> magnetic (grade C8 bariumeferrite<br />

magnets) opening. Results demonstrated that ceramic magnets<br />

were efficient at manipulating the swimming behavior <strong>of</strong><br />

N. brevirostris; however, barrier trials <strong>and</strong> tonic immobility trials<br />

demonstrated that habituation to the magnetic <strong>field</strong>s occurs due to<br />

repeated stimulation over a short duration (Fig. 7). It is uncertain<br />

how these results would relate to experiments conducted on wild<br />

<strong>shark</strong>s, but was deemed unlikely to be as high as the frequency <strong>of</strong><br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10 7<br />

Fig. 7. <strong>The</strong> total behaviors from three lemon <strong>shark</strong>s (Negaprion brevirostris), with the treatment variables (C¼Control; M ¼ Magnet) listed on the x-axis. a) <strong>The</strong> total avoidance<br />

behaviors (p < 0.001) <strong>of</strong> the three <strong>shark</strong>s in trial one <strong>and</strong> trial two. b) <strong>The</strong> total number <strong>of</strong> entrances (p < 0.001) <strong>of</strong> the three <strong>shark</strong>s in trial one <strong>and</strong> trial two. Illustration taken from<br />

O’Connell et al. (2011a).<br />

interaction seen in this experiment <strong>and</strong> thus is unknown if sensory<br />

habituation would occur.<br />

3. Factors affecting repellent success<br />

Prior to drawing conclusions from the current body <strong>of</strong> research<br />

pertaining to <strong>electrosensory</strong> repellents <strong>and</strong> as a means to underst<strong>and</strong><br />

inconsistent results, it is imperative that scientists monitor<br />

a variety <strong>of</strong> biological <strong>and</strong> environmental parameters which may<br />

influence repellent success. Although it is important that scientists<br />

analyze the basic behavioral responses <strong>of</strong> elasmobranchs toward<br />

various <strong>electrosensory</strong> stimuli, studying parameters that may alter<br />

success will give scientists valuable insight as to methods <strong>and</strong><br />

conditions which will yield the most successful implementation.<br />

3.1. Satiation<br />

<strong>The</strong> availability <strong>of</strong> prey can be influenced by a variety <strong>of</strong> factors,<br />

including: climate change (e.g. Loeb et al., 1997; Polovina, 1996),<br />

fisheries (e.g. Bearzi et al., 2006), <strong>and</strong> predator abundance (e.g.<br />

Connell, 1998). With changing prey densities, predator satiation<br />

may concurrently change which may greatly influence predator<br />

behavior (Williamson, 1980). To determine how the level <strong>of</strong> satiation<br />

could effect <strong>electrosensory</strong> repellent success, several laboratory<br />

studies were conducted (Stoner <strong>and</strong> Kaimmer, 2008; Tallack<br />

<strong>and</strong> M<strong>and</strong>elman, 2009). Stoner <strong>and</strong> Kaimmer (2008) demonstrated<br />

that EPM repellent effectiveness on S. acanthias was found<br />

to be highly correlated with food deprivation where increasing<br />

levels <strong>of</strong> food deprivation (2 <strong>and</strong> 4 days) resulted in decreased<br />

repellent success. Similarly, Tallack <strong>and</strong> M<strong>and</strong>elman (2009)<br />

demonstrated that the responsiveness <strong>of</strong> S. acanthias to EPM<br />

stimuli was inversely proportional to the level <strong>of</strong> food deprivation.<br />

<strong>The</strong>refore one influential factor which may be pertinent to the<br />

success <strong>of</strong> <strong>electrosensory</strong> deterrents is animal satiation. It is unclear<br />

how these laboratory findings would translate to the <strong>field</strong> <strong>and</strong><br />

future experimentation is needed; however, if these findings hold<br />

true, situations where the level <strong>of</strong> satiation may be highest (e.g.<br />

seasonality <strong>of</strong> prey abundance, lack <strong>of</strong> fishing pressure, healthy<br />

ecosystem) may increase the likelihood <strong>of</strong> repellent success <strong>and</strong><br />

thus be pertinent to future implementation strategies.<br />

3.2. Water visibilityeturbidity <strong>and</strong> light intensity<br />

Little is known about the effect <strong>of</strong> water visibility on elasmobranch<br />

sensory allocation; however, preliminary experimentation<br />

with the bull <strong>shark</strong> (C. leucas) illustrates that increased turbidity<br />

<strong>and</strong> low water visibility conditions increase the repellent effectiveness<br />

<strong>of</strong> ceramic magnets (O’Connell et al., submitted for<br />

publication). Similar to where blind humans show enhanced<br />

auditory localization capabilities in comparison to sighted humans<br />

(Lessard et al., 1998; Muchnik et al., 1991; Rice, 1970), it may be<br />

possible that the <strong>electrosensory</strong> capabilities <strong>of</strong> elasmobranchs are<br />

altered in conditions when vision is compromised, therefore<br />

maximizing the effectiveness <strong>of</strong> <strong>electrosensory</strong> repellents.<br />

Upon further consideration, although turbidity may drastically<br />

alter the extent <strong>of</strong> the visual <strong>field</strong> <strong>of</strong> elasmobranchs <strong>and</strong> may yield<br />

differences in sensitivity to <strong>electrosensory</strong> repellents, the <strong>electrosensory</strong><br />

differences due to variations in light intensity (e.g. day<br />

versus night) may be minimal due to the light-reflecting layer<br />

behind the retina known as the tapetum lucidum (Gruber <strong>and</strong><br />

Cohen, 1985). This structure maximizes light available for lowlight<br />

conditions. <strong>The</strong>refore although future studies should not<br />

neglect light intensity, turbidity may be a more influential factor in<br />

<strong>electrosensory</strong> repellent success <strong>and</strong> may provide pertinent information<br />

as to the locations which will yield the most success (e.g.<br />

turbid coastal environments).<br />

3.3. Presence <strong>of</strong> conspecifics <strong>and</strong> heterospecifics<br />

Intra- <strong>and</strong> inter-specific competition is highly reported in the<br />

literature (Crombie, 1947; Connell, 1961; Polis, 1981; Stiling et al.,<br />

1984; Munday et al., 2001) <strong>and</strong> illustrates that the presence <strong>of</strong><br />

conspecifics <strong>and</strong>/or heterospecifics induces a competitive<br />

mentality, <strong>of</strong>ten leading to an abrupt change in behavior. Brill et al.<br />

