Biochemical mechanisms of biomineralization and elemental incorporation in otoliths: implications for fish and fisheries research

Oliver R. B. Thomas https://orcid.org/0000-0002-8497-670X Submitted in total fulfilment of the requirements of the degree of Doctor of Philosophy December 2018 School of BioSciences The University of Melbourne

1

ABSTRACT

All vertebrates have small bioinorganic “earstones” in their inner ear labyrinth that are essential for hearing and balance. While otoliths play a vital anatomical role in fish, their true value to science is as biochronometers, largely due to their unique pattern of growth. Otoliths first form in embryo and continue to grow throughout the life of an individual, with a double-banded increment composed of a calcium carbonate-rich region and a protein-rich region being deposited daily. In addition to this, they are considered to be metabolically inert, and do not undergo remodelling or resorption. Consequently, otoliths are employed in a variety of ways in fish ecology. Firstly, an individual fish’s age and growth rate can be estimated through counting increments and measuring their widths. Secondly, analysis of increment trace element:calcium ratios, such as by laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-MS), can allow for the reconstruction of environmental histories, aiding in the determination of natal origin, movement, habitat use, diet and the impacts of climate change. The utility of specific trace elements as indicators of environmental change, however, is unclear as there is considerable uncertainty as to whether a given trace element is interacting with the mineral or protein components of an increment. This uncertainty is a consequence of otolith research having been largely focussed upon either microstructure or inorganic chemistry, with very few studies on the protein-rich regions of the otolith. As a result, very little is understood about the biochemical mechanisms of biomineralization or trace element incorporation. This is important, as the mechanisms that govern otolith formation and growth underpin the assumptions made in traditional increment analyses. In this thesis, I initially undertook a systematic review of all the literature pertaining to otolith biochemistry, revealing the significant gaps that exist in otolith biochemistry as a discipline. Importantly, I determined that fewer than a score of otolith proteins had been identified – a stark contrast to the hundreds or thousands of proteins that have been

i

identified in comparable biomineral systems such as enamel or bone. Working on black bream

(Acanthopagrus butcheri), an extensively studied species endemic to southern Australia, I used size exclusion chromatography coupled with ICP-MS to determine the trace element:protein interactions in endolymph, the inner ear fluid that otoliths are submerged in, and the source of all of its constituents. In this study, I assayed 22 elements, and determined that 12 were solely present in a protein-bound form, 6 were present as free ions, and 4 were present in both forms.

This allowed me to make recommendations as to their utility in environmental reconstructions. In my next study, I created a unique, multi-disciplinary workflow that combined transcriptomics with proteomics. In this study, I sequenced the transcriptome of the black bream inner ear and used this to identify proteins from the separated organic phase of otoliths and endolymph from wild caught adult black bream. This resulted in the discovery of hundreds of previously unknown proteins, providing new insights into the likely biochemical mechanisms involved in otolith formation and growth. In my final study, I tested the utility of trace element ratios in environmental reconstructions. Specifically, I compared the ability of different cluster analysis approaches to resolve spatial and temporal differences in the likely spawning and larval nursery habitats of juvenile black bream in the Gippsland Lakes, Australia. The results from my thesis have allowed me to make recommendations as to the utility of trace elements in environmental reconstructions and have revealed exciting new avenues of research that fuse ecology and biochemistry.

ii

DECLARATION

This is to certify that:

The thesis comprises only my original work towards the PhD except where otherwise indicated.

Due acknowledgement has been made in the text to all other material used.

The thesis is fewer than 100 000 words in length, exclusive of tables, maps, bibliographies and appendices.

Oliver R. B. Thomas

December 2018

Cover image: Sagittal otolith from juvenile black bream (Acanthopagrus butcheri), D. Chamberlain, 2016

iii

PREFACE

I am the primary author and principle contributor on all chapters presented in this thesis. My primary supervisor, Stephen E. Swearer, is a co-author on all chapters. My supervisor Blaine R. Roberts is a co-author on Chapters 3 and 4. My third supervisor, Gregory P. Jenkins is a co-author on Chapter 5.

Article publication status and co-author contributions

Chapter Two: Unpublished material not yet submitted for publication. Co-authored by Stephen E. Swearer. ORBT and SES conceived and designed the review; ORBT read, assessed and selected manuscripts to be included in the review; ORBT wrote the manuscript; SES provided editorial comments.

Contributions: ORBT 85%, SES 15%

Chapter Three: Published in Metallomics on March 2017. Co-authored by Stephen E. Swearer, Katherine Ganio and Blaine R. Roberts. BRR, SES and ORBT conceived and designed the experiments; ORBT conducted the experiments and collected data with assistance from KG and BRR; ORBT analysed the data and wrote the manuscript; SES, KG and BRR provided editorial comments.

Contributions: ORBT 75%, KG 5%, BRR 10%, SES 10%

Chapter Four: Accepted for publication in FEBS Journal, December 2018. Co-authored by Stephen E. Swearer, Eugene A. Kapp, Po Peng, Gerry Q. Tonkin-Hill, Anthony Papenfuss, Anne Roberts, Pascal Bernard and Blaine R. Roberts. ORBT, BRR and SES conceived and designed the experiments. ORBT conducted the experiments and collected data with assistance from GQT, AP, AR and PB. ORBT, EAK and PP analysed data and designed workflows. ORBT wrote the manuscript. SES, EAK and BRR provided editorial comments.

Contributions: ORBT 75%, SES 5%, EAK 5%, PP 3%, GQT 3%, AP 1%, AR 2%, PB 1%, BRR 5%

Chapter Five: Unpublished material not submitted for publication. Co-authored by Gregory P. Jenkins, Katherine V. Thomas and Stephen E. Swearer. ORBT, GPJ and SES conceived and designed the study. ORBT conducted the experiments and collected data with assistance from GPJ and SES. KVT provided GIS expertise. ORBT wrote the manuscript. SES and GPJ provided editorial comments.

Contributions: ORBT 75%, GPJ 10%, KVT 5%, SES 10%

iv

This research was funded by grants from the Holsworth Wildlife Research Endowment (Chapter Five), the Australian Research Council (Chapters Three, Four and Five), the Research on the Ecology and Evolution (REEF) Lab (all chapters), and the University of Melbourne’s John and Allen Gilmour Award (Chapter Four). All research was conducted in accordance with the animal ethics requirements of the University of Melbourne (Chapters Three, Four, Five). Permits were obtained from the Victorian state government for collection of fish (Chapters Three, Four and Five; RP 1204, 1135 and 1141).

v

ACKNOWLEDGMENTS

Firstly, a huge thank you to my supervisors for all the support and friendship that they have provided over the past years. Greg, I have really appreciated your insight, advice and guidance in the field! Blaine, thank you for taking the plunge and joining the world of fish inner ears and, in turn, welcoming me into the Metalloproteomics group. Things would’ve been very different if you hadn’t agreed to collaborate with us! Steve, thank you for taking the gamble on me back in 2013. You’re a great supervisor, mentor and friend. I owe you much.

My colleagues and mates in BioSciences, the Florey and the REEF, NCCC and Metalloproteomics laboratory groups (past and present) provided great advice and feedback on many of my manuscripts and presentations. Not only have they been fantastically collegial, they’ve also been stalwart friends! Of special note are those that made me feel part of everything: Po, Luke, Rob, Valeriya, Cleves, Sievers, Joao, Juanma, Matt, Fletch, Dean, Nicole, Kath, Jess, Eric, Soumya, Stephan, Larissa, Katie, Anne and Eugene.

Many people were generous in sharing their time with me, volunteering to help me in the laboratory or the field: Dean Chamberlain, Luke Barrett, Michael Sievers, Jessica French, Kathryn Hassell, Anne Roberts, Soumya Mukherjee, Akiva Gebler, Ryan Woodland, Fiona Warry, David Brehm. Apologies if I’ve omitted anyone here! A special mention goes to my collaborators Po Peng and Eugene Kapp who co-designed a bioinformatics workflow with me and provided consistent support and friendship.

My family and friends have been hugely supportive of me, giving me the space to write and think…and talk, talk, talk. My sister, Kath, provided expert advice and counsel. (She also made some fantastic maps!) My children, Jasmine, Frederick and Dashiell have all inadvertently become experts on fish ears through patiently listening to me. My mother and father, first introduced me to the amazing marine world through SCUBA – thank you!

Finally, to Carolyn, my best friend. I couldn’t have done this without you. You’ve made more sacrifices than anyone so that I could follow my passion. Thank you.

vi

DEDICATION

For Carolyn

vii

CONTENTS

List of Tables x

List of Figures xi

Chapter One | General Introduction

General Introduction 1

References 11

Chapter Two | Otolith Biochemistry – A review

Abstract 25

Introduction 26

Otolith Formation and Growth 29

Otolith Chemistry 36

The Otolith Organic Matrix 39

Models of Otolith Growth 72

Implications and Future Directions 77

Conclusion 80

References 81

Chapter Three | Trace element–protein interactions in endolymph from the inner ear of fish: implications for environmental reconstructions using fish otolith chemistry

Abstract 111

Introduction 112

Methods 116

Results and Discussion 121

Conclusion 131

References 133

viii

Chapter Four | The inner ear proteome of fish

Abstract 143

Introduction 143

Results 148

Discussion 154

Conclusion 163

Materials and Methods 164

References 172

Chapter Five | Spatio-temporal resolution of spawning and larval nursery habitats using otolith microchemistry is element dependent

Abstract 183

Introduction 184

Materials and Methods 188

Results 196

Discussion 216

Conclusion 226

References 227

Chapter Six | General Discussion

General discussion 245

Future directions 251

References 258

Appendices

Appendix 3.1. 265

Appendix 3.2. 277

Appendix 4.1. 281

Appendix 5.1. 282

Appendix 5.2. 284

ix

LIST OF TABLES

Table 2.1. Summary of proteins directly identified in otoliths or endolymph of fish 44

Table 2.2. Summary of trace elements found in otoliths and endolymph of fish 79

Table 3.1. Typical operating parameters of SEC-ICP-MS 119

Table 3.2. Average concentration, standard deviation, range, coefficient of variation, and percentage enrichment of 22 trace elements from paired otolith and endolymph samples from adult Acanthopagrus butcheri (n=3), determined by inductively coupled plasma-mass spectrometry. 122

Table 4.1. Comparison of unique peptides occurring in starting material versus

RP-HPLC fractions for pooled otolith and endolymph samples from A. butcheri. 148

Table 4.2. Results of BLASTX v2.60+ alignment of A. butcheri cDNA FASTA with

RefSeq proteins. 148

Table 4.3. Parameters employed for Mascot searches of proteomic data. 170

Table 5.1. Summaries of the 2015 cluster analyses of LA-ICP-MS data from sagittal otoliths extracted from juvenile black bream (A. butcheri). 202

Table 5.2. Summaries of the 2016 cluster analyses of LA-ICP-MS data from sagittal otoliths extracted from juvenile black bream (A. butcheri). 203

x

LIST OF FIGURES

Figure 1.1. Schematic of inner ear of cichlid, Steatocranus tinanti. 2

Figure 2.1. Distribution of saccular epithelial cells in the inner ear of teleost fish. 30

Figure 2.2. Results of keyword searches using Web of Science. 40

Figure 3.1. Size exclusion ICP-MS chromatograms of endolymph. 126

Figure 4.1. Schematic of workflow for creation of a custom species-specific proteomics sequence database. 147

Figure 4.2. Comparison of protein distributions, based on unique peptides, occurring in otolith and endolymph samples from adult black bream. 149

Figure 4.3. All proteins, based on proteotypical peptides, found in the otoliths of adult black bream. 150

Figure 4.4. Heat map of selected proteins occurring in the inner ear of black bream. 151

Figure 4.5. Constellation plot illustrating the clustering of inner ear proteins from a range of fish species. 153

Figure 4.6. Schematic of hypothetical ion transport mechanisms in the inner ear of black bream. 155

Figure 4.7. Proposed regulation of diel-entrained organic matrix deposition. 161

Figure 4.8. Proteins common to bone and other biomineralization systems found in the inner ear of adult black bream. 162

xi

Figure 5.1. Fishing sites for sampling black bream, Gippsland Lakes estuary;

Feb-Apr, 2015 and 2016. 190

Figure 5.2. Schematic illustrating how LA-ICP-MS data were used as observations in the cluster analyses of juvenile black bream otoliths. 195

Figure 5.3. Size frequency distribution of juvenile black bream caught in

Lake King, Victoria, Australia over February, March and April of 2015 and 2016. 197

Figure 5.4. Pie charts showing the number and monthly proportion of juvenile black bream caught per standard haul in 2015 and 2016. 199

Figure 5.5. Pie charts showing proportional distribution of clusters based on otolith primordium chemistry for black bream for the first and second cohort for 2015. 205

Figure 5.6. Pie charts showing proportional distribution of clusters based on otolith larval period chemistry for black bream for the first and second cohort for 2015. 207

Figure 5.7. Comparison of dominant clusters based on otolith chemistry in the larval period and primordium for black bream caught in Lake King, Australia, 2015. 209

Figure 5.8. Pie charts showing proportional distribution of clusters based on otolith primordium chemistry for black bream for the first and second cohort for 2016. 211

Figure 5.9. Pie charts showing proportional distribution of clusters based on otolith larval period chemistry for black bream for the first and second cohort for 2016. 213

Figure 5.10. Comparison of dominant clusters based on otolith chemistry in the larval period and primordium for black bream caught in Lake King, Australia, 2016. 215

xii

Figure 5.11. Interannual comparison of otolith larval chemistry for juvenile black bream caught in Lake King, Australia, in 2015 and 2016. 224

Figure 6.1. Expression of Matrix metalloproteinase 2 in the inner ear of zebrafish (Danio rerio). 252

Figure 6.2. Z-stack of CLARITY-treated otolith from adult black bream. 254

Figure 6.3. MALDI images of major tissues of the juvenile black bream eye. 256

xiii

CHAPTER ONE: GENERAL INTRODUCTION

In the labyrinth of the inner ear of all vertebrates are small, biomineral structures. These structures – called either otoliths (in fish) or otoconia (in other taxa) – are composed of calcium carbonate and various structural and accessory biomolecules (Carlstrom, 1963). Otoliths and otoconia are fundamental to an organism’s sense of balance as they allow for the fine-detail perception of gravity (Tanturri, 1926). The mechanism of gravity perception (for both otoliths and otoconia) is largely the same, irrespective of taxa. The otolith (or otoconia) is in a fluid-filled compartment of the inner ear, where it is positioned against sensory hairs (kinocilia). These compartments (along with their contained otoliths/otoconia) are often referred to as “otolithic end organs”. As an organism experiences a shift in acceleration (due to gravity, or, in the case of fish, a change to water movement in the water column), the otolith/otoconia moves against the kinocilia, producing shearing force, which is detected by the otic nerve (Popper et al., 1982, Ross and Pote, 1984, Popper and Saidel, 1994, Popper and Fay, 2011).