(2009) conducted a laboratory experiment which examined the<br />

feeding response <strong>of</strong> juvenile s<strong>and</strong>bar <strong>shark</strong>s (C. plumbeus) attwo<br />

different density levels (n ¼ 7; n ¼ 14) toward an EPM alloy. During<br />

high-density trials (e.g. 14 <strong>shark</strong>s), the deterrent effects were<br />

minimal <strong>and</strong> short-lived; however, C. plumbeus demonstrated<br />

a sensitivity to EPM alloys during low-density trials <strong>and</strong> therefore<br />

the researchers concluded that competition plays a role in the<br />

effectiveness <strong>of</strong> EPM alloys. (Brill et al., 2009). In a more recent<br />

study conducted by Robbins et al. (2011), the efficacy <strong>of</strong> seven rareearth<br />

(neodymiumeironeboron) magnet, two ferrite magnet, <strong>and</strong><br />

two EPM (NdePr: neodymiumepraseodymium) alloy configurations<br />

on reducing the depredation rate <strong>of</strong> the Galapagos <strong>shark</strong> (C.<br />

galapagensis) on baited lines was examined. During high density<br />

levels (greater than 3 <strong>shark</strong>s) results illustrate that EPMs were<br />

ineffective; however, in low density trials behavior was exploratory<br />

<strong>and</strong> cautious. With these findings, this study concluded that<br />

conspecific density greatly impacted repellent success <strong>and</strong> suggested<br />

that these repellents may be better suited for recreational<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


8<br />

C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10<br />

<strong>and</strong> commercial fishing industries where <strong>shark</strong> densities are<br />

sufficiently low. Lastly, Jordan et al. (2011) examined the deterrent<br />

capabilities <strong>of</strong> an EPM stimulus (NeodymiumeNd) on the smooth<br />

dogfish (M. canis). When tested in groups, M. canis preferred to feed<br />

from EPM-associated baits, whereas individually tested M. canis<br />

were significantly deterred by EPM-associated baits. This demonstrates<br />

that the effectiveness <strong>of</strong> EPMs, or more specifically,<br />

Neodymium on M. canis, may be heavily influenced by conspecific<br />

density. <strong>The</strong>refore, future studies should account for animal density<br />

as these studies illustrate that density-induced behavioral changes<br />

toward <strong>electrosensory</strong> deterrents is plausible.<br />

3.4. Salinity<br />

In regards to EPMs, the conductive medium is an essential<br />

component to the successful transfer <strong>of</strong> cations from the surface <strong>of</strong><br />

the metal to the approaching electronegative <strong>shark</strong>. <strong>The</strong> lower the<br />

resistivity, or opposition <strong>of</strong> electric current, as seen in highly saline<br />

environments (Chave et al., 1991; Eidesmo et al., 2002), will result<br />

in the successful transfer <strong>of</strong> these cations <strong>and</strong> prove the metals to<br />

be more effective (Stroud, 2008a). It is essential that future studies<br />

monitor salinity, as deploying metals in open ocean regions (i.e.<br />

peak salinity) may produce drastically different results in<br />

comparison to estuarine or nearshore environments that are<br />

heavily influenced by freshwater input.<br />

It is uncertain how salinity may impact elasmobranch magnetoreception.<br />

In order for elasmobranch magnetoreception to occur,<br />

two key elements are required for active <strong>and</strong>/or passive electromagnetic<br />

induction: a magnetic <strong>field</strong> (B) <strong>and</strong> velocity (v) (Kalmijn,<br />

1982, 1984). For the active mode <strong>of</strong> electromagnetic induction, the<br />

velocity corresponds to movement <strong>of</strong> the organism through B h<br />

(horizontal portion <strong>of</strong> Earth’s geomagnetic <strong>field</strong>); however, in the<br />

passive mode, the velocity component refers to the velocity <strong>of</strong> the<br />

ocean currents through B v (vertical portion <strong>of</strong> Earth’s geomagnetic<br />

<strong>field</strong>). Although v <strong>and</strong> B are key elements required for elasmobranch<br />

magnetic <strong>field</strong> detection, another element is the conductive<br />

medium that the animal is swimming within (Kalmijn, 1974, 1984,<br />

1997). <strong>The</strong> resistivity <strong>of</strong> seawater <strong>and</strong> freshwater drastically differs,<br />

with freshwater being characterized <strong>of</strong> having a resistivity <strong>of</strong> 2e<br />

200 U m <strong>and</strong> seawater 0.3 U m(Pashley et al., 2005; Chave et al.,<br />

1991; Eidesmo et al., 2002). In lower salinity conditions, the electric<br />

potential gradients between the pore (apical) <strong>and</strong> internal<br />

(basal) structures that are responsible for neural activity, tend to be<br />

minimal thus making the induction mechanism less efficient in<br />

freshwater (Kalmijn, 1974, 1981, 1984). <strong>The</strong>refore, salinity may<br />

impact the magnetoreception capabilities <strong>of</strong> elasmobranchs <strong>and</strong><br />

estuarine, brackish water, or even freshwater ecosystems may not<br />

be ideal ecosystems for magnetic deterrents (Kalmijn, 1974, 1984).<br />

3.5. Animal size/canal length<br />

Ontogentic shifts in electroreceptor sensitivity has been<br />

demonstrated with several elasmobranch species (Sisneros et al.,<br />

1998; Sisneros <strong>and</strong> Tricas, 2002). It is suggested that these shifts<br />

or increases in sensitivity as an animal matures is directly related to<br />

increasing ampullary canal legnth. <strong>The</strong> ampullae <strong>of</strong> Lorenzini<br />

function on voltage gradients, with neurotransmitters being<br />

released to primary afferent neurons when a difference exists<br />

between the pore (apical) <strong>and</strong> internal (basal) potentials (Bennett,<br />

1971; Tricas, 2001). Thus enhanced sensitivity is a function <strong>of</strong> canal<br />

length since greater voltage gradients exist between the pore<br />

surface <strong>and</strong> internal ampullary canal as ampullary canals lengthen.<br />

In addition, previous studies pertaining to the Atlantic stingray<br />

(Dasyatis sabina) <strong>and</strong> clearnose skate (Raja eglanteria) demonstrate<br />

that with maturity there is a gain <strong>of</strong> <strong>electrosensory</strong> primary<br />

afferents <strong>and</strong>, presumably, neural sensitivity (Sisneros et al., 1998;<br />

Sisneros <strong>and</strong> Tricas, 2002). In R. eglanteria, the neural sensitivity<br />

was eight times greater in adults <strong>and</strong> five times greater in juveniles<br />

in comparison to embryos (Sisneros et al., 1998). Similarly, in<br />

D. sabina, the neural sensitivity was four times greater in adults <strong>and</strong><br />

three times greater in juveniles when being compared to embryos<br />

(Sisneros <strong>and</strong> Tricas, 2002). <strong>The</strong>refore, animal size which is related<br />

to ampullary canal length, <strong>and</strong> in some cases, an increase in the<br />

quantity <strong>of</strong> afferent neurons, may influence the success <strong>of</strong> <strong>electrosensory</strong><br />

deterrents, with larger <strong>and</strong> more mature animals having<br />

enhanced sensitivity <strong>and</strong> thus a greater potential to be deterred.<br />

4. Commercial application<br />

Besides assessing the effectiveness <strong>of</strong> these repellents on<br />

interacting species, it is essential to also consider a variety <strong>of</strong> other<br />

characteristics/issues pertaining to these repellents, such as: cost,<br />

safety, pollution, <strong>and</strong> h<strong>and</strong>ling logistics.<br />

4.1. Cost <strong>of</strong> repellents<br />

<strong>The</strong> use <strong>of</strong> certain EPM <strong>and</strong> magnets may be prohibitively<br />

expensive when deployed in large quantities, such as for<br />

commercial fishing hooks. Jordan et al. (2011) proposed the use <strong>of</strong><br />

neodymium metal due to its <strong>field</strong> strength <strong>and</strong> relative cost in<br />

comparison to other lanthanide metals, but it is likely that using<br />

a pure lanthanide metal in a commercial fishery is economically<br />

impractical. While the PEW Institute (Cos<strong>and</strong>ey-Goden <strong>and</strong><br />

Morgan, 2011) conclude that “significant cost” precludes EPMs as<br />

an “economically viable option”, the authors fail to consider<br />

abundant <strong>and</strong> cost-effective reactive metals. Metals with a higher<br />

reduction potential at a fraction <strong>of</strong> the cost <strong>of</strong> pure lanthanides have<br />

been shown to be effective (O’Connell et al., submitted for<br />

publication). Magnesium <strong>and</strong> certain magnesium alloys, at $0.05<br />

US/gram in quantity for 99% þ purity, are much more practical <strong>and</strong><br />

easier to fabricate for fishing gear requirements.<br />

4.2. Magnet-associated issues<br />

Besides varying successes with magnetic repellents (Rigg et al.,<br />

2009; O’Connell et al., 2010, 2011a, 2011b), there exists a variety <strong>of</strong><br />

issues which one must consider prior to implementation <strong>and</strong>/or<br />

application. (1) When working in close proximity to neodymiume<br />

ironeboron <strong>and</strong> bariumeferrite magnets on a research vessel or<br />

platform, these magnets can be prohibitively difficult to work with<br />

<strong>and</strong> potentially very dangerous <strong>and</strong> cause bodily harm if a finger or<br />

other body part becomes trapped between two magnets. (2) <strong>The</strong>se<br />

materials have the ability to permanently destroy electronic<br />

equipment (e.g. cameras <strong>and</strong> computers) <strong>and</strong> thus should be<br />

h<strong>and</strong>led with care. (3) When applying magnets directly to a fishing<br />

hook, the presence <strong>of</strong> this large visual stimulus (e.g. the magnet)<br />

may potentially interfere with target catch unless hooks are<br />

magnetized thus removing the visual component. (4) If applying<br />

magnets directly to fishing hooks or for beach net-associated<br />

applications, it is inevitable that some magnets will be lost<br />

causing pollution. (5) Lastly, it is currently not understood how<br />

exposing organisms to strong magnetic <strong>field</strong>s may impact their<br />

electroreception <strong>of</strong> navigational capabilities <strong>and</strong> thus prior to<br />

commercial application this topic should be extensively studied.<br />

4.3. EPM-associated issues<br />

EPMs have shown promise in both laboratory (Stoner <strong>and</strong><br />

Kaimmer, 2008; Brill et al., 2009; Jordan et al., 2011) <strong>and</strong> <strong>field</strong><br />

experimentation (Kaimmer <strong>and</strong> Stoner, 2008; Wang et al., 2008).<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10 9<br />

However, several issues arise when working with these materials:<br />

(1) <strong>The</strong>se materials are flammable <strong>and</strong> this presents a variety <strong>of</strong><br />

challenges both on l<strong>and</strong> <strong>and</strong> at sea. (2) EPMs undergo rapid hydrolysis<br />

when placed in seawater. Although EPM degradation varies<br />

depending on reactivity, it is uncertain how cost-effective these<br />

materials will be if in constant need <strong>of</strong> replacement. (3) Similar to<br />

magnets, when applied to a hook an EPM ingot creates a large visual<br />

stimulus which may impact target catch. (4) <strong>The</strong> constant utilization,<br />

degradation <strong>and</strong> potential loss <strong>of</strong> materials in seawater is<br />

a source <strong>of</strong> pollution <strong>and</strong> must be taken into consideration. (5) Most<br />

importantly, it is unknown how these materials may physiologically<br />

affect interacting organisms in short <strong>and</strong> long term basis.<br />

4.4. Semiochemical-associated issues<br />

To date, little information exists pertaining to <strong>semiochemical</strong>s.<br />