In terrestrial vertebrates, the otoconia are usually a paste of calcite “eardust” (Pote and Ross,

1991). In contrast to this, in fish, otoliths are relatively large “earstones” that can be either aragonite (Mugiya et al., 1981), calcite (Gauldie, 1993, Oliveira et al., 1996), vaterite (Campana,

1983b, Gauldie, 1986, Oliveira et al., 1996), or a mix of polymorphs (Tzeng et al., 2007). Calcite, aragonite and vaterite, as polymorphs of calcium carbonate, differ in crystal geometry: calcite – rhombohedral; aragonite – orthorhombic; vaterite – hexagonal (de Leeuw and Parker, 1998).

Interestingly, precipitation of aragonite, the prevalent polymorph in fish otoliths, is favoured at temperatures above 30°C in vitro (Kinsman and Holland, 1969, Ogino et al., 1987) yet aragonite precipitates in the fish inner ear at lower temperatures (Bath et al., 2000, Takagi, 2002). It is presumed that the organic constituents of the otolith are responsible for facilitating aragonite

1 precipitation at this temperature, but the specifics are unclear (Söllner et al., 2003, Tohse et al.,

2009, Wojtas and Dobryszycki, 2011).

In fish, there are three pairs of otoliths that differ in size and location within the inner ear. The largest (usually) are the sagittae and are found in the saccula. Sagittae tend to be aragonite

(Carlstrom, 1963, Degens et al., 1969, Mann et al., 1983, Morales-Nin, 1986b, Oliveira et al., 1996), though they can have vaterite impurities, which may cause impairment of hearing and gravity detection (Sweeting et al., 2004, Tomas et al., 2004, Reimer et al., 2016). Sagittae have been the focus of the majority of ecological studies (Campana and Neilson, 1985, Campana, 1999). The next largest are the lapilli, found in the utriculi , which also are often aragonite (Campana, 1999). The smallest otoliths, the asterisci, are found in the lagenae, and are usually vaterite (Lowenstam,

1989, Oliveira et al., 1996, Li and Feng, 2007). Although the size hierarchy tends to be sagittae>lapilli>asterisci, this can vary between taxa (Secor et al., 1992). A representation of the inner ear labyrinth of a bony fish is shown in Fig. 1.1.

Figure 1.1. Schematic of inner ear of cichlid, Steatocranus tinanti. Semicircular canals: anterior (asc), horizontal (hsc), and posterior (psc) canals; ampulla of the anterior (aa), horizontal (ha), and posterior (pa) canals; cc, common canal (= crus commune); (2) otolith end organs: utricle, saccule, lagena; (3) sensory epithelia: cristae of the anterior (ca), horizontal (ch), posterior (cp) canals; maculae of the utricle (mu), saccule (ms), and lagena (ml); (4) otoliths: uot, utricular otolith (= lapillus), sot, saccular otolith (= sagitta), and lot, lagenar otolith (= asteriscus). a, anterior; d, dorsal. Reprinted from Schulz-Mirbach et al. (2018).

2

In fish early development, otoliths are among the first structures to form. In zebrafish, for example, otolith formation occurs as early as 24 hours post fertilisation (Riley et al., 1997).

Following formation of the primordium, (the otolith core, which is rich in proteins and glycogen), alternating protein-rich and mineral-rich bands are deposited each day (Pannella, 1971). The structure of otolith increments is that of a layer of rod-shaped aragonite crystals perpendicularly placed against an intermeshed organic layer, often referred to as the “otolith organic matrix”

(Dunkelberger et al., 1980, Zhang, 1992). In addition to the daily cycle of increment deposition, otoliths are both acellular and metabolically inert (Campana and Neilson, 1985) and do not undergo resorption due to acidity (Geen et al., 1985) or exertion (Campana, 1983a). They have been found to be reabsorbed only under extreme anaerobic stress (Mugiya and Uchimura, 1989).

Furthermore, otoliths continue to grow even in periods of starvation (Marshall and Parker, 1982,

Campana, 1983b). Currently there are no clear explanations for the mechanisms underpinning its continual cycle of daily increment deposition or its metabolic inertness.

A useful consequence of increment deposition and metabolic inertness, however, is that otoliths can be treated as archives of chronological information. This is desirable in ecological research: the bands can be counted, allowing for estimation of a fish’s age (Campana and Neilson, 1985); the varying widths of individual, or groups of, bands can be measured to provide a useful proxy for growth (Brothers et al., 1976); and remarkably, the change in ratios of trace elements to calcium across increments can be used to reconstruct environmental chronologies (Campana, 1999, Izzo et al., 2016).

Fish are generally in motion throughout their lives, either swimming throughout a given waterway or travelling across a salinity gradient. As a given individual fish changes its physical location, the water content with regards to minor and trace elements will also change depending on environmental conditions (Campana and Thorrold, 2001). Otolith increments have been observed

3 to have small, but significant, concentrations of elements other than calcium, carbon or oxygen present (Campana, 1999, Thorrold and Swearer, 2009, Izzo et al., 2015). The prevailing view in otolith research is that the fluctuations of trace element amounts in the water column are

“recorded” by the otolith. It was suggested by Campana (1999) that there are four main routes that trace elements can take with respect to incorporation: substitution for calcium (or carbonate) in the aragonite lattice; as a mineral impurity (such as witherite, strontianite); trapping in the interstitial spaces between crystals; and interaction with the otolith organic components. The specifics of trace element incorporation, however, are still largely unknown. To date, only one study has demonstrated a definitive mechanism of incorporation. Using analysis of extended X-ray absorption fine structure (EXAFS), Doubleday et al. (2014) demonstrated that strontium directly substitutes for calcium in the aragonite lattice. The route by which a trace element enters a fish is similarly unclear. Trace elements may enter the blood plasma either through branchial uptake in the gut or gills (Payan et al., 2004), or as part of diet (Watanabe et al., 1997, Ranaldi and Gagnon,

2008b). They must then be transported from the plasma, through the saccular epithelial cells, into the endolymph – although the mechanisms by which this happens have not been elucidated. Once in the endolymph, a trace element can potentially be incorporated into an otolith increment as it forms.

Despite the mechanisms of uptake and incorporation for individual elements being still largely unresolved, otolith trace element:calcium ratios have nevertheless been used to create a variety of environmental reconstructions. They have been employed in: determination of larval dispersal

(Swearer et al., 1999, Brophy and Danilowicz, 2002); detection of migratory pathways (Thresher et al., 1994, Limburg and Siegel, 2006, Miles et al., 2009); reconstruction of temperature chronologies (Hoie et al., 2004, Dorval et al., 2011); as a means of tagging individuals (Campana et

4 al., 1994, Warren-Myers et al., 2014); determination of natal source (Thorrold et al., 2001, Barbee and Swearer, 2007); and, as indicators of habitat use (Thorrold et al., 1998).

Thorrold and Swearer (2009) found that, in the majority of otolith microchemistry studies, the organic material was not removed prior to analysis, and that most researchers opted to simply rinse their otoliths in ultrapure water. As we currently are unaware of the specifics by which elements bind to the organic phase, the presence of organic material during analyses likely adds unwanted variability to datasets, reducing confidence in any predictions made. It is therefore vital to determine which trace elements are indicative of water chemistry, and which exist as necessary constituents (i.e. cofactors) of organic molecules within the otolith organic matrix.

There are considerable difficulties associated with determining metal partitioning with regards to the organic component of an otolith. In order to examine any macromolecules (and determine trace elements, if any, associated with them), the organic material must be first separated from the inorganic matrix. Extraction of the organic constituents is a simple enough task. One can choose to remove the crystalline phase through dissolution in either Ethylenediaminetetraacetic acid (EDTA) (Asano and Mugiya, 1993), or other acids (Davies et al., 2011)). Conversely, the organic material can be removed through employing a non-specific lyase such as Proteinase-K (Shiao et al.,

1999). Unfortunately, all of these methods result in the decoupling of trace elements from the proteinaceous portion of the otolith, destroying any information regarding the specifics of trace element:organic interactions. For example, EDTA is not solely a calcium-chelator. Whilst its binding affinity for calcium is very high, it will also form complex ions with other metals such as magnesium or manganese, effectively ripping these elements away from any proteins they are associated with. Similarly, acid-dissolution or etching based approaches involve the disruption of tertiary structure, a mechanism that will almost certainly cause the dissociation of any loosely bound elements from any proteins present. An alternative approach exists, however, for

5 determining metal:protein interactions in the otolith. If otolith proteins are identified, we can then determine their interactions with trace elements due to structural similarity with known vertebrate proteins.

Unfortunately, we currently have limited knowledge about the protein components of the otolith, and their associated trace metal interactions. There have been some early studies focussed on the structural anatomy and biochemistry of otoliths and endolymph, as well as attempts to quantify and qualify the proteinaceous material (Degens et al., 1969, King, 1978, Söllner et al., 2003, Miller et al., 2006, Tohse et al., 2008, Weigele et al., 2016), the content of endolymph (Mugiya and

Takahashi, 1985, Kalish, 1991, Payan et al., 1999, Borelli et al., 2001, Tohse and Mugiya, 2001), and, the factors necessary for biogenic crystallisation of aragonite (Takagi, 2002, Jolivet et al.,

2008, Melancon et al., 2008). The findings of some of these studies, however, were limited by the experimental methods available to them. Although it is now possible to sequence hundreds of proteins present in a sample, the first otolith protein to be sequenced, inner ear-specific collagen

(now called Otolin-1), was the result of an exhaustive study (Davis et al., 1995). Similarly, while early estimations of fluctuations in protein abundances were determined through low resolution bulk analyses (Degens et al., 1969, Mugiya and Takahashi, 1985, Morales-Nin, 1986a), it is possible now, with the advances in separation techniques and mass spectrometry, to obtain far finer resolution. As a direct consequence of this, over the past twenty years only roughly a dozen proteins have been identified in the otolith (Lundberg et al., 2015). This is starkly contrasted by the thousands of proteins documented in osteoblasts (Xiao et al., 2007) and hundreds in enamel

(Charone et al., 2016, Salmon et al., 2017, Lima Leite et al., 2018).

Fortunately, the last two decades has seen an explosion in the number of techniques available for the identification of proteins and other organics. Fields of note here are proteomics, transcriptomics, and metalloproteomics. Proteomics is the study of the function, structure and

6 localization of proteins – referred to as the “proteome” (Shi and Chance, 2008). In proteomics, proteins are digested to fragments (peptides). Peptides are then used to identify proteins present through being aligned with large peptide or gene databases (Zhang et al., 2013). In regards to determining its organic matrix, an otolith can be dissolved in a suitable reagent (e.g. trichloracetic acid), the proteins extracted and subsequently digested into constituent peptides. These peptides are then aligned (e.g. with BLAST, MASCOT) against a peptide database. However, this method,

(“bottom-up” proteomics) only provides evidence for the existence of a protein and does not necessarily allow for the determination of the entirety of its primary sequence. The ability for a researcher to “see” a peptide is influenced by whether it is produced from trypsin (or similar) digestion, as well as its readiness to ionize within the mass spectrometer (Nesvizhskii et al., 2007,

Bantscheff et al., 2007). It is desirable then, to work with the sequenced genome, as it allows for the determination of protein primary sequences. Unfortunately, very few fish species have had their genomes sequenced (https://www.ncbi.nlm.nih.gov/genome, accessed 10 December 2018).

A possible solution is to make use of transcriptomics: the sequencing and assembly of the transcriptome – that portion of the genome that directly codes for proteins (Velculescu et al.,

1997). This approach produces a species-specific search database and allows for the determination of otolith protein primary sequences. Entire sequences can then be aligned with known proteins to determine homology, as well as act as an essential resource in both genetic manipulation and expression studies on inner ear proteins. Finally, metalloproteomics specifically refers to investigations that combine traditional proteomics with metallomics – the thorough and total understanding of metals and metalloids within cells and tissues (Szpunar, 2005).

Metalloproteomics, then, involves the unravelling of the explicit and implicit relationships between biologically occurring metals and proteins (Lothian et al., 2013) – essential for determining the specifics of trace element interactions in otoliths. In short, the use of proteomics,

7 transcriptomics and their related disciplines can result in powerful outcomes for otolith research: key biomineralization proteins can be determined and their function inferred, hypotheses can be justified regarding the otolith’s unique properties (e.g. diel-entrained increment growth, aragonite selection, metabolic inertness), and one can identify otolith metalloproteins and thus be able to discern which trace elements are indicative of endogenous processes, and which are indicative of exogeny.

THESIS AIMS AND STRUCTURE

Otoliths are frequently utilised in a variety of ways in fish and fisheries ecology (e.g. in stock assessments, growth estimations, environmental reconstructions) yet the biochemical mechanisms that underpin the basic assumptions of otolith microstructure and microchemical analyses are largely unknown. The aims of this thesis are to increase our current knowledge of the biochemistry of otolith formation, growth and trace element incorporation and use this to guide ecological interpretations of otolith data.

I worked with black bream, Acanthopagrus butcheri, an estuarine species that has been extensively studied in terms of its otolith chemistry (Elsdon and Gillanders, 2002, Elsdon and

Gillanders, 2003, Elsdon and Gillanders, 2004, de Vries et al., 2005, Elsdon and Gillanders, 2005a,

Elsdon and Gillanders, 2005b, Elsdon and Gillanders, 2005c, Elsdon and Gillanders, 2006, Ranaldi and Gagnon, 2008a, Webb et al., 2012, Doubleday et al., 2014, Williams et al., 2018). Black bream is a southern Australian fish that is relatively long-lived, abundant, and occupies its natal estuary for the majority of its life (Butcher, 1945, Butcher et al., 1962).

Chapters 2-5 were written as standalone manuscripts, with minor changes to fit the thesis format.

Consequently, some repetition of fundamental concepts – particularly as pertains to the use of otoliths in fish ecological research – has been unavoidable.

8

In Chapter Two, to determine the current state of knowledge relating to otolith biochemistry, I conducted a systematic review. In this review, I examined over 9000 manuscripts resulting from a keyword search for “otolith* and (chemistry or biochemistry or protein* or structure)” on Web of

Science (http://www.webofknowledge.com) and filtered these to 218 studies relating to otolith biochemistry. I describe the biochemical factors relating to otolith growth, otolith chemistry and the otolith organic matrix, with a special emphasis on the various proteins involved in the latter. In addition to describing the proteins in the otolith, I also describe the suggested mechanisms for otolith nucleation, growth and polymorph selection, comment on the utility of trace elements in otolith research and suggest potential lines of future otolith scientific research.

Chapter Three was previously published in Metallomics. The intention of this study was to determine the specifics of trace element:protein interactions in the otolith, and use these to inform future ecological studies. As described above, the dissolution of otoliths in traditional reagents (e.g. EDTA, acids) results in the unwanted decoupling of trace elements from organic constituents. However, as the otolith gets all of its constituents – both organic and inorganic – from endolymph (Payan et al., 2004), any trace element:protein interactions occurring in the otolith are also likely to be present in this fluid. To answer this question, I used size exclusion chromatography-inductively coupled plasma mass spectrometry (SEC-ICP-MS) to determine the trace element:protein interactions occurring in endolymph of adult black bream. The basic idea was that, as fractions from a given biological tissue exit the separation medium (i.e. a size exclusion gel chromatography column), they then enter into an ICP-MS instrument and are detected for metals. As a result, one can then link individual metals to specific protein molecular weight fractions. I assayed 22 elements and found that 12 elements are only present bound to proteins, 6 are present in the salt phase, and 4 are present in both salt and protein-bound forms.

9

Chapter Four is currently in press, FEBS Journal. In this study, I determined the proteome of both the otolith and endolymph of black bream. Proteomics approaches typically involve the use of model organisms such as Homo sapiens, Mus musculus or Danio rerio – which have all had their genomes sequenced. In a proteomic approach, as stated above, peptides are matched with large peptide or genome databases. Previous proteomic studies on otoliths have been either on D. rerio

(Söllner et al., 2003, Kaplon et al., 2008, Petko et al., 2008, Yariz et al., 2012), or made use of large databases (such as ) for alignment of peptides (Weigele et al., 2016). This proves a significant challenge for the study of non-model organisms like black bream as the false discovery rate (FDR) for peptides is increased. More importantly, large numbers of mass spectra end up being unassigned, causing numerous novel proteins to be overlooked. Normally then, the genome would be sequenced. This, however, is prohibitive in terms of time and cost. To deal with this issue, I created a multi-omics workflow that allowed for identification of candidate inner ear proteins from black bream. My approach first involved the sequencing, assembly and annotation of its inner ear and brain transcriptome. I then used reverse-phase fractionation coupled with tandem mass spectrometry (RP-HPLC-MS/MS) to identify peptides in otoliths and endolymph from wild juvenile and adult black bream and matched these to its transcriptome. I subsequently created a species-specific peptide search database for conducting quantitative proteomics, which was used to uncover the proteome of the black bream inner ear. In this study, I developed a comprehensive protein landscape, increasing the number of identified otolith proteins by a factor of 30 and additionally identified several novel proteins – many of which would have not been identified using a traditional proteomics approach. The results of my proteomics study allowed me to come up with likely explanations for several unresolved aspects of otolith biomineralization: inorganic constituent supply to the endolymph; the selection of aragonite as the calcium carbonate polymorph; diel-entrained increment deposition; and, its metabolic inertness.

10

In Chapter Five, I investigated the utility of trace element ratios obtained by laser ablation- inductively coupled plasma-mass spectrometry (LA-ICP-MS) in creating physico-chemical environmental reconstructions. I used the results of my previous studies (Chapters Two and Three) to explicitly investigate how the choice of elements influences the predictive power of multi- elemental signatures in resolving spatio-temporal differences in otolith chemistry. Specifically, I assessed the performance of Ward’s hierarchical clustering using three different approaches: (1) all elements analysed through LA-ICP-MS; (2) elements identified as being primarily in the salt- fraction of the endolymph; and, (3) elements identified as being associated with the protein- fraction of endolymph. Following this, I assessed the predictive power of my clustering approaches through multinomial logistic regression. I then used these results to make inferences relating to natal source and dispersal of juvenile black bream in the Gippsland Lakes, Australia for the years

2015 and 2016. My findings were consistent with previous hypotheses relating to black bream juvenile recruitment but also revealed a previously unknown spawning group, suggesting an unexpected source of black bream larvae.

Finally, Chapter Six provides a general discussion of this body of work, summarizing the key findings. I conclude by discussing a number of fruitful lines of research available to otolith scientists.

REFERENCES

Asano, M., Mugiya, Y. (1993) Biochemical and calcium-binding properties of water-soluble proteins

isolated from otoliths of the tilapia, Oreochromis niloticus. Comparative Biochemistry and

Physiology Part B: Comparative Biochemistry 104, 201-205.

11

Bantscheff, M., Schirle, M., Sweetman, G., Rick, J., Kuster, B. (2007) Quantitative mass

spectrometry in proteomics: a critical review. Analytical and Bioanalytical Chemistry 389,

1017-1031. [In English].

Barbee, N.C., Swearer, S.E. (2007) Characterizing natal source population signatures in the

diadromous fish Galaxias maculatus, using embryonic otolith chemistry. Marine Ecology

Progress Series 343, 273-282.

Bath, G.E., Thorrold, S.R., Jones, C.M., Campana, S.E., McLaren, J.W., Lam, J.W.H. (2000) Strontium

and barium uptake in aragonitic otoliths of marine fish. Geochimica et Cosmochimica Acta

64, 1705-1714.

Borelli, G., Mayer-Gostan, N., De Pontual, H., Boeuf, G., Payan, P. (2001) Biochemical relationships

between endolymph and otolith matrix in the trout (Oncorhynchus mykiss) and turbot

(Psetta maxima). Calcified Tissue International 69, 356-364.

Brophy, D., Danilowicz, B.S. (2002) Tracing populations of Atlantic herring (Clupea harengus L.) in

the Irish and Celtic Seas using otolith microstructure. ICES Journal of Marine Science 59,

1305-1313.

Brothers, E.B., Mathews, C.P., Lasker, R. (1976) Daily growth increments in otoliths from larval and

adult fishes. Fishery Bulletin 74, 1-8. [In English].

Butcher, A.D. (1945) The Gippsland Lakes bream fishery. Australian Fisheries Newsletter 4(3).

Butcher, A.D., Ling, J.K., Victoria, Fisheries, Wildlife, D. (1962) Bream tagging experiments in East

Gippsland during April and May 1944, Vol., Fisheries and Wildlife Dept., Victoria,

[Melbourne].

12

Campana, S.E. (1983a) Calcium deposition and otolith check formation during periods of stress in

coho salmon, Oncorhynchus kisutch. Comparative Biochemistry and Physiology a-

Physiology 75, 215-220. [In English].

Campana, S.E. (1983b) Feeding periodicity and the production of daily growth increments in

otoliths of steelhead trout (Salmo gairdneri) and starry flounder (Platichthys stellatus).

Canadian Journal of Zoology-Revue Canadienne De Zoologie 61, 1591-1597. [In English].

Campana, S.E. (1999) Chemistry and composition of fish otoliths: pathways, mechanisms and

applications. Marine Ecology Progress Series 188, 263-297.

Campana, S.E., Fowler, A.J., Jones, C.M. (1994) Otolith elemental fingerprinting for stock

identification of Atlantic cod (Gadus morhua) using Laser-Ablation ICP-MS. Canadian

Journal of Fisheries and Aquatic Sciences 51, 1942-1950. [In English].

Campana, S.E., Neilson, J.D. (1985) Microstructure of fish otoliths. Canadian Journal of Fisheries

and Aquatic Sciences 42, 1014-1032. [In English].

Campana, S.E., Thorrold, S.R. (2001) Otoliths, increments, and elements: keys to a comprehensive

understanding of fish populations? Canadian Journal of Fisheries and Aquatic Sciences 58,

30-38.

Carlstrom, D.D. (1963) Crystallographic study of vertebrate otoliths. Biological Bulletin 125, 441-&.

[In English].

Charone, S., De Lima Leite, A., Peres-Buzalaf, C., et al. (2016) Proteomics of secretory-stage and

maturation-stage enamel of genetically distinct mice. Caries Research 50, 24-31. [In

English].

13

Davies, C.A., Brophy, D., Jeffries, T., Gosling, E. (2011) Trace elements in the otoliths and dorsal

spines of albacore tuna (Thunnus alalunga, Bonnaterre, 1788): An assessment of the

effectiveness of cleaning procedures at removing postmortem contamination. Journal of

Experimental Marine Biology and Ecology 396, 162-170.

Davis, J.G., Oberholtzer, J.C., Burns, F.R., Greene, M.I. (1995) Molecular cloning and

characterization of an inner ear-specific structural protein. Science 267, 1031-1034. de Leeuw, N.H., Parker, S.C. (1998) Surface structure and morphology of calcium carbonate

polymorphs calcite, aragonite, and vaterite: An atomistic approach. Journal of Physical

Chemistry B 102, 2914-2922. [In English]. de Vries, M.C., Gillanders, B.M., Elsdon, T.S. (2005) Facilitation of barium uptake into fish otoliths:

Influence of strontium concentration and salinity. Geochimica et Cosmochimica Acta 69,

4061-4072.

Degens, E., Deuser, W., Haedrich, R. (1969) Molecular structure and composition of fish otoliths.

Marine Biology 2, 105.

Dorval, E., Piner, K., Robertson, L., Reiss, C.S., Javor, B., Vetter, R. (2011) Temperature record in the

oxygen stable isotopes of Pacific sardine otoliths: Experimental vs. wild stocks from the

Southern California Bight. Journal of Experimental Marine Biology and Ecology 397, 136-

143.

Doubleday, Z.A., Harris, H.H., Izzo, C., Gillanders, B.M. (2014) Strontium randomly substituting for

calcium in fish otolith aragonite. Analytical Chemistry 86, 865-869.

Dunkelberger, D.G., Dean, J.M., Watabe, N. (1980) The ultrastructure of the otolithic membrane

and otolith in the juvenile mummichog, Fundulus heteroclitus. Journal of Morphology 163,

367-377. [In English].

14

Elsdon, T.S., Gillanders, B.M. (2002) Interactive effects of temperature and salinity on otolith

chemistry: challenges for determining environmental histories of fish. Canadian Journal of

Fisheries and Aquatic Sciences 59, 1796-1808.

Elsdon, T.S., Gillanders, B.M. (2003) Relationship between water and otolith elemental

concentrations in juvenile black bream Acanthopagrus butcheri. Marine Ecology Progress

Series 260, 263-272.

Elsdon, T.S., Gillanders, B.M. (2004) Fish otolith chemistry influenced by exposure to multiple

environmental variables. Journal of Experimental Marine Biology and Ecology 313, 269-

284.

Elsdon, T.S., Gillanders, B.M. (2005a) Alternative life-history patterns of estuarine fish: barium in

otoliths elucidates freshwater residency. Canadian Journal of Fisheries and Aquatic

Sciences 62, 1143-1152. [In English].

Elsdon, T.S., Gillanders, B.M. (2005b) Consistency of patterns between laboratory experiments and

field collected fish in otolith chemistry: an example and applications for salinity

reconstructions. Marine and Freshwater Research 56, 609-617.

Elsdon, T.S., Gillanders, B.M. (2005c) Strontium incorporation into calcified structures: separating

the effects of ambient water concentration and exposure time. Marine Ecology Progress

Series 285, 233-243.

Elsdon, T.S., Gillanders, B.M. (2006) Identifying migratory contingents of fish by combining otolith

Sr : Ca with temporal collections of ambient Sr : Ca concentrations. Journal of Fish Biology

69, 643-657.

Gauldie, R.W. (1986) Vaterite otoliths from chinook salmon (Oncorhynchus tshawytscha). New

Zealand Journal of Marine and Freshwater Research 20, 209-217. [In English].

15

Gauldie, R.W. (1993) Polymorphic crystalline-structure of fish otoliths. Journal of Morphology 218,

1-28. [In English].

Geen, G.H., Neilson, J.D., Bradford, M. (1985) Effects of pH on the early development and growth

and otolith microstructure of chinook salmon, Oncorhynchus tshawytscha. Canadian

Journal of Zoology-Revue Canadienne De Zoologie 63, 22-27.

Hoie, H., Otterlei, E., Folkvord, A. (2004) Temperature-dependent fractionation of stable oxygen

isotopes in otoliths of juvenile cod (Gadus morhua L.). ICES Journal of Marine Science 61,

243-251.

Izzo, C., Doubleday, Z.A., Grammer, G.L., et al. (2016) Fish as proxies of ecological and

environmental change. Reviews in Fish Biology and Fisheries 26, 265-286. [In English].

Izzo, C., Doubleday, Z.A., Schultz, A.G., Woodcock, S.H., Gillanders, B.M. (2015) Contribution of

water chemistry and fish condition to otolith chemistry: comparisons across salinity

environments. Journal of Fish Biology 86, 1680-1698.