Of the evidence that does exist, it was determined that very small<br />

quantities <strong>of</strong> <strong>semiochemical</strong>s, approximately 50 10 6 L are<br />

needed to evoke a repellent response while a <strong>shark</strong> is in tonic<br />

immobility (Stroud, 2008b). But, even with this minimal introduction<br />

<strong>of</strong> chemicals into seawater, a variety <strong>of</strong> issues can arise<br />

unless the chemicals are derived from natural sources. (1) <strong>The</strong><br />

chemicals being introduced into the ecosystem are synthetic or<br />

superpotent <strong>and</strong> thus may serve as an environmental pollutant,<br />

although as stated in Stroud (2008b) components <strong>of</strong> chemical<br />

repellents are compliant with regulations pertaining to the United<br />

States Environmental Protection Agency, the Food <strong>and</strong> Drug<br />

Administration, the Department <strong>of</strong> Transportation, <strong>and</strong> the<br />

National Marine Fisheries Service. (2) In addition with chemicals,<br />

their success will be heavily dependent on currents <strong>and</strong><br />

geographical features <strong>of</strong> the area. In situations where currents are<br />

slack or minimal, this will lead to minimal chemical dispersion <strong>and</strong><br />

may make the chemical nearly ineffective at far distances. (3) In<br />

situations when <strong>shark</strong>s are approaching from the opposite direction<br />

from the source <strong>of</strong> the <strong>semiochemical</strong>, it may also minimize the<br />

effectiveness <strong>of</strong> the chemicals. (4) Lastly, it is unknown how the<br />

<strong>semiochemical</strong>s may affect the olfactory <strong>and</strong>/or gustatory receptors<br />

or other biological tissues <strong>of</strong> interacting organisms. <strong>The</strong>refore<br />

extensive future research is needed on these aspects.<br />

5. Conclusion <strong>and</strong> future directions<br />

With the wide range <strong>of</strong> experiments examining the effects <strong>of</strong><br />

<strong>electrosensory</strong> stimuli on elasmobranchs, evidence supports that<br />

elasmobranchs can detect the stimuli (e.g. Kaimmer <strong>and</strong> Stoner,<br />

2008; Rigg et al., 2009), but the species-specificity <strong>of</strong> the repellent<br />

responses (e.g. O’Connell et al., 2011b, Jordan et al., 2011) remains<br />

unclear. It is essential for future research to underst<strong>and</strong>: (1) the<br />

environmental or physiological characteristics which may be most<br />

responsible for <strong>electrosensory</strong> sensitivity to magnetic or EPM<br />

repellents, (2) the best applications for each type <strong>of</strong> repellent with<br />

the recreational fishery being a possible c<strong>and</strong>idate for future<br />

research due to minimal experimentation thus far, <strong>and</strong> (3) to<br />

underst<strong>and</strong> the physiological effects <strong>of</strong> these repellents on interacting<br />

elasmobranchs. This <strong>field</strong> should no longer be solely assessing<br />

whether <strong>electrosensory</strong> repellents work as evidence now demonstrates<br />

the efficacy <strong>of</strong> repellents (Kaimmer <strong>and</strong> Stoner, 2008; Brill<br />

et al., 2009; Rigg et al., 2009; O’Connell et al., 2010; O’Connell<br />

et al., 2011a,b; Jordan et al., 2011), but this <strong>field</strong> may be enhanced<br />

through rigorous laboratory investigations, or <strong>field</strong> investigations<br />

when possible, which focus on what parameters maximize repellent<br />

effectiveness so future implementation will maximize the likelihood<br />

for success. Lastly, little evidence has been presented on the<br />

behavioral effects <strong>of</strong> current <strong>shark</strong> repellent technologies on teleost,<br />

mammalian, or chelonian species. Studies must be conducted on<br />

these species, as it will be essential prior to the final implementation<br />

<strong>of</strong> any repellents technology to fishing gears or beach nets.<br />

Acknowledgments<br />

We first would like to thank the School <strong>of</strong> Marine Science <strong>and</strong><br />

Technology (SMAST) at the University <strong>of</strong> Massachusetts Dartmouth<br />

for their support. We thank Dr. Adrianus J. Kalmijn for sharing his<br />

invaluable knowledge pertaining to the mechanisms underlying<br />

elasmobranch magnetoreception. Lastly, we would like to thank Dr.<br />

John M<strong>and</strong>elman, Dr. Laura Jordan <strong>and</strong> Dr. Steve Kajiura for their<br />

editorial comments as they greatly benefitted this manuscript.<br />

References<br />

Bastian, J., 1994. Electrosensory organisms. Phys. Today 47, 30e37.<br />

Baum, J.K., Myers, R.A., 2004. Shifting baselines <strong>and</strong> the decline <strong>of</strong> pelagic <strong>shark</strong>s in<br />

the Gulf <strong>of</strong> Mexico. Ecol. Lett. 7, 135e145.<br />

Baum, J.K., Myers, R.A., Kehler, D.G., Worm, B., Harley, S.J., Doherty, P.A., 2003.<br />

Collapse <strong>and</strong> conservation <strong>of</strong> <strong>shark</strong> populations in the Northwest Atlantic.<br />

Science 299, 389e392.<br />

Bearzi, G., Politi, E., Agazzi, S., Azzellino, A., 2006. Prey depletion caused by overfishing<br />

<strong>and</strong> the decline <strong>of</strong> marine megafauna in eastern Ionian Sea coastal<br />

waters (central Mediterranean). Biol. Conserv. 127 (4), 373e382.<br />

Bennett, M.V.L., 1971. Electroreception. In: Hoar, W.S., R<strong>and</strong>all, D.J. (Eds.), Fish<br />

Physiology, vol. 5. Academic Press, New York, pp. 493e574.<br />

Bonfil, R., 1994. Overview <strong>of</strong> world elasmobranch fisheries. FAO Fish. Tech. Paper No.<br />

341, 119 p. FAO, Rome.<br />

Boord, R.L., Campbell, C.B.G., 1977. Structural <strong>and</strong> functional organization <strong>of</strong> the<br />

lateral line system <strong>of</strong> <strong>shark</strong>s. Amer. Zool. 17, 431e441.<br />

Brill, R., Bushnell, P., Smith, L., Speaks, C., Sundaram, R., Stroud, E., Wang, J., 2009.<br />

<strong>The</strong> repulsive <strong>and</strong> feeding deterrent effects <strong>of</strong> electropositive metals on juvenile<br />

s<strong>and</strong>bar <strong>shark</strong>s (Carcharhinus plumbeus). Fish. Bull. 107, 298e307.<br />