Jolivet, A., Bardeau, J.-F., Fablet, R., Paulet, Y.-M., de Pontual, H. (2008) Understanding otolith

biomineralization processes: new insights into microscale spatial distribution of organic

and mineral fractions from Raman microspectrometry. Analytical and Bioanalytical

Chemistry 392, 551-560.

Kalish, J.M. (1991) Determinants of otolith chemistry - seasonal-variation in the composition of

blood-plasma, endolymph and otoliths of bearded rock cod Pseudophycis barbatus.

Marine Ecology Progress Series 74, 137-159. [In English].

Kaplon, T.M., Rymarczyk, G., Nocula-Lugowska, M., et al. (2008) Starmaker exhibits properties of

an intrinsically disordered protein. Biomacromolecules 9, 2118-2125.

16

King, K., Jr. (1978) Distribution of gamma-carboxyglutamic acid in calcified tissues. Biochimica et

Biophysica Acta 542, 542-546.

Kinsman, D.J.J., Holland, H.D. (1969) Co-precipitation of cations with CaCO3 .4. Co-precipitation of

Sr2+ with aragonite between 16 degrees and 96 degrees C. Geochimica et Cosmochimica

Acta 33, 1-&. [In English].

Li, Z., Feng, Q. (2007) Analysis of polymorphs of carp's otoliths. Rare Metal Materials and

Engineering 36, 47-49.

Lima Leite, A., Silva Fernandes, M., Charone, S., Whitford, G.M., Everett, E.T., Buzalaf, M.A.R.

(2018) Proteomic mapping of dental enamel matrix from inbred mouse strains: Unraveling

potential new players in enamel. Caries Research 52, 78-87. [In English].

Limburg, K.E., Siegel, D.I. (2006) The hydrogeochemistry of connected waterways: The potential of

linking geology to fish migrations. Northeastern Geology and Environmental Sciences 28,

254-265.

Lothian, A., Hare, D.J., Grimm, R., Ryan, T.M., Masters, C.L., Roberts, B.R. (2013)

Metalloproteomics: principles, challenges and applications to neurodegeneration.

Frontiers in Aging Neuroscience 5, 35. [In English].

Lowenstam, H.A. (1989) On biomineralization / Heinz A. Lowenstam, Stephen Weiner, Vol.

Accessed from https://nla.gov.au/nla.cat-vn1683340), Oxford University Press, New York.

Lundberg, Y.W., Xu, Y., Thiessen, K.D., Kramer, K.L. (2015) Mechanisms of otoconia and otolith

development. Developmental Dynamics 244, 239-253.

Mann, S., Parker, S.B., Ross, M.D., Skarnulis, A.J., Williams, R.J.P. (1983) The ultrastructure of the

calcium carbonate balance organs of the inner-ear - an ultrahigh resolution electron-

17

microscopy study. Proceedings of the Royal Society Series B-Biological Sciences 218, 415-+.

[In English].

Marshall, S.L., Parker, S.S. (1982) Pattern identification in the microstructure of sockeye salmon

(Oncorhynchus nerka) otoliths. Canadian Journal of Fisheries and Aquatic Sciences 39, 542-

547.

Melancon, S., Fryer, B.J., Gagnon, J.E., Ludsin, S.A. (2008) Mineralogical approaches to the study of

biomineralization in fish otoliths. Mineralogical Magazine 72, 627-637.

Miles, N.G., West, R.J., Norman, M.D. (2009) Does otolith chemistry indicate diadromous lifecycles

for five Australian riverine fishes? Marine and Freshwater Research 60, 904-911.

Miller, M.B., Clough, A.M., Batson, J.N., Vachet, R.W. (2006) Transition metal binding to cod otolith

proteins. Journal of Experimental Marine Biology and Ecology 329, 135-143.

Morales-Nin, B. (1986a) Chemical composition of the otoliths of the sea bass Dicentrarchus labrax

pisces serranidae. Cybium 10, 115-120.

Morales-Nin, B.Y.O. (1986b) Structure and composition of otoliths of cape hake Merluccius

capensis. South African Journal of Marine Science, 3-10.

Mugiya, Y., Takahashi, K. (1985) Chemical properties of the saccular endolymph in the rainbow

trout Salmo gairdneri. Bulletin of the Faculty of Fisheries Hokkaido University 36, 57-63. [In

English].

Mugiya, Y., Uchimura, T. (1989) Otolith resorption induced by anaerobic stress in the goldfish,

Carassius auratus. Journal of Fish Biology 35, 813-818. [In English].

18

Mugiya, Y., Watabe, N., Yamada, J., Dean, J.M., Dunkelberger, D.G., Shimizu, M. (1981) Diurnal

rhythm in otolith formation in the goldfish, Carassius auratus. Comparative Biochemistry

and Physiology a-Physiology 68, 659-662. [In English].

Nesvizhskii, A.I., Vitek, O., Aebersold, R. (2007) Analysis and validation of proteomic data

generated by tandem mass spectrometry. Nature Methods 4, 787-797. [In English].

Ogino, T., Suzuki, T., Sawada, K. (1987) The formation and transformation mechanism of calcium

carbonate in water. Geochimica et Cosmochimica Acta 51, 2757-2767.

Oliveira, A.M., Farina, M., Ludka, I.P., Kachar, B. (1996) Vaterite, calcite, and aragonite in the

otoliths of three species of piranha. Naturwissenschaften 83, 133-135.

Pannella, G. (1971) Fish otoliths: daily growth layers and periodical patterns. Science (New York,

N.Y.) 173, 1124-1127.

Payan, P., De Pontual, H., Boeuf, G., Mayer-Gostan, N. (2004) Endolymph chemistry and otolith

growth in fish. Comptes Rendus Palevol 3, 535-547.

Payan, P., Edeyer, A., De Pontual, H., Borelli, G., Boeuf, G., Mayer-Gostan, N. (1999) Chemical

composition of saccular endolymph and otolith in fish inner ear: lack of spatial uniformity.

American Journal of Physiology-Regulatory Integrative and Comparative Physiology 277,

R123-R131.

Petko, J.A., Millimaki, B.B., Canfield, V.A., Riley, B.B., Levenson, R. (2008) Otoc1: A novel Otoconin-

90 ortholog required for otolith mineralization in zebrafish. Developmental Neurobiology

68, 209-222.

Popper, A.N., Fay, R.R. (2011) Rethinking sound detection by fishes. Hearing Research 273, 25-36.

19

Popper, A.N., Platt, C., Saidel, W.M. (1982) Acoustic functions in the fish ear. Trends in

Neurosciences 5, 276-280. [In English].

Popper, A.N., Saidel, W.M. (1994) Sensory cells of the fish ear: Puzzle of hairy cells. Sensornye

Sistemy 8, 29-41.

Pote, K.G., Ross, M.D. (1991) Each otoconia polymorph has a protein unique to that polymorph.

Comparative Biochemistry and Physiology B-Biochemistry & Molecular Biology 98, 287-

295. [In English].

Ranaldi, M.M., Gagnon, M.M. (2008a) Trace metal incorporation in otoliths of black bream

(Acanthopagrus butcheri munro), an indicator of exposure to metal contamination. Water

Air and Soil Pollution 194, 31-43. [In English].

Ranaldi, M.M., Gagnon, M.M. (2008b) Zinc incorporation in the otoliths of juvenile pink snapper

(Pagrus auratus Forster): The influence of dietary versus waterborne sources. Journal of

Experimental Marine Biology and Ecology 360, 56-62.

Reimer, T., Dempster, T., Warren-Myers, F., Jensen, A.J., Swearer, S.E. (2016) High prevalence of

vaterite in sagittal otoliths causes hearing impairment in farmed fish. Scientific Reports 6,

25249. [In English].

Riley, B.B., Zhu, C.W., Janetopoulos, C., Aufderheide, K.J. (1997) A critical period of ear

development controlled by distinct populations of ciliated cells in the zebrafish.

Developmental Biology 191, 191-201. [In English].

Ross, M.D., Pote, K.G. (1984) Some properties of otoconia. Philosophical Transactions of the Royal

Society of London Series B-Biological Sciences 304, 445-&.

20

Salmon, C.R., Giorgetti, A.P.O., Leme, A.F.P., et al. (2017) Microproteome of dentoalveolar tissues.

Bone 101, 219-229. [In English].

Schulz-Mirbach, T., Ladich, F., Plath, M., Hess, M. (2018) Enigmatic ear stones: what we know

about the functional role and evolution of fish otoliths. Biological Reviews of the

Cambridge Philosophical Society 0.

Secor, D., Dean, J.M., Laban, E.H. (1992) Otolith removal and preparation for microstructural

examination. Otolith microstructure examination and analysis. Canadian special

publication of fisheries and aquatic sciences 117, 19-57.

Shi, W., Chance, M.R. (2008) Metallomics and metalloproteomics. Cellular and Molecular Life

Sciences 65, 3040-3048. [In English].

Shiao, J.C., Tzeng, C.S., Leu, C.L., Chen, F.C. (1999) Enhancing the contrast and visibility of daily

growth increments in fish otoliths etched by Proteinase K buffer. Journal of Fish Biology

54, 302-309.

Söllner, C., Burghammer, M., Busch-Nentwich, E., et al. (2003) Control of crystal size and lattice

formation by Starmaker in otolith biomineralization. Science 302, 282-286.

Swearer, S.E., Caselle, J.E., Lea, D.W., Warner, R.R. (1999) Larval retention and recruitment in an

island population of a coral-reef fish. Nature 402, 799-802. [In English].

Sweeting, R.M., Beamish, R.J., Neville, C.M. (2004) Crystalline otoliths in teleosts: Comparisons

between hatchery and wild coho salmon (Oncorhynchus kisutch) in the Strait of Georgia.

Reviews in Fish Biology and Fisheries 14, 361-369.

21

Szpunar, J. (2005) Advances in analytical methodology for bioinorganic speciation analysis:

metallomics, metalloproteomics and heteroatom-tagged proteomics and metabolomics.

Analyst 130, 442-465. [In English].

Takagi, Y. (2002) Otolith formation and endolymph chemistry: a strong correlation between the

aragonite saturation state and pH in the endolymph of the trout otolith organ. Marine

Ecology Progress Series 231, 237-245.

Tanturri, V. (1926) The labyrinthine reflexes of position. Morgagni 68, 161-166.

Thorrold, S.R., Jones, C.M., Swart, P.K., Targett, T.E. (1998) Accurate classification of juvenile

weakfish Cynoscion regalis to estuarine nursery areas based on chemical signatures in

otoliths. Marine Ecology Progress Series 173, 253-265.

Thorrold, S.R., Latkoczy, C., Swart, P.K., Jones, C.M. (2001) Natal homing in a marine fish

metapopulation. Science 291, 297-299.

Thorrold, S.R., Swearer, S.E. (2009) Otolith Chemistry. In: Tropical Fish Otoliths: Information for

Assessment, Management and Ecology. Vol. 11. Reviews-Methods and Technologies in Fish

Biology and Fisheries (Eds. B.S. Green, B.D. Mapstone, G. Carlos, G.A. Begg), pp. 249-295.

Thresher, R.E., Proctor, C.H., Gunn, J.S., Harrowfield, I.R. (1994) An evaluation of electron-probe

microanalysis of otoliths for stock delineation and identification of nursery areas in a

southern temperature groundfish, Nemadactylus macropterus (Cheilodactylidae). U S

National Marine Fisheries Service Fishery Bulletin 92, 817-840.

Tohse, H., Mugiya, Y. (2001) Effects of enzyme and anion transport inhibitors on in vitro

incorporation of inorganic carbon and calcium into endolymph and otoliths in salmon

Oncorhynchus masou. Comparative Biochemistry and Physiology a-Molecular and

Integrative Physiology 128, 177-184.

22

Tohse, H., Saruwatari, K., Kogure, T., Nagasawa, H., Takagi, Y. (2009) Control of polymorphism and

morphology of calcium carbonate crystals by a matrix protein aggregate in fish otoliths.

Crystal Growth & Design 9, 4897-4901.

Tohse, H., Takagi, Y., Nagasawa, H. (2008) Identification of a novel matrix protein contained in a

protein aggregate associated with collagen in fish otoliths. FEBS Journal 275, 2512-2523.

Tomas, J., Geffen, A.J., Allen, I.S., Berges, J. (2004) Analysis of the soluble matrix of vaterite otoliths

of juvenile herring (Clupea harengus): Do crystalline otoliths have less protein?

Comparative Biochemistry and Physiology a-Molecular & Integrative Physiology 139, 301-

308.

Tzeng, W.-N., Chang, C.-W., Wang, C.-H., et al. (2007) Misidentification of the migratory history of

anguillid eels by Sr/Ca ratios of vaterite otoliths. Marine Ecology Progress Series 348, 285-

295.

Velculescu, V.E., Zhang, L., Zhou, W., et al. (1997) Characterization of the yeast transcriptome. Cell

88, 243-251.

Warren-Myers, F., Dempster, T., Fjelldal, P.G., Hansen, T., Jensen, A.J., Swearer, S.E. (2014) Stable

isotope marking of otoliths during vaccination: a novel method for mass-marking fish.

Aquaculture Environment Interactions 5, 143-154.

Watanabe, T., Kiron, V., Satoh, S. (1997) Trace minerals in fish nutrition. Aquaculture 151, 185-207.

[In English].

Webb, S.D., Woodcock, S.H., Gillanders, B.M. (2012) Sources of otolith barium and strontium in

estuarine fish and the influence of salinity and temperature. Marine Ecology Progress

Series 453, 189-199.

23

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2016) Not all inner ears are the same: Otolith matrix

proteins in the inner ear of sub-adult cichlid fish, Oreochromis mossambicus, reveal

insights into the biomineralization process. Anatomical Record-Advances in Integrative

Anatomy and Evolutionary Biology 299, 234-245.

Williams, J., Jenkins, G.P., Hindell, J.S., Swearer, S.E. (2018) Fine-scale variability in elemental

composition of estuarine water and otoliths: Developing environmental markers for

determining larval fish dispersal histories within estuaries. Limnology and Oceanography

63, 262-277. [In English].