Brown, B.R., Hughes, M.E., Hutchison, J.C., 2003. Extracellular signal fluctuations in<br />

<strong>shark</strong> electrosensors. In: Fluctuations <strong>and</strong> Noise in Biological, Biophysical, <strong>and</strong><br />

Biomedical Systems. SPIE Series, vol. 5110. Bellingham, WA.<br />

Budker, P., 1971. <strong>The</strong> Life <strong>of</strong> Sharks. Weidenfeld <strong>and</strong> Nicolson, London, p. 222.<br />

Carey, F.G., Scharold, J.V., 1990. Movements <strong>of</strong> blue <strong>shark</strong>s (Prionace glauca) in depth<br />

<strong>and</strong> course. Mar. Biol. 106, 329e342.<br />

Chave, A.D., Constable, S.C., Edwards, R.N., 1991. Electrical exploration methods for<br />

the seafloor. In: Nabbighian, M.N. (Ed.), Electromagnetic Methods in Applied<br />

Geophysics, vol. 2, pp. 931e966. Tulsa.<br />

Cliff, G., Dudley, S.F.J., 1992. Protection against <strong>shark</strong> attack in South Africa, 1952-90.<br />

Aust. J. Mar. Fresh. Res. 43, 263e272.<br />

Connell, J.H., 1961. <strong>The</strong> influence <strong>of</strong> interspecific competition <strong>and</strong> other factors on<br />

the distribution <strong>of</strong> the Barnacle Chthamalus Stellatus. Ecology 42, 710e723.<br />

Connell, S.D., 1998. Effects <strong>of</strong> predators on growth, mortality <strong>and</strong> abundance <strong>of</strong><br />

a juvenile reef-fish: evidence from manipulations <strong>of</strong> predator <strong>and</strong> prey abundance.<br />

Mar. Ecol. Prog. Ser. 169, 251e261.<br />

Cos<strong>and</strong>ey-Goden, A., Morgan, A., 2011. Fisheries Bycatch <strong>of</strong> Sharks: Options for<br />

Mitigation. In: Ocean Science Series. PEW Environmental Group Tech Memo.<br />

Crombie, A.C., 1947. Interspecific competition. J. Anim. Ecol. 16 (1), 44e73.<br />

Dijkgraaf, S., Kalmijn, A.J., 1963. Untersuchungen über die Funktion der Lorenzinischen<br />

Ampullen an Haifischen. Z. Vergl. Physiol. 47, 438e456.<br />

Dijkgraaf, S., Kalmijn, A.J., 1966. Versuche zur biologischen Bedeutung der Lorenzinischen<br />

Ampullen bei den Elasmobranchiern. Z. Vergl. Physiol. 53, 187e194.<br />

Dudley, S.F.J., 1997. A comparison <strong>of</strong> the <strong>shark</strong> control programs <strong>of</strong> New SouthWales<br />

<strong>and</strong> Queensl<strong>and</strong> (Australia) <strong>and</strong> KwaZulu-Natal (South Africa). Ocean Coast.<br />

Manag. 34, 1e27.<br />

Dudley, S.F.J., Cliff, G., 1993. Some effects <strong>of</strong> <strong>shark</strong> nets in the Natal nearshore<br />

environment. Environ. Biol. Fish. 36, 243e255.<br />

Eidesmo, T., Ellingsrud, S., MacGregor, L.M., Constable, S.C., Sinha, M.C., Johansen, S.,<br />

Kong, F.N., Westerdahl, H., 2002. Sea Bed Logging (SBL), a new method for<br />

remote <strong>and</strong> direct identification <strong>of</strong> hydrocarbon filled layers in deepwater areas.<br />

News Feature 20 (3), 144e152.<br />

Gilman, E., Clarke, S., Brothers, N., Alfaro-Shigueto, J., M<strong>and</strong>elman, J., Mangel, J.,<br />

Petersen, S., Piovano, S., Thomson, N., Dalzell, P., Donoso, M., Goren, M.,<br />

Werner, T., 2007. Shark Depredation <strong>and</strong> Unwanted Bycatch in Pelagic Longline<br />

Fisheries: Industry Practices <strong>and</strong> Attitudes, <strong>and</strong> Shark Avoidance Strategies.<br />

Western Pacific Regional Fishery Management Council.<br />

Gruber, S.H., Cohen, J.L.,1985. Visual system <strong>of</strong> the white <strong>shark</strong>, Carcharodon carcharias,<br />

with emphasis on retinal structure. Mem. South. Calif. Acad. Sci. 9, 61e72.<br />

Harrington, J.M., Myers, R.A., Rosenberg, A.A., 2005. Wasted fishery resources:<br />

discarded by-catch in the USA. Fish. Fish. 6, 350e361.<br />

Holl<strong>and</strong>, K.N., Wetherbee, B.M., Lowe, C.G., Meyer, C.G., 1999. Movements <strong>of</strong> tiger<br />

<strong>shark</strong>s (Galeocerdo cuvier) in coastal Hawaiian waters. Mar. Biol. 134, 665e673.<br />

Jordan, L.J., M<strong>and</strong>elman, J.W., Kajiura, S.M., 2011. Behavioral responses to weak<br />

electric <strong>field</strong>s <strong>and</strong> a lanthanide metal in two <strong>shark</strong> species. J. Exp. Mar. Biol. Ecol.<br />

409, 345e350.<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005


10<br />

C.P. O’Connell et al. / Ocean & Coastal Management xxx (2012) 1e10<br />

Kaimmer, S., Stoner, A.W., 2008. Field investigation <strong>of</strong> rare-earth metal as a deterrent<br />

to spiny dogfish in the Pacific halibut fishery. Fish. Res. 94, 43e47.<br />

Kajiura, S.M., 2003. Electroreception in neonatal bonnethead <strong>shark</strong>s, Sphyrna tiburo.<br />

Mar. Biol. 143 (3), 603e611.<br />

Kajiura, S.M., Holl<strong>and</strong>, K.N., 2002. Electroreception in juvenile scalloped hammerhead<br />

<strong>and</strong> s<strong>and</strong>bar <strong>shark</strong>s. J. Exp. Biol. 205 (23), 3609e3621.<br />

Kalmijn, A.J., 1966. Electro-perception in <strong>shark</strong>s <strong>and</strong> rays. Nature 212, 1232e1233.<br />

Kalmijn, A.J., 1971. <strong>The</strong> electric sense <strong>of</strong> <strong>shark</strong>s <strong>and</strong> rays. J. Exp. Biol. 55, 371e383.<br />

Kalmijn, A.J., 1973. Electro-orientation in <strong>shark</strong>s <strong>and</strong> rays: theory <strong>and</strong> experimental<br />

evidence. Scripps Inst. Oceanogr. 73 (39), 1e22.<br />

Kalmijn, A.J., 1974. <strong>The</strong> detection <strong>of</strong> electric <strong>field</strong>s from inanimate <strong>and</strong> animate<br />

sources other than electric organs. In: Electroreceptors <strong>and</strong> other Specialized<br />

Receptors in Lower Vertebrates. H<strong>and</strong>book <strong>of</strong> Sensory Physiology, vol. 3.<br />

Springer-Verlag, Berlin (Germany), pp. 147e200.<br />

Kalmijn, A.J., 1981. Biophysics <strong>of</strong> geomagnetic <strong>field</strong> detection. IEEE Trans. Magn. 17,<br />

1113e1124.<br />

Kalmijn, A.J., 1982. Electric <strong>and</strong> magnetic <strong>field</strong> detection in elasmobranch fishes.<br />

Science 218 (4575), 916e918.<br />

Kalmijn, A.J., 1984. <strong>The</strong>ory <strong>of</strong> electromagnetic orientation: a further analysis. In:<br />

Keynes, R.D., Maddrell, S.H.P. (Eds.), Comparative Physiology <strong>of</strong> Sensory<br />

Systems. Cambridge University Press, pp. 525e560.<br />

Kalmijn, A.J., 1997. Electric <strong>and</strong> near-<strong>field</strong> acoustic detection, a comparative study.<br />

Acta Physiol. Sc<strong>and</strong>. 638, 25e38.<br />

Kalmijn, A.J., 2000. Detection <strong>and</strong> processing <strong>of</strong> electromagnetic <strong>and</strong> near-<strong>field</strong> acoustic<br />

signals in elasmobranch fishes. Philos. Trans. R. Soc. Lond. B 355, 1135e1141.<br />

Klimley, A.P., 1993. Highly directional swimming by scalloped hammerhead <strong>shark</strong>s,<br />

Sphyrna lewini, <strong>and</strong> subsurface irradiance, temperature, bathymetry, <strong>and</strong><br />

geomagnetic <strong>field</strong>. Mar. Biol. 117, 1e22.<br />

Klimley, A.P., Beavers, S.C., Curtis, T.H., Jorgensen, S.J., 2002. Movements <strong>and</strong><br />

swimming behavior <strong>of</strong> three species <strong>of</strong> <strong>shark</strong>s in La Jolla Canyon, California.<br />

Environ. Biol. Fish. 63, 117e135.<br />

Lessard, N., Paree, A., Lepore, F., Lassonde, M., 1998. Early-blind human subjects<br />

localize sound sources better than sighted subjects. Nature 395, 278e280.<br />

Lewison, R.L., Crowder, L.B., Read, A.J., Freeman, S.A., 2004. Underst<strong>and</strong>ing impacts<br />

<strong>of</strong> fisheries bycatch on marine megafauna. Trends Ecol. Evol. 19, 598e604.<br />

Loeb, V., Siegel, V., Holm-Hansen, O., Hewitt, R., Fraserk, W., Trivelpiecek, W.,<br />

Trivelpiecek, S., 1997. Effects <strong>of</strong> sea-ice extent <strong>and</strong> krill or salp dominance on the<br />

Antarctic food web. Nature 387, 897e900.<br />

[MAFMC] Mid-Atlantic Fishery Management Council, 1998. Spiny Dogfish Fishery<br />

Management Plan, Draft. August 1998. MAFMC [Dover, DE]. 225 pp. þ<br />

appendices.<br />

Manire, C.A., Gruber, S.H., 1990. Many <strong>shark</strong>s may be headed towards extinction.<br />

Conserv. Biol. 4, 10e11.<br />

Montgomery, J.C., Walker, M.M., 2001. Orientation <strong>and</strong> navigation in elasmobranchs:<br />

which way forward Environ. Biol. Fish. 60, 109e116.<br />

Muchnik, C., Efrati, M., Nemeth, E., Malin, M., Hildesheimer, M., 1991. Central<br />

auditory skills in blind <strong>and</strong> sighted subjects. Sc<strong>and</strong>. Audiol. 20, 19e23.<br />

Munday, P.L., Jones, G.P., Caley, M.J., 2001. Interspecific competition <strong>and</strong> coexistence<br />

in a guild <strong>of</strong> coral-dwelling fishes. Ecology 82, 2177e2189.<br />

Murray, R.W., 1957. Evidence for a mechanoreceptive function <strong>of</strong> the ampullae <strong>of</strong><br />

Lorenzini. Nature 179, 106e107.<br />

Murray, R.W., 1960. <strong>The</strong> response <strong>of</strong> the ampullae <strong>of</strong> Lorenzini <strong>of</strong> elasmobranchs to<br />

mechanical stimulation. J. Exp. Biol. 37, 417e424.<br />

Nieberding, C.M., de Vos, H., Schneider, M.V., Lassance, J.M., Estramil, N., et al., 2008.<br />

<strong>The</strong> male sex pheromone <strong>of</strong> the butterfly Bicyclus anynana: towards an<br />

evolutionary analysis. PLoS ONE 3 (7), 2751.<br />

NMFS, 2010. Final Amendment 3 to the Consolidated Atlantic Highly Migratory<br />

Species Fishery Management Plan. National Oceanic <strong>and</strong> Atmospheric Administration,<br />

National Marine Fisheries Service, Office <strong>of</strong> Sustainable Fisheries,<br />

Highly Migratory Species Management Division, Silver Spring, MD, Public<br />

Document. p. 632.<br />

O’Connell, C.P., 2008. Investigation <strong>of</strong> grade C8 barium ferrite (BaFe2O4) permanent<br />

magnets as a possible elasmobranch bycatch reduction system. In: Swimmer, Y.,<br />