Wojtas, M., Dobryszycki, P. (2011) Hyperphosphorylation of intrinsically disordered Starmaker

protein increases its biomineralization activity. FEBS Journal 278, 482-482.

Xiao, Z., Camalier, C.E., Nagashima, K., et al. (2007) Analysis of the extracellular matrix vesicle

proteome in mineralizing osteoblasts. Journal of Cellular Physiology 210, 325-335. [In

English].

Yariz, K.O., Duman, D., Zazo Seco, C., et al. (2012) Mutations in OTOGL, encoding the inner ear

protein Otogelin-like, cause moderate sensorineural hearing loss. American Journal of

Human Genetics 91, 872-882.

Zhang, Y., Fonslow, B.R., Shan, B., Baek, M.C., Yates, J.R., 3rd (2013) Protein analysis by

shotgun/bottom-up proteomics. Chemical Reviews 113, 2343-2394. [In English].

Zhang, Z. (1992) Ultrastructure of otolith increments and checks in the teleost fish Oreochromis

niloticus. Journal of Morphology 211, 213-220. [In English].

24

CHAPTER TWO: OTOLITH BIOCHEMISTRY – A REVIEW

ABSTRACT

Otoliths are bioinorganic minerals within the inner ear labyrinth of all bony fishes. Otoliths grow incrementally, laying down alternating protein-rich and mineral rich bands daily. As they accrete for the entire life of an individual, otoliths are considered natural chronometers, archiving the growth and physico-chemical histories of individual fish. They are thus frequently employed in fisheries stock assessments, age and growth estimations, as well as proxies for movement, habitat use and the impact of climate change. The majority of otolith research has been focussed in these areas, either through the measurement of otolith increments for ageing studies, or the analysis of trace-element:calcium ratios for environmental reconstructions. In contrast, little attention has been paid to the proteinaceous constituents of the otolith, their mechanisms of incorporation or their implications for otolith formation, growth, chemical composition and function. To provide a comprehensive picture of the otolith organic matrix, I reviewed 218 papers relating to otolith biochemistry. Here I provide a comprehensive summary of the current state-of-knowledge with regards to the biochemistry of the endolymph (inner ear fluid), saccular epithelial tissues, and the otolith, with particular attention to the specific proteins present. I discuss the structure and function of 28 proteins and 5 proteoglycans identified thus far in the fish inner ear. I additionally comment on the appropriateness and utility of particular trace elements as environmental proxies, considering the likelihood that they occur primarily as metalloproteins in the otolith organic matrix, and suggest fruitful lines of future research in otolith science.

25

INTRODUCTION

Otoliths play a vital physiological role in bony fish – the reception of sound and gravity through the detection of changes in ambient water pressure (Popper et al., 1982). Specifically, the otolith interacts with hair cells (kinocilia) on the macula that grow through a membrane that lies in a groove (sulcus acusticus) on the otolith face (Popper et al., 2005). As the otolith moves against the kinocilia, it produces shearing forces, that are detected by mechanoreceptors (Lundberg et al.,

2006). Indeed, all vertebrates have calcified structures present in the inner ear labyrinth that are employed in this way for gravity sensation (Lowenstein, 1936). In each inner ear labyrinth are otolith end organs, the sacculum, the lagena, and the utriculus – three compartments filled with endolymph, a fluid rich in organic and inorganic components essential for otolith growth (Mugiya and Takahashi, 1985, Kalish, 1989, Kalish, 1991, Romanek and Gauldie, 1996, Payan et al., 1997,

Payan et al., 1999, Borelli et al., 2001). While amphibians and fish have all three compartments, in mammals only the saccula and utriculi are present. Fish also differ from other vertebrates in the size and number of biomineralized inner ear structures. In fish, there are three pairs of relatively large acellular “earstones”, termed the sagittae, the lapilli and asterisci otoliths, that respectively interact with cilia of the sensory maculae of the saccula, utriculi and lagenae (Popper et al., 1982,

Campana and Neilson, 1985). In contrast to otoliths, in other vertebrates an “eardust” paste composed of numerous tiny otoconia crystals serves the same function (Popper et al., 2005).

Although otoconia and otoliths are all calcium carbonate biominerals, their crystal geometry differ.

Otoconia are made of calcite (trigonal), sagittae and lapilli tend to be aragonite (orthorhombic), and asterisci are often vaterite (hexagonal). It is generally assumed that the differences in crystal polymorph of biominerals are driven by the biochemistry of their associated macromolecules

(Weiner and Addadi, 1997, Falini et al., 2005, Reimer et al., 2017). Consequently, knowledge of

26 otolith biochemistry is fundamental to understanding the relationships between an otolith’s shape and morphology and its function (Reimer et al., 2016).

The initial suggestion that the hard parts of fish might be used for determining age is ancient (it was first suggested by Aristotle), but the development of an actual quantitative method, for otoliths, however, was not articulated until Reibisch published his seminal work in 1899 (Aristotle et al., ca. 340BCE (transl. 1965), Reibisch, 1899). Further important methodological developments occurred throughout the 20th century; early explorations of the organic and inorganic components of otoliths were undertaken (Dannevig, 1956, Irie, 1960, Degens et al., 1969, Mugiya, 1974,

Dunkelberger et al., 1980, Mugiya, 1987), and most importantly, a novel method linking trace element ratios (specifically Sr:Ca) with daily increments was trialled and developed (Radtke, 1985).

In the 21st century, the measurement of otolith increments, in conjunction with the determination of their associated trace element:calcium ratios, has become an essential and powerful tool in the reconstruction of fish life histories (Limburg, 1995, Campana and Thorrold, 2001, Elsdon et al.,

2008, Sturrock et al., 2012, Izzo et al., 2016b). As otolith microstructural and microchemical data are so frequently and broadly employed, it is important to understand the biochemical mechanisms underpinning these observations. Only with proper knowledge of otolith protein machinery can these biochronologies be correctly interpreted.

The intention of this review is to provide an overview of what macromolecules have been identified in otoliths, describe what their likely role is in otolith formation and growth based on their function in other biominerals and discuss the implications of this for otolith research on hearing, microstructure, and chemistry. To this end, I first present an overview of the structures of the saccular epithelium and the otolith membrane, particularly in how they relate to the supply of inorganic and organic components necessary for otolith growth. Following this, I describe the composition of the endolymph, followed by a comprehensive summary and discussion of each of

27 the main protein types present in the otolith organic matrix, with special reference to their structure (if known) and role in biomineralization. Finally, as there are currently only a few studies on the specifics of trace element interactions with the organic components of otoliths or endolymph (Miller et al., 2006, Izzo et al., 2016a, Thomas et al., 2017), I discuss the implications of these interactions in the context of reconstructing environmental histories using otolith microchemistry.

In order to review all available literature relating to the biochemistry of otoliths, I searched the

Web of Science (http://www.webofknowledge.com) on April 19, 2016 (and repeated again on

September 17, 2018), using the search string “otolith* and (chemistry or biochemistry or protein* or structure)”. This search yielded over 9000 entries. As a first pass, I reviewed all titles, disqualifying those papers that were solely concerned with human health or microstructural analyses (e.g. ageing). I additionally discarded papers focussed on biochronology reconstructions, unless specific reference was made to the otolith organic matrix. I also discarded conference abstracts. The second pass of assessment involved reading the abstract for each study, ensuring that papers to be included provided information with regards to the biochemical mechanisms and organic constituents of the inner ear. Finally, I searched the reference lists of included papers to identify any potentially missed studies. This process resulted in a total of 218 studies included in the review.

Many of the studies reviewed were undertaken without the benefit of the large protein and nucleotide databases that exist today. Consequently, I performed Basic Local Alignment Search

Tool (BLAST) analyses on available otolith protein entries from the National Center for

Biotechnology Information (NCBI). By undertaking BLAST analyses, I was able to identify areas of conservation between proteins as well as any protein domains of interest. BLAST is a powerful bioinformatics tool that compares a target primary sequence (e.g. nucleotide or amino acid) with a

28 library of all known sequences. From the result of a BLAST analysis, any homologous proteins may be identified along with a measurement of the significance of their alignment with the target.

OTOLITH FORMATION AND GROWTH

The most striking feature about otoliths, particularly the sagittae (which are typically the largest in most taxa) is their mechanism of growth – the daily formation of alternating mineral-rich and protein-rich bands (Pannella, 1971, Jolivet et al., 2008). The calcium carbonate-dominant band is often referred to as the “incremental zone”, whereas the protein-rich band is called the

“discontinuous zone” (Dunkelberger et al., 1980, Mugiya et al., 1981). The specific biochemical mechanisms that underpin the apparently circadian process of alternating deposition of these two zones have not yet been fully elucidated.

An understanding of the structure of the sacculum and its epithelial cells is therefore vital to the construction of a comprehensive model of otolith growth. As the otolith is acellular, it is likely that all necessary components for its growth are secreted by specialized epithelial cells into the endolymph in the extracellular space. The epithelium of the sacculum can be described as having four distinct areas: the sensory zone (macula), a meshwork zone of large ionocytes flanking the macula, a transitional zone, and a zone containing discrete groupings of smaller ionocytes facing the macula (the “patches” zone, sometimes referred to as the “squamous epithelia”) (Davis et al.,

1995b, Mayer-Gostan et al., 1997, Pisam et al., 1998, Takagi and Takahashi, 1999, Weigele et al.,

2017). Protein secretion has been detected throughout the saccular epithelium – in hair cells, supporting cells, transitional epithelial cells and squamous cells (Mugiya, 1968, Mugiya, 1974,

Baba et al., 1991, Takagi and Takahashi, 1999, Takagi, 2000, Stooke-Vaughan et al., 2015, Weigele

29 et al., 2017). Figure 2.1, from Pisam et al. 1998, shows the distribution of cells in the saccular epithelium.

Figure 2.1. Distribution of saccular epithelial cells in the inner ear of teleost fish showing the “patches”, “intermediate” and “meshwork” areas. The macula is also indicated. Here, the macula has numerous hair cells (HC), that interact with nerve endings (NE). The macula also contains supporting cells (SC) and granular cells (GC). The “meshwork area” contains large ionocytes (LI) whereas the “patches area” has small ionocytes (SI). Reprinted from Pisam et al., 1998.

The macula contains hair cells and supporting cells. It was initially thought that the hair cells and supporting cells were restricted, respectively, to the apical and basal epithelial layers of the macula. However, Weigele et al. (2016a) found two different types of supporting cells in the inner ear of larval O. mossambicus. While Type I supporting cells were found solely in the basal layer,

Type II were elongated and protruded into the apical layer, being able to interface with the endolymph. In a study conducted by Pisam et al. (1998), “granular” cells were identified on the periphery of the macula that were implicated in protein secretion to the endolymph. It is likely that these “granular” cells and Type II supporting cells are the same. It has also been observed,

30 through scanning electron microscopy, that protein strands exude from the macula (Sokolowski,

1986). Weigele et al. (2016a) determined that otolith matrix proteins are expressed throughout the saccular epithelium, with two (Otolin-1 and OMP-1) being co-expressed in cells in the ventral meshwork zone, and a further three (otogelin, alpha tectorin and Sparc) being chiefly expressed in the macula. These findings are consistent with similar studies of rainbow trout (Murayama et al.,

2004), flatfish (Paralichthys lethostigma) (Wang et al., 2011) and zebrafish (Stooke-Vaughan et al.,

2015).

In addition to protein secretion, the meshwork and “patches” zones are also likely to be involved in ion supply to the endolymph. It has been suggested, considering the differences in cell size and distribution, that the two different zones have different functions (e.g. K+ supply to the endolymph, Ca2+/H+ exchange; (Mayer-Gostan et al., 1997)). Both the meshwork zone and the

“patches” zone have been found to be rich in mitochondria as well as containing Na+/K+-ATPase and isozymes of carbonic anhydrase (Mayer-Gostan et al., 1997, Takagi, 1997).

The otolith interfaces with the hair cells on the macula through an acellular gelatinous membrane

(Dunkelberger et al., 1980, Sokolowski, 1986, Khan and Drescher, 1990, Benser et al., 1993,

Hughes et al., 2004). This membrane consists of two discrete areas, a structured gelatinous layer and an unstructured subcupular meshwork area. The former extends from the groove along the inner face (sulcus acusticus) to the hair cells of the macula. It lies over roughly half of the proximal, or inner, face of the otolith, with pores in the gelatinous layer corresponding to protrusions in the sulcus acusticus (Dunkelberger et al., 1980). In contrast to this, the subcupular meshwork is made up of a network of fibres that extend from the apical epithelium to the otolith. These fibres are far more abundant in the nonsensory epithelium, and are incorporated into the otolith organic matrix

31

(Dunkelberger et al., 1980). Dunkelberger et al. (1980) suggested that the subcupular meshwork was more likely to be the functional zone than the gelatinous layer.

In mammals, the otolith membrane is fairly well characterised, with several of its constituent proteins – Otolin, Otogelin, Otogelin-like, Alpha- and Beta-tectorin – having been identified

(Goodyear and Richardson, 2002, Deans et al., 2010, Yariz et al., 2012). Unsurprisingly, homologous proteins have also been identified in the teleostean otolith membrane. For instance,

Otolin-1 has been identified in the otolith membrane of bluegill sunfish (Davis et al., 1995a), rainbow trout (Murayama et al., 2004) and zebrafish (Murayama et al., 2005) through immunohistochemistry. In addition to this, Alpha-tectorin, as well as otogelin and otogelin-like have been implicated in the membrane through knockdown studies (Yariz et al., 2012, Stooke-

Vaughan et al., 2015). Unlike the avian and mammalian inner ear, however, Beta-tectorin appears to be purely expressed in the macula, and is not a component of the otolith membrane in fish

(Yang et al., 2011). The details of these proteins, as well as the chaperone, GP96, are described below.

Otoliths first form in embryo as a cluster of precursor particles that are tethered to the saccular epithelial tissue (Riley et al., 1997). These particles contain both glycoproteins and glycogen, and eventually form the primordium (Pisam et al., 2002, Beier and Anken, 2006, Dauphin and Dufour,

2008, Stooke-Vaughan et al., 2015). Analyses of cores of otoliths have indicated that the primordia are almost purely composed of organic constituents which control the direction of mineral precipitation (Mann et al., 1983, Jolivet et al., 2008). After initial formation of the primordium, biomineralization commences, with alternating protein-rich and mineral-dominant increments being laid down over a 24h period. The expression of otolith proteins as well as increment growth have been demonstrated to be dependent on thyroid hormone action (Shiao and Hwang, 2004,

32

Wang et al., 2006, Shiao et al., 2008, Schreiber et al., 2010, Wang et al., 2011, Coffin et al., 2012).

Due to this, and the continual accretion of increment bands throughout a fish’s life, increment number and widths can be used to determine the age of an individual as well as changes in somatic growth (Campana and Neilson, 1985). The validation of the deposition of annual and daily increments is a vital part of both effective fisheries management and age and growth studies

(Devries and Frie, 1996). However, as the mechanisms underpinning increment deposition are poorly understood, current ageing studies rely on assumptions gleaned from otolith marking studies (Campana, 2001).

Several studies have sought to link endolymph protein concentration with otolith daily growth.

Vertebrate endolymph, as a body fluid, however, is comparable to blood serum with regards to number, diversity and concentrations of proteins present. Indeed, Thomas et al. (2019) found the presence of hundreds of different proteins in the endolymph, with the most abundant (for example Apolipoprotein A1, (ApoA1)) not necessarily playing functional roles in biomineralization.

Studies interpreting endolymph protein concentration fluctuations with regards to increment formation should therefore be viewed with care (Kalish, 1991, Payan et al., 1998, Payan et al.,

1999, Edeyer et al., 2000). These studies typically quantified bulk protein concentrations, so it is unclear whether a concentration change is due to changes in expression of otolith matrix proteins, particularly as these are orders of magnitude lower than non-biomineralization proteins (Thomas et al., 2019). Bearing this in mind, there are observed differences in protein concentration in endolymph during the night:day period, with protein levels being generally higher during the former. From this, as well as elevated daytime CO2 concentrations, along with immunohistochemical studies of otolith matrix proteins, it is generally believed that the proteinaceous band begins forming at dusk and continues throughout the night, whereas aragonite precipitation occurs at dawn (Mugiya, 1987, Zhang and Runham, 1992, Borelli et al.,

33

2003a, Takagi et al., 2005, Tohse et al., 2006). An inversion of this has been observed in some studies, however, suggesting that the timing of incremental and discontinuous zone formation may be species specific (Edeyer et al., 2000, Parmentier et al., 2007). The fluctuations in levels of

CO2 and protein expression in the endolymph encouraged investigations into carbonic anhydrase.

Carbonic anhydrase (CA) isozymes are key proteins in otolith growth (Mugiya et al., 1979, Mugiya and Takahashi, 1985, Mayer-Gostan et al., 1997, Tohse and Mugiya, 2001, Beier et al., 2002,

Borelli et al., 2003b, Tohse et al., 2004, Beier and Anken, 2006, Tohse et al., 2006). Their sequences are highly conserved across vertebrates, and utilise zinc as a co-factor (Tohse et al.,

2006). Functionally, CA isozymes act to catalyze the conversion of carbon dioxide to bicarbonate:

− + 퐶푂2 (푎푞) + 퐻2푂(푙) → 퐻퐶푂3 (푎푞) + 퐻(푎푞) Equation 1

CA is thought to play a regulatory role in otolith growth, particularly with regards to symmetry and size. Anken et al. (2004) found that CA activity is down-regulated in hypergravitational conditions, suggesting that modulation of CA activity occurs as a response to the gravity vector. Similar to the otolith matrix component, Otolin-1, CA has also been found to be expressed diurnally, with levels being significantly lower during the day (Takagi et al., 2005, Tohse et al., 2006).

At least one type of CA has been identified in the saccular epithelium, being found in the transitional and squamous epithelial cells (Mugiya et al., 1979, Tohse et al., 2004). In contrast to this, Beier and Anken (2006) found that it was expressed with organic seeding particles before ionocyte formation, suggesting it may be involved in accretion of precursor particles during otolith initial formation (nucleation). As they also identified it in proximity to kinocilia, it may also play a role in the spatial distribution of bicarbonate and daily increment formation (Tohse et al., 2006).

Interestingly, Thomas et al. (2019) identified two types in the inner ear: CA 1 in the endolymph

(which is likely the result of saccular epithelial cell lysis) and CA 6 occurring solely in the otolith. It

34 is likely that CA 1 acts in bicarbonate transportation, ensuring that the endolymph is saturated with regards to aragonite precipitation (Tohse and Mugiya, 2004). As shown in Equation 2, aragonite precipitation results in carbon dioxide and water by-products. Thomas et al. (2019) postulated that CA 6 acts at the surface of the growing otolith to maintain local bicarbonate supersaturation, allowing further aragonite precipitation (Equation 1). However, as a pH gradient has been observed between endolymph and blood plasma (Payan et al., 1997), any protons produced in this way must be subsequently removed from the endolymph. Through experiments employing amiloride (a Na+ transport inhibitor) and ouabain (an inhibitor of Na+/K+-ATPase), Payan et al. (1997) suggested the presence of an energy dependent-Na+/H+ exchanger in the inner ear.

Although Na+/K+-ATPase has been implicated in inner ear development through knockout or pharmacological studies (Malicki et al., 1996, Whitfield et al., 1996, Blasiole et al., 2004, Blasiole et al., 2006, Gibert et al., 2011), the presence of a Na+/H+ exchange protein has not been directly confirmed. Thomas et al. (2019) proposed a different mechanism for pH control, that protons were removed from the endolymph through the action of V-type ATPase (Lagerstromfermer et al.,

1995). A similar mechanistic partnership between CA 6 and V-type ATPase (i.e. simultaneous regulation of bicarbonate and pH) has been thought to also occur in enamel formation (Smith et al., 2006, Lacruz et al., 2012, Yin and Paine, 2017). It is likely that CA 6 and V-type ATPase work in concert to maintain not only the bicarbonate concentration, but also the pH necessary for aragonite precipitation.

2+ − 퐶푎(푎푞) + 2퐻퐶푂3 (푎푞) → 퐶푎퐶푂3 (푠) + 퐶푂2 (푎푞) + 퐻2푂(푙) Equation 2

Bicarbonate ions produced in the saccular epithelial cells must be transported, either passively or actively, into the endolymph. In a comprehensive study employing a range of inhibitors (i.e. acetazolamide, disulfonate stilbenes, thiocyanate, and ouabain), Tohse and Mugiya (2001)

35 hypothesized that the transport of bicarbonate was accomplished by a dual mechanism of passive

- exchange with chloride ions concurrently with the action of HCO3 -ATPase. Although the latter has

- - not been directly identified, Thomas et al. (2019) determined the presence of a Cl and HCO3 exchanger, Band 3 Anion Exchange protein (B3AE). This protein has been previously found in erythrocytes as well as the basolateral membrane of distal tubule and collecting duct cells of the kidney, where it plays a role in regulation of calcification (Drenckhahn et al., 1985). It is likely that otolith increment deposition requires a tightly regulated environment, both at the growing otolith face, and in the endolymph. A thorough understanding of the intricacies of the proteins involved in ion transport, as well as those necessary for pH regulation, is crucial. This is particularly relevant when making ecological inferences from otolith data (e.g. increment widths) that rely on growth assumptions.

OTOLITH CHEMISTRY

Fish are often exposed to water bodies with different physico-chemical properties over their lifetime and these histories can be recorded in their otoliths as changes in the concentrations of minor and trace elements. The initial underlying assumption of otolith microchemistry is that the otoliths, by virtue of being acellular and metabolically inert, act to record the physico-chemical history of a fish (Campana, 1999, Campana and Thorrold, 2001). The incorporation of specific elements has been shown to be influenced by pH, salinity, temperature and concentration gradients (Mugiya and Tanaka, 1995, Elsdon et al., 2008, Sturrock et al., 2012, Izzo et al., 2015).

Currently, this information is typically extracted by determining otolith trace element-to-calcium ratios via laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-MS). In fish and fisheries research, these data can then be used in a variety of ways; to determine dispersal of

36 larvae (Swearer and Shima, 2010) , detection of diadromy or other migratory pathways (Hale and

Swearer, 2008, Hicks et al., 2010), reconstruction of temperature histories (Hoie et al., 2004,

Dorval et al., 2011), or as a natural or artificial tag (Campana et al., 1994, Warren-Myers et al.,

2014).

Several mechanisms have been suggested to explain the incorporation of trace elements into increments: substitution for calcium in the aragonite lattice, as has been demonstrated with strontium (Doubleday et al., 2014); being “trapped” as an increment forms, either as a free ion or through interaction with a similarly trapped macromolecule; or as part of the organic matrix, either through adhesion to a macromolecule or as a co-factor (Campana, 1999). For example,

Weigele et al. (2016b) found evidence of an iron-binding protein, Transferrin (also called serotransferrin), in the otolith matrix (Weigele et al., 2016). These are important distinctions because although the mechanisms of calcium substitution or random trapping of free ions may reflect changes in an individual’s physicochemical environment, it has been suggested that some trace element interactions with the organic matrix are more likely to be indicative of physiological events (Sturrock et al., 2014, Sturrock et al., 2015). Consequently, a deeper understanding of the biochemical mechanisms of inorganic material supply and trace element incorporation as well as the inorganic and proteinaceous composition of the otolith growth medium, the endolymph, is essential to the interpretation of microchemical data and subsequent reconstructions of environmental histories.

Endolymph has many characteristics in common with other extracellular fluids (such as

+ - + - comparable Na and Cl concentrations). However, it is distinguished by having high K and HCO3 concentrations, and by being relatively basic with regards to its pH (Kalish, 1991, Payan et al.,

1997). In comparison with blood plasma levels, it has slightly higher Cl- levels, and is less concentrated in terms of Na+, Mg2+ and Ca2+ (Enger, 1964, Fange et al., 1972, Miyamoto, 1978,

37

Mugiya and Takahashi, 1985, Kalish, 1991, Takagi, 1997, Payan et al., 1997). Importantly, the

2+ - endolymph has been reported as being supersaturated with Ca and CO2 (HCO3 ) with respect to aragonite precipitation (Takagi, 2002, Takagi et al., 2005). In addition to these electrolytes, the endolymph contains many other trace metals and anions (Melancon et al., 2009, Thomas et al.,

2017) as well as numerous proteins and proteoglycans, some of which have been implicated in biomineralization (Borelli et al., 2001, Thomas et al., 2019). Interestingly, the distributions of some of these components have been reported as being non-uniform throughout the sacculum, with

3- 2+ + PO4 , proteins and Mg favouring the macula-facing (proximal) side and K and total CO2 being more abundant in the distal side (Payan et al., 1999, Borelli et al., 2001). Furthermore, although

2+ - - + + the mechanisms of supply of some ions have been suggested (notably Ca , HCO3 , Cl , K , Na ) the transport mechanisms for trace elements (e.g. Li+, Sr2+, Ba2+, Mn2+ and others) remain unresolved.

It is currently assumed that they are either “swept up” as part of active transport of more abundant ions, or enter the endolymph via paracellular pathways (Campana, 1999, Payan et al.,

2002).

The question of how calcium is supplied to the endolymph has been of keen interest, for obvious reasons. Several studies investigating this have been published, but with opposing viewpoints. The first hypothesis is that calcium is generally supplied to the endolymph passively, travelling paracellularly through tight junctions in the saccular epithelium, and that active transport of calcium is not a feature of otolith biomineralization. Payan et al. (2002) presented this argument, justified through experiments employing ATP inhibitors. This hypothesis is supported by the observation of Ibsch et al. (2004) of calcium precipitates on the proximal face of the otolith, between it and the macula. In that study, the authors suggested the transport of calcium via a paracellular route. In contrast to this, Mugiya and Yoshida (1995) found that the use of a range of

ATP inhibitors resulted in the disruption of otolith calcification, and that no paracellular transport

38 mechanism was evident (as visualized through mannitol labelling). These two contradictory viewpoints were eventually resolved by Cruz et al. (2009), who incisively noted that these hypotheses (like so many early otolith studies) were based on pharmacology, histology or kinetic observations, rather than through direct detection of the proteins in question. In their study, the presence of Plasma Membrane Ca2+-ATPase (PMCA) was positively confirmed in the inner ear through RNA probes coupled with in situ hybridization. Furthermore, knockout of its related gene,

ATP2b1a, resulted in aberrant otolith phenotypes. However, as calcium accretion was still observed to occur in mutants, they concluded that calcium supply could also be additionally occurring through either paracellular transport or the presence of another PMCA isoform. In support of this, Thomas et al. (2019) identified Na+/K+-ATPase, Calmodulin, PMCA as well as

Na+/Ca2+exchanger in the inner ear of black bream (Acanthopagrus butcheri). Considering all of this, it is likely that intracellular Ca2+ is transferred to PMCA and Na+/Ca2+exchanger by Calmodulin, with the electrochemical gradient required for active transport being supplied by Na+/K+-ATPase.

Calcium is then transported to the endolymph through a combination of active transport and ion exchange.

THE OTOLITH ORGANIC MATRIX

Although some investigations have been carried out into the organic components of otoliths, this field is still in its infancy. The overwhelming majority of otolith studies are on age and growth (i.e., otolith microstructure). This is in stark contrast to a far smaller fraction of papers on inorganic chemical composition (i.e., otolith microchemistry) and the scant number that are focused upon the biochemistry of the fish inner ear (Fig. 2.2). In these latter studies, a picture of the protein

39 landscape of the otolith and endolymph has begun to develop, with several proteins being identified and partially characterised.

Otolith publications 1921-2017 450 400 350 300 250 200 150

Number of of publications Number 100 50

0

1921 1942 1963 1984 2005 1924 1927 1930 1933 1936 1939 1945 1948 1951 1954 1957 1960 1966 1969 1972 1975 1978 1981 1987 1990 1993 1996 1999 2002 2008 2011 2014 2017 Year

All otolith papers Otolith microstructure papers Otolith microchemistry papers Otolith organic matrix papers

Figure 2.2. Results of keyword searches using Web of Science. All searches employed the keywords “otolith* and fish*”. In addition to this, the following keywords were employed for specific searches: Otolith microstructure papers – “and (microstruc* or increment* or ageing)”; Otolith microchemistry papers – “and (microchem* or trace element or fingerprint*)”; Otolith organic matrix papers – “and matrix”. All search data retrieved on 17 September 2018 from http://www.webofknowledge.com. While Reibisch (and others) first noted the presence of organic material between the annual bands of plaice (Pleuronectes platessa) otoliths (Reibisch, 1899), the first foray into otolith protein analyses did not occur until 1956, when Dannevig (1956) investigated the constituents of Atlantic cod (Gadus morhua) otoliths. In that study, a protein “conchiolin”, imagined to be related to molluscan shell proteins, was suggested. This study was also remarkable in noting an increased prevalence of organic material in the ‘winter’ compared to ‘summer’ seasonal bands. Despite these early studies of the organic constituents within the otolith, the first comprehensive analysis of the biochemistry of otoliths did not occur until the seminal 1969 study conducted by Degens,

40

Deuser and Heidrich (1969). In this study, the authors investigated the otoliths of 25 species of fish, and identified a potential protein, which they dubbed “otolin”. This protein was described as fibrous with a molecular weight exceeding 150 kDa. Using ion-exchange chromatography techniques, they were able to discern the number, types and amounts of amino acids present.

“Otolin” was found to be highly acidic, with an abundance of aspartic and glutamic acid residues.

The authors proposed that this specific feature would allow “otolin” to facilitate mineralisation.

Indeed, they argued that the presence of acidic residues (and therefore carboxyl side chains) would not only serve to coordinate calcium ions into the growing crystal lattice, but that the large number of oxygens present would encourage the selection of aragonite over calcite. They hypothesised that the carboxyl side chains would bind calcium cations, encouraging aragonite’s 9- fold coordination, eventually being stabilised by their substitution with endolymph hydrogen carbonate anions, initiating nucleation. Notably, Degens et al. (1969) also classed “otolin” as a potential collagen, remarking on the relative similarity of its amino acid content with both keratin

(similar abundances, presence of cysteine) and collagen (low levels of basic amino acids, presence of hydroxyproline). They also found that “otolin” degraded to 70-80 kDa units containing identical amino acid distributions as those in the parent fraction upon treatment with urea/hydroxylamine.

The detection of “otolin” in so many teleost species, by this study, stimulated further research in this area. Further studies were undertaken that cited organic components of the otolith with comparable amino acid compositions to otolin. In two 1986 studies, Morales-Nin found that amino acid composition as well as overall protein amounts differed between juvenile and adult otoliths, first hinting at the different roles that the organic matrix plays over the course of a fish’s life

(Morales-Nin, 1986a, Morales-Nin, 1986b). Later studies began to show that the otolith matrix was not as simple as containing a single, dominant protein, with staining techniques indicating at least

3 candidate proteins (Gauldie et al., 1990), and sodium dodecyl sulfate-polyacrylamide gel

41 electrophoresis (SDS-PAGE) analyses revealing the presence of more than 15 in the water-soluble fraction alone (Baba et al., 1991, Asano and Mugiya, 1993, Sasagawa and Mugiya, 1996). A caveat, however, must be made about these early studies. These investigations made use of HCl in protein hydrolysis prior to amino acid sequencing. An unavoidable side effect of this type of treatment is the conversion of Asparagine and Glutamine residues to their acidic counterparts, Aspartic Acid and Glutamic Acid. Consequently, these studies were unable to describe the true amino acid composition of inner ear proteins.

To date, only one deep sequencing study has been done on the proteome of the fish inner ear. In that study, Thomas et al. (2019) found the presence of over 380 proteins in the otolith, including all that had been previously identified. Prior to that study, the following protein types were identified in the otolith and inner ear: (1) Inner ear collagens and their accessories, such as Otolin-

1 (Degens et al., 1969, Davis et al., 1995a, Murayama et al., 2002) and Precerebellin-like protein

(Cblnl) (Kang et al., 2008); (2) Transferrin-family proteins, such as Otolith Matrix Protein-1 (OMP-1)

(Murayama et al., 2000, Murayama et al., 2004, Murayama et al., 2005) and Transferrin (Weigele et al., 2016); (3) Intrinsically disordered proteins implicated in crystal polymorph selection, such as

Starmaker (Stm) (Söllner et al., 2003), Starmaker-like (Stm-l) (Bajoghli et al., 2009) and Otolith

Matrix Macromolecule-64 (OMM-64) (Tohse et al., 2008, Tohse et al., 2009, Poznar et al., 2017);

(4) An ortholog of mammalian Otoconin 90 (Petko et al., 2008); (5) Secreted protein acid rich in cysteine (Sparc, also called Osteonectin) (Kang et al., 2008, Weigele et al., 2017, Weigele et al.,

2016); (6) A serine protease inhibitor, (Neuroserpin (Kang et al., 2008, Weigele et al., 2016); (7)

Myosin Light Chain-9 (Weigele et al., 2016, Weigele et al., 2017); (8) Calcium transporters,

Calmodulin and Calbindin (Motta et al., 2009); and (9) Otolith membrane and hair cell constituents such as Otogelin, Otogelin-like and Alpha- and Beta-Tectorin (Yariz et al., 2012, Stooke-Vaughan et al., 2015, Weigele et al., 2016). In addition to these, several studies have indicated the presence of

42 other organic molecules in the otolith matrix, including glycosaminoglycans, polysaccharides, and lipoproteins (Mugiya, 1968, Wright, 1991, Asano and Mugiya, 1993, Takagi et al., 2000, Dauphin and Dufour, 2003), as well as glycogen-containing particles thought to be involved in “seeding” the nascent otolith (Pisam et al., 2002, Beier and Anken, 2006, Dauphin and Dufour, 2008, Stooke-

Vaughan et al., 2015).

As well as matrix molecules, it has been shown that otolith development and growth also rely on accessory macromolecules present in both saccular epithelial cells and hair cells in the cilia.

Several proteins have been hypothesised as being involved in the supply of materials from the saccular epithelial cells to the endolymph as well as maintaining the physical conditions for crystallisation. These include calcium and anion exchangers, a variety of ATPases (Mugiya, 1986,

Mugiya and Yoshida, 1995, Payan et al., 1997, Tohse and Mugiya, 2001, Blasiole et al., 2004, Shiao et al., 2005, Blasiole et al., 2006, Cruz et al., 2009, Gibert et al., 2011) and at least one isozyme of

CA (Mugiya, 1977, Mugiya et al., 1979, Mugiya and Takahashi, 1985, Tohse and Mugiya, 2001,

Beier et al., 2002, Borelli et al., 2003b, Tohse et al., 2004, Tohse and Mugiya, 2004, Shiao et al.,

2005). Thomas et al. (2019) found the presence of a Sodium/Calcium Exchanger 2, Plasma

Membrane Ca2+-ATPase (PMCA), Band 3 Anion Exchange Protein (B3AE), V-type Proton ATPase (V-

ATPase) and two separate isozymes of CA, 1 and 6. Additionally, Osterix and the “chaperonin”

GP96 have both also been implicated as necessary for fish otic development (Sumanas et al., 2003,

DeLaurier et al., 2010, Renn and Winkler, 2014). Inner ear proteins thus far identified are summarised in Table 2.1.

43

Table 2.1. Proteins directly identified (either through mRNA labelling, gene sequencing or proteomics) in otoliths or endolymph of fish. Structural details listed include domains, families or superfamilies identified through the conserved domain search tool, https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi, accessed on February 14 2018. References listed indicate the first direct identification of a given protein in each species.

Protein Species Structural details Function Family/Grouping Reference Otolin-1 Lepomis macrochirus C1q domain, Ca2+-binding, structural C1q family Davis et al., 1995 Oncorhynchus keta Collagen X/VIII collagen, forms hubs Murayama et al., 2002 Oncorhynchus mykiss domain, trimeric Murayama et al., 2004 Danio rerio structure, Murayama et al., 2005 Oryzias latipes glycoprotein Nemoto et al., 2008 Oreochromis mossambicus Weigele et al., 2016a;

Acanthopagrus butcheri Thomas et al., 2019 Precerebellin-like O. mykiss C1q domain Interacts with Otolin-1 C1q family Kang et al., 2008 protein Otolith matrix protein-1 O. mykiss Transferrin domain, Structural protein, Fe3+- Transferrin family Murayama et al, 2004 D. rerio HEXXH Zn-binding binding, putative Zn2+- Murayama et al, 2005 - Gadus morhua domain, glycoprotein binding, HCO3 binding Andersen et al., 2011 O. mossambicus Weigele et al., 2016 A. butcheri Thomas et al., 2019 Transferrin O. mossambicus Type 2 periplasmic Fe transport, Fe3+- Transferrin family Weigele et al., 2016 - A. butcheri binding fold binding, HCO3 binding Thomas et al., 2019 superfamily Starmaker; D. rerio Na+/Ca2+ exchanger Ca2+-binding, crystal Intrinsically disordered Söllner et al, 2003 Starmaker-like; O. latipes domain, polymorph selection, proteins Bajoghli et al., 2009 Otolith matrix O. mykiss phosphorylated crystal nucleation Tohse et al., 2008 macromolecule-64 A. butcheri Thomas et al., 2019 Extracellular A. butcheri FAM20C Superfamily Mn2+-binding, FAM20C Thomas et al., 2019 serine/threonine phosphorylates protein kinase FAM20C biomineralization proteins (SIBLING)

44

Protein Species Structural details Function Family/Grouping Reference Otoconin 90-like; Otoc1 D. rerio Phospholipase A2- Ca2+-binding, necessary Otoconins Petko et al., 2008 O. mossambicus like domain, Brain for otolith seeding Weigele et al., 2016 A. butcheri acid soluble protein Thomas et al., 2019 superfamily Sparc; Osteonectin O. mykiss EF-hand, extracellular Ca2+-binding, Collagen- Sparc family Kang et al., 2008 D. rerio calcium binding binding, enhances Nemoto et al., 2008 Ictalurus punctatus motif; follistatin-like plasminogen activity Weigele et al, 2016 O. latipes Sparc domain Thomas et al., 2019 O. mossambicus A. butcheri Neuroserpin O. mykiss Serine protease Regulates proteolytic Serpin family Kang et al., 2008 A. butcheri inhibitor family cascades Thomas et al., 2019 Matrix A. butcheri Hemopexin-repeats, Zn2+-binding, gelatin- Neuroserpin inhibited Thomas et al., 2019 metalloproteinase 2 Matrixin family, binding, cleaves collagen proteolytic cascade (MMP 2); Fibronectin type 2 types I, IV, V, VII, X, XI, 72 kDa Type IV domain XIV, gelatin collagenase Tissue inhibitor of A. butcheri Tissue inhibitor of Inhibitor of matrix Inhibits MMP 2 Thomas et al., 2019 matrix metalloproteinase metalloproteinases metalloproteinase 2 family, Metzincin- (TIMP 2) binding interface, hemopexin-domain- binding interface Plasminogen A. butcheri Trypsin-like serine Proteolytic factor in a Neuroserpin inhibited Thomas et al., 2019 protease family, five variety of processes, proteolytic cascade Kringle domains activates collagenases, cleaves von Willebrand factors Tissue type plasminogen A. butcheri Trypsin-like serine Cleaves plasminogen to Neuroserpin inhibited Thomas et al., 2019 activator protease family, form plasmin proteolytic cascade calcium-binding EGF domain, 2 Kringle domains

45

Protein Species Structural details Function Family/Grouping Reference Otopetrin 1 D. rerio Otopetrin Intracellular Ca2+ Saccular Hughes et al., 2004 Paralichthys lethostigma superfamily regulation and protein transmembrane protein Wang et al., 2011 A. butcheri trafficking Thomas et al., 2019 Otogelin, Otogelin-like P. lethostigma Von Willebrand Adheres otolith to Otolith membrane Wang et al., 2011 D. rerio factor D domains, macula protein Yariz et al., 2012 O. mossambicus glycoprotein Weigele et al., 2016 A. butcheri Thomas et al., 2019

HSP90/GP96 D. rerio HSP 90 superfamily Processes integrins or Otolith membrane Sumanas et al., 2003 A. butcheri otogelin/otogelin-like protein Thomas et al., 2019 Alpha/beta-tectorin P. lethostigma Zona Pellucida, Von Collagen binding, Otolith membrane Wang et al., 2011 D. rerio Willebrand factor D adheres otolith to protein Yang et al., 2011 O. mossambicus domains, N-terminus kinocilia Weigele et al., 2016 A. butcheri Nidogen domain Thomas et al., 2019 Myosin 9 Light Chain O. mossambicus EF-hand motif Ca2+-binding, Motor Myosins Weigele et al., 2016 A. butcheri protein, possible Thomas et al., 2019 component of kinocilia Biglycan A. butcheri C-terminal novel E3- Proteoglycan, binds to Proteoglycans Thomas et al., 2019 ligase domain, type I and II collagen, Leucine rich repeat collagen fibre assembly N-terminal domain Decorin A. butcheri C-terminal novel E3- Proteoglycan, binds to Proteoglycans Thomas et al., 2019 ligase domain, type I and type II Leucine rich repeat collagen, fibronectin and N-terminal domain TGF-beta. Thought to be the result of gene duplication with biglycan Osteoglycin; Mimecan A. butcheri C-terminal novel E3- Induces bone formation Proteoglycans Thomas et al., 2019 ligase domain, leucine rich repeats Epiphycan A. butcheri C-terminal novel E3- Role in bone formation, Proteoglycans Thomas et al., 2019 ligase domain, organization of cartilage leucine rich repeats matrix