Wang, J.H., McNaughton, L. (Eds.), Shark Deterrent <strong>and</strong> Incidental Capture<br />

Workshop, 10e11 April 2008, pp. 47e50. US Department <strong>of</strong> Commerce, NOAA<br />

Technical Memor<strong>and</strong>um NOAA-TM-NMFS-PIFSC-16. 72 pp.<br />

O’Connell, C.P., Gruber, S.H., Guttridge, T.L., O’Connell, T.J., Johnson, G., Grudecki, K.,<br />

<strong>and</strong> He, P. <strong>The</strong> use <strong>of</strong> permanent magnets to reduce elasmobranch encounter<br />

with a simulated beach net. 1. <strong>The</strong> Bull Shark (Carcharhinus leucas). Ocean Coast.<br />

Manag., submitted for publication.<br />

O’Connell, C.P., Abel, D.C., Rice, P.H., Stroud, E.M., Simuro, N.C., 2010. Responses <strong>of</strong><br />

the southern stingray (Dasyatis americana) <strong>and</strong> the nurse <strong>shark</strong> (Ginglymostoma<br />

cirratum) to permanent magnets. Mar. Freshw. Behav. Phy. 43, 63e73.<br />

O’Connell, C.P., Gruber, S.H., Abel, D.C., Stroud, E.M., Rice, P.H., 2011a. <strong>The</strong> responses<br />

<strong>of</strong> juvenile lemon <strong>shark</strong>s, Negaprion brevirostris, to a magnetic barrier. Ocean<br />

Coast. Manag. 54 (3), 225e230.<br />

O’Connell, C.P., Abel, D.C., Stroud, E.M., 2011b. Analysis <strong>of</strong> permanent magnets as<br />

elasmobranch bycatch reduction devices in hook-<strong>and</strong>-line <strong>and</strong> longline trials.<br />

Fish. Bull. 109 (4), 394e401.<br />

Pashley, R.M., Rzechowicz, M., Pashley, L.R., Francis, M.J., 2005. De-gassed water is<br />

a better cleaning agent. J. Phys. Chem. B 109 (3), 1231.<br />

Paulin, M.G., 1995. Electroreception <strong>and</strong> the compass sense <strong>of</strong> <strong>shark</strong>s. J. <strong>The</strong>or. Biol.<br />

174, 325e339.<br />

Polis, G.A., 1981. <strong>The</strong> evolution <strong>and</strong> dynamics <strong>of</strong> intraspecific predation. Annu. Rev.<br />

Ecol. Syst. 12, 225e251.<br />

Polovina, J.J., 1996. Decadal variation in the trans-Pacific migration <strong>of</strong> northern<br />

bluefin tuna (Thunnus thynnus) coherent with climate-induced change in prey<br />

abundance. Fish Oceanogr. 5, 114e119.<br />

Pratt, H.L., Casey, J.G., 1990. Shark reproductive strategies as a limiting factor in<br />

directed fisheries, with a review <strong>of</strong> Holden’s method <strong>of</strong> estimating growth<br />

parameters. In: Pratt Jr., H.L., Gruber, S.H., Taniuchi, T. (Eds.), Elasmobranchs as<br />

Living Resources: Advances in the Biology, Ecology, <strong>and</strong> Systematics, <strong>and</strong> the<br />

Status <strong>of</strong> the Fisheries. U.S. Dep. Commer., pp. 97e109. NOAA Tech. Rep. NMFS 90.<br />

Rago, P., Sosobee, K.A., 2009. Briefing item B: update on the status <strong>of</strong> spiny dogfish<br />

in 2009 <strong>and</strong> initial evaluation <strong>of</strong> alternative harvest strategies. In: Fisheries<br />

Management Council Memo, pp. 1e52.<br />

Rasmussen, L.E.L., Schmidt, M.J., 1992. Are <strong>shark</strong>s chemically aware <strong>of</strong> crocodiles<br />

In: Doty, R.L., Muller-Schwarze, D. (Eds.), Chemical Signals in Vertebrates IV.<br />

Plenum Press, New York, pp. 335e342.<br />

Rice, C.E., 1970. Early experience <strong>and</strong> perceptual enhancement. Res. Bull. Am. Found.<br />

Blind 22, 1e22.<br />

Rice, P., 2008. A shocking discovery: how electropositive metals (EPMs) work <strong>and</strong><br />

their effects on elasmobranchs. In: Swimmer, Y., Wang, J.H., McNaughton, L.<br />

(Eds.), Shark Deterrent <strong>and</strong> Incidental Capture Workshop, 10e11 April 2008, pp.<br />

31e35. US Department <strong>of</strong> Commerce, NOAA Technical Memor<strong>and</strong>um NOAA-<br />

TM-NMFS-PIFSC-16. 72 pp.<br />

Rigg, D.P., Peverell, S.C., Hearndon, M., Seymour, J.E., 2009. Do elasmobranch<br />

reactions to magnetic <strong>field</strong>s in water show promise for bycatch mitigation Mar.<br />

Freshw. Res. 60, 942e948.<br />

Robbins, W.D., Peddemors, V.M., Kennelly, S.J., 2011. Assessment <strong>of</strong> permanent<br />

magnets <strong>and</strong> electropositive metals to reduce the line-based capture <strong>of</strong> Galapagos<br />

<strong>shark</strong>s, Carcharhinus galapagensis. Fish. Res. 109, 100e106.<br />

Schluessel, V., Bennett, M.B., Bleckmann, H., Blomberg, S., Collin, S.P., 2008.<br />

Morphometric <strong>and</strong> ultrastructural comparison <strong>of</strong> the olfactory system in elasmobranchs:<br />

the significance <strong>of</strong> structure-function relationships based on<br />

phylogeny <strong>and</strong> ecology. J. Morphol. 269, 1365e1386.<br />

Shepherd, T.D., Myers, R.A., 2005. Direct <strong>and</strong> indirect fishery effects on small coastal<br />

elasmobranchs in the northern Gulf <strong>of</strong> Mexico. Ecol. Lett. 8, 1095e1104.<br />