46

Protein Species Structural details Function Family/Grouping Reference Aggrecan A. butcheri C-type lectin domain, Hyaluronic acid binding, Proteoglycans Thomas et al., 2019 Immunoglobin component of cartilage domain, link extracellular matrix domains, complement control protein superfamily domain - Carbonic anhydrase 1, 6 O. mykiss Carbonic anhydrase Converts CO2 into HCO3 , Carbonic anhydrases Mayer-Gostan, 1997 Scophthalmus maximus alpha superfamily regulates pH, Zn2+- Mayer-Gostan, 1997 O. mossambicus (Type 1, 6); binding Beier et al., 2002 D. rerio Laminin G-like Shiao et al., 2005 A. butcheri domain (Type 6) Thomas et al., 2019 Plasma membrane D. rerio E1-E2 ATPase Transports Ca2+ across Ion transport Cruz et al., 2009 calcium ATPase A. butcheri superfamily plasma membrane Thomas et al., 2019 Band 3 anion exchange A. butcheri Band 3 cytoplasmic Cell membrane protein, Ion transport Thomas et al., 2019 - - protein domain, glycoprotein exchanges Cl and HCO3 Na+/Ca2+ exchanger A. butcheri Calx-beta superfamily Cell membrane protein, Ion transport Thomas et al., 2019 exchanges Ca2+ and Na+ Calmodulin A. butcheri Penta-EF-hand family Mediates ion channels Ion transport Thomas et al., 2019 through Ca2+-binding V-type proton ATPase A. butcheri V-type ATPase Controls cellular acidity Ion transport Thomas et al., 2019 superfamily through H+ transport

47

C1q family proteins

Otolin-1

Although the otolith organic matrix had long been assumed to contain a collagen as a primary constituent, clear evidence for what this might be was not provided until two comprehensive studies conducted by Davis et al. (1995a, 1997) on bluegill sunfish (Lepomis macrochirus). They identified a 2.0kb sequence from the cDNA of bluegill sunfish through differential screening and, from this, were able to deduce a potential primary sequence for this protein, which they named

“saccular collagen”. Saccular collagen in L. macrochirus is 423 amino acids (AA) in length, with a large 217 AA collagenous domain containing many (70) Gly-x-y repeats, flanked by non- collagenous N-terminus and C-terminus domains. The C-terminus domain was found to be homologous with the non-collagenous domains of Collagen types VIII and X. Their analyses showed a lack of a lectin-binding domain, indicating the likelihood of saccular collagen being a collectin, or short-chain collagen. Saccular collagen was found to be a 95 kDa protein that is secreted apically from the supporting cells on the outer periphery of the saccular macula, with immunohistochemical results suggesting it to be a glycoprotein present in the gelatinous layer of the otolith membrane. The determination of the primary sequence of saccular collagen, encouraged Murayama et al. (2002) to search for its homolog in chum salmon (Oncorhynchus keta). In this study, the authors extracted protein from the gelatinous, EDTA-insoluble portion of otoliths. This fraction was de-N-glycosylated with a glycopeptidase and then analysed with both

SDS-PAGE and Matrix Assisted Laser Desorption/Ionization-Time of Flight-Mass Spectrometry

(MALDI-TOF-MS). Through their work, they were able to extensively determine both the primary amino acid and cDNA sequences of this protein in chum salmon, which they named “Otolin-1”.

48

Chum Salmon Otolin-1 is a glycoprotein of 100 kDa (deglycosylated to 90 kDa), with 68% homology to bluegill sunfish saccular collagen. Similarly to saccular collagen, it also has a C-terminal domain with high homology with collagen types VIII and X, and its collagenous domain also contains a comparable number (74) of Gly-x-y repeats. Murayama et al. (2002) speculated that if Otolin-1 behaves like other collagens, it should oligomerise into a 3D hexagonally-arranged lattice, through the C-terminal and N-terminal domains acting as nodes with the collagenous domain as a spacer.

Since these studies, further homologs have been determined in other fish species (Murayama et al., 2004, Murayama et al., 2005, Nemoto et al., 2008, Weigele et al., 2016), while still more have been added to the NCBI database, having been determined through genomic pipeline projects.

Otolin-1 shares sequence homology, also, with Homo sapiens Otolin, suggesting that it is reasonably conserved within vertebrates (Deans et al., 2010). While the specific primary sequence differs among species, they all contain a non-collagenous N-terminus domain, a repeated Gly-x-y motif in the middle, “spacer” collagen domain, and a globular C-terminal domain with high homology to type VIII and type X collagens (now referred to as the C1q superfamily domain).

With regards to biomineralization, Otolin-1 is a key player in the formation of daily increments. In zebrafish (Danio rerio), its incorporation into the otolith happens within 72 hours post fertilisation

(hpf), coinciding with increment formation (Murayama et al., 2005). It has also been demonstrated, in rainbow trout (Oncorhynchus mykiss), that its expression follows a diurnal pattern (Borelli et al., 2003a, Takagi et al., 2005). It has been found to be expressed in the ventral meshwork area of the saccule (but not the utricle) of Mozambique tilapia (Oreochromis mossambicus), suggesting a crucial role in the formation of sagittae, but not lapilli (Weigele et al.,

2015b). Furthermore, in zebrafish, it has been shown to interact with OMP-1, being essential for both proper biomineralization and correct tethering of the otolith to the sensory macula, and that

49 in embryo, morpholino knockdown of Otolin-1 leads to reduced otolith growth rate, larger otoliths, and fusion of sagittae with lapilli within 72 hpf (Murayama et al., 2005).

The globular C1q domain in collagens has been shown, in humans, to have a high affinity for both calcium ions as well as other macromolecules (Gaboriaud et al., 2012). It has been found that short-chain collagens trimerize at their C1q domains, creating a reticular network of “hubs” with radial collagen fibre “spokes” (Kwan et al., 1991, Sutmuller et al., 1997). The presence of such a reticular network in otoliths has been observed in Nile tilapia (Oreochromis niloticus) (Zhang,

1992). Furthermore, in a study by Hołubowicz et al. (2017), it was hypothesised that this domain in zebrafish Otolin-1 similarly acts as a “hub” for protein interactions during biomineralization. In that study, the authors found that, in solution, the presence of calcium ions induced trimerization of the C1q domain. They additionally found that there was a concentration dependent effect of calcium on stabilization of the C1q domain, with the optimum concentration being comparable to that in fish endolymph. In humans, Otolin-1 has been shown to bind with fundamental otoconial matrix proteins Otoconin-90 and Cerebellin-1 (Deans et al., 2010). Importantly, Hołubowicz et al.

(2017) hypothesised that Otolin-1 acts to mediate the biomineralization process through trimerizing and interacting with analogous matrix proteins in the fish otolith (e.g. OMP-1,

Starmaker, OMM-64) as well as binding calcium ions necessary for aragonite precipitation.

Together, the above indicates that Otolin-1 likely plays a stabilising role in the otolith, by compacting the organic matrix in the discontinuous zone, while at the same time allowing for coordination of the calcium ions in the incremental zone. Additionally, it also likely plays a role in correctly anchoring the otolith to the sensory macula (Murayama et al., 2005).

50

Precerebellin-like protein

A single study identified the presence of precerebellin-like protein (Cblnl, also sometimes called

Complement C1q protein). In that study, Kang et al. (2008) found five peptides that aligned with two separate isoforms of Cblnl in otoliths from adult rainbow trout. Structurally, Cblnl contains a

C1q domain, and is relatively short, having a molecular weight of 19.8 kDa. Unlike Otolin-1, however, cblnl does not contain a collagen domain. It is thought that its role is to interact with

Otolin-1 during assembly, either by creating a “cap” on the surface of the above described trimer

“hubs” or through reduction of the internal connectivity of these “hubs” (Kang et al., 2008).

Transferrin-family proteins

Otolith Matrix Protein-1

Otolith matrix protein-1 was first characterised by Murayama et al. (2000). Indeed, although it was among the first matrix proteins to be identified, relatively little is known about it. OMP-1 is likely to be directly involved in otolith growth through coordinating hydrogen carbonate and is additionally a necessary component for correct deposition of Otolin-1 to the face of the growing discontinuous zone.

It has been found, through several studies, that OMP-1 interacts with Otolin-1, potentially playing a role as a scaffold for the formation of the incremental zone, with its knockdown causing Otolin-1 to not be expressed as well (Murayama et al., 2005). Although OMP-1 is likely to be essential for increment formation, it is unlikely to play a major role in nucleation of the otolith. Histochemical studies of the expression of otolith proteins in rainbow trout, and zebrafish have found that OMP-

1 expression coincides with increment deposition, first occurring well after the primordium has

51 formed (Murayama et al., 2004, Murayama et al., 2005). In contrast to this, Petko et al. (2008) found that OMP-1 expression in zebrafish first occurs at 15 hpf, before increment formation, but is most strongly expressed by 24 hpf. Similarly, OMP-1 has been found to be expressed in larval

Mozambique tilapia (Weigele et al., 2017) prior to formation of the primordium, and continues to be expressed throughout development. The expression of OMP-1 has been found to occur simultaneously with Otolin-1 expression and in the non-sensory dorsal and ventral meshwork regions of the macula. In cod (Gadus morhua) (Andersen et al., 2011), its expression has been found to be solely restricted to the otolith. These regions are both near to ionocytes as well as co- locating with carbonic anhydrase expression (Weigele et al., 2015b). Finally, unlike Otolin-1, in rainbow trout (Takagi et al., 2005), OMP-1 is expressed continually, not following a diurnal pattern.

Structurally, OMP-1 is a monomeric glycoprotein and member of the transferrin family (Andersen et al., 2011). Unlike other transferrins, however, it contains only one transferrin lobe rather than two. It was originally thought to contain five Fe binding sites (Murayama et al., 2000). However, these potential sites in OMP-1 show strong homology with the non-iron binding C-terminus lobe of human melanotransferrin (Lambert, 2012). Specifically, in OMP-1, the canonical iron-binding residues, His, have been replaced with Asn (Gaffney and Valentine, 2012). Consequently, it is likely that OMP-1 has a role other than iron transfer (Murayama et al., 2005, Andersen et al., 2011,

Gaffney and Valentine, 2012). In rainbow trout (Murayama et al., 2000) and zebrafish (Murayama et al., 2005), the OMP-1 homologs contain a HEXXH-motif, marking their capability to bind Zn.

Interestingly, another member of the transferrin family is inhibitor of carbonic anhydrase

(Lambert, 2012). A BLASTP alignment of the murine form of carbonic anhydrase inhibitor showed

34% homology with rainbow trout OMP-1 (accessed on February 14 2018 https://blast.ncbi.nlm.nih.gov/blast/Blast.cgi). It is thus possible that OMP-1 has a dual role during

52 increment deposition. Firstly, its presence allows for the expression of Otolin-1, and the formation of the collagenous scaffold described above. Secondly, its transferrin domain is likely to allow coordination of bicarbonate ions during precipitation whilst potentially inhibiting further activity of carbonic anhydrase in situ.

Serotransferrin

In two recent studies, serotransferrin (sometimes called simply “transferrin”) was identified in the organic matrix (Weigele et al., 2016, Thomas et al., 2019). In vertebrates, the function of transferrin is unsurprisingly as an iron transporter, binding and releasing Fe3+ as a pH-specific response to receptors (Lambert, 2012). In a study of transferrin variants in cod (Andersen et al.,

2011), two separate serum transferrin genes, Tf1 and Tf2 were both found to be expressed in the brain, albeit with the latter showing lower expression levels when compared with the former. In that study, neither was found to be expressed in the otolith. Furthermore, Thomas et al. (2019) found serotransferrin to be present in the endolymph in extremely high levels, being at the same order of magnitude as other serum proteins, ApoA1 and Haemoglobin (Hb). In that study, the authors also found these, and other “typical” serum proteins, present in the otolith organic matrix at high abundances. The presence of transferrin in the otolith is likely due to the trapping of endolymph – a fluid with strong similarities to cerebral spinal fluid – during increment formation, and therefore is not necessarily indicative of it playing a functional role in the organic matrix.

However, considering that transferrin also has the capacity to bind bicarbonate, similar to OMP-1, it may have a role in mineral formation (Lambert, 2012).

53

Intrinsically Disordered Proteins

An intrinsically disordered protein (IDP) is one which lacks a specific tertiary structure, but that upon binding to a ligand, often folds into a globular or other specified shape (Dyson and Wright,

2002, Fink, 2005, Kaplon et al., 2008, Kaplon et al., 2009, Uversky and Dunker, 2010). IDPs, due to their very nature, can be difficult to identify through sequence alignment as functional homologs from two separate species may have the same biochemical role but vastly dissimilar primary sequences. In proteomics, a typical Mascot search of otolith proteins versus the ray-finned fish proteome is unlikely to uncover an IDP (or, indeed, any other novel sequences). For example, in the zebrafish otolith, the Starmaker (Stm) protein has been established as playing a key role in polymorph selection, nucleation and regulation of crystal growth. A BLASTP of its equivalent protein in rainbow trout, OMM-64, reveals 37% homology with across 35% of the length of its sequence (accessed on February 14 2018 https://blast.ncbi.nlm.nih.gov/blast/Blast.cgi). I suggest that searches against Actinopterygii proteome, rather than a species-specific genome or transcriptome, accounts for its lack of discovery in studies (such as that conducted by Weigele et al. on O. mossambicus (2016b)) rather than its absence from the otolith, as others (e.g., Rozycka et al., 2014, Thomas et al., 2019) have found putative Stm functional homologs with low coverage and homology with Stm or OMM-64. Despite the difficulty in identifying these proteins, understanding their roles in nucleation and growth is crucial to our understanding of otolith biomineralisation.

Starmaker

Starmaker was the first IDP found through the screening of the zebrafish genome for any potential homologs of human dentin sialophosphoprotein (DSPP) (Söllner et al., 2003). DSPP is an important protein in tooth biomineralization