Sisneros, J.S., Tricas, T.C., 2002. Ontogenetic changes in the response properties <strong>of</strong><br />

the peripheral <strong>electrosensory</strong> system in the Atlantic stingray (Dasyatis sabina).<br />

Brain Behav. Evol. 59, 130e140.<br />

Sisneros, J.A., Tricas, T.C., Luer, C.A.,1998. Response properties <strong>and</strong> biological function <strong>of</strong><br />

the skate <strong>electrosensory</strong> system during ontogeny. J. Comp. Physiol. A 183, 87e99.<br />

Stevens, J.D.R., Bonfil, R., Dulvy, N.K., Walker, P.A., 2000. <strong>The</strong> effects <strong>of</strong> fishing on<br />

<strong>shark</strong>s, rays, <strong>and</strong> chimaeras (chondrichthyans), <strong>and</strong> the implications for marine<br />

ecosystems. ICES J. Mar. Sci. 57, 476e494.<br />

Stiling, P.D., Brodbeck, B.V., Strong, D.R., 1984. Intraspecific competition in Hydrellia<br />

valida (Diptera: Ephydridae), a leaf miner <strong>of</strong> Spartina alterniflora. Ecology 65 (2),<br />

660e662.<br />

Stoner, A.W., Kaimmer, S.M., 2008. Reducing elasmobranch bycatch: laboratory<br />

investigation <strong>of</strong> rare earth metal <strong>and</strong> magnetic deterrents with spiny dogfish<br />

<strong>and</strong> Pacific halibut. Fish. Res. 92, 162e168.<br />

Stroud, E.M., 2008a. A small demonstration <strong>of</strong> rare earth galvanic cell. In: Swimmer,<br />

Y., Wang, J.H., McNaughton, L. (Eds.), Shark Deterrent <strong>and</strong> Incidental Capture<br />

Workshop, 10e11 April 2008, pp. 40e42. US Department <strong>of</strong> Commerce, NOAA<br />

Technical Memor<strong>and</strong>um NOAA-TM-NMFS-PIFSC-16. 72 pp.<br />

Stroud, E.M. 2008b. Chemical <strong>shark</strong> repellents: identifying the actives <strong>and</strong><br />

controlling their release. In: Swimmer, Y., Wang, J.H., McNaughton, L. (Eds.),<br />

Shark Deterrent <strong>and</strong> Incidental Capture Workshop, 10e11 April 2008, pp. 43e<br />

46. US Department <strong>of</strong> Commerce, NOAA Technical Memor<strong>and</strong>um NOAA-TM-<br />

NMFS-PIFSC-16. 72 pp.<br />

Stroud, E.M., O’Connell, C.P., Rice, P.H., Snow, N., Barnes, B.B., Hanson, J.E. Existence<br />

<strong>of</strong> a <strong>shark</strong> necromone derived from putrefied <strong>shark</strong> tissue. Ocean Coast. Manag.,<br />

submitted for publication.<br />

Tallack, S.M., M<strong>and</strong>elman, J.W., 2009. Do rare-earth metals deter spiny dogfish A<br />

feasibility study on the use <strong>of</strong> electropositive “mischmetal” to reduce the<br />

bycatch <strong>of</strong> Squalus acanthias by hook gear in the Gulf <strong>of</strong> Maine. ICES J. Mar. Sci.<br />

66, 315e322.<br />

Tricas, T.C., 2001. <strong>The</strong> neuroecology <strong>of</strong> the elasmobranch <strong>electrosensory</strong> world: why<br />

peripheral morphology shapes behavior. Environ. Biol. Fish. 60, 77e92.<br />

Tricas, T.C., Kajiura, S.M., Summers, A.P., 2009. Response <strong>of</strong> the hammerhead <strong>shark</strong><br />

olfactory epithelium to amino acid stimuli. J. Comp. Physiol. A 195, 947e954.<br />

Tumlinson, J.H., Turlings, T.C.J., Lewis, W.J., 1993. Semiochemically mediated<br />

foraging behavior in beneficial parasitic insects. Arch. Insect Biochem. Physiol.<br />

22, 385e391.<br />

Verheggen, F.J., Haubruge, E., Mescher, M.C., 2010. Alarm pheromones. In:<br />

Litwack, G. (Ed.), Pheromones. Elsevier, Amsterdam, <strong>The</strong> Netherl<strong>and</strong>s.<br />

von Frisch, K., 1938. <strong>The</strong> sense <strong>of</strong> hearing in fish. Nature 14, 8e11.<br />

Wang, J., McNaughton, L., Swimmer, Y., 2008. Galapagos <strong>and</strong> s<strong>and</strong>bar <strong>shark</strong> aversion<br />

to electropositive metal (Pr-Nd alloy). In: Swimmer, Y., Wang, J.H., McNaughton,<br />

L. (Eds.), Shark Deterrent <strong>and</strong> Incidental Capture Workshop, 10e11 April 2008,<br />

pp. 31e35. US Department <strong>of</strong> Commerce, NOAA Technical Memor<strong>and</strong>um NOAA-<br />

TM-NMFS-PIFSC-16. 72 pp.<br />

Williamson, C.E., 1980. <strong>The</strong> predatory behavior <strong>of</strong> mesocyclops edax: predator<br />

preferences, prey defenses, <strong>and</strong> starvation-induced changes. Limnol. Oceanogr.<br />

25 (5), 903e909.<br />

WWF, 2006. World Wildlife Fund Smart Gear Competition [Internet]. World<br />

Wildlife Fund, Washington (DC). Available from: http://www.smartgear.org/<br />

smartgearwinners/smartgearwinner2006/smartgearwinner2006gr<strong>and</strong>/index.<br />

cfm [cited 2006].<br />

Please cite this article in press as: O’Connell, C.P., et al., <strong>The</strong> <strong>emerging</strong> <strong>field</strong> <strong>of</strong> <strong>electrosensory</strong> <strong>and</strong> <strong>semiochemical</strong> <strong>shark</strong> repellents: Mechanisms <strong>of</strong><br />

detection, overview <strong>of</strong> past studies, <strong>and</strong> future directions, Ocean & Coastal Management (2012), http://dx.doi.org/10.1016/<br />

j.ocecoaman.2012.11.005

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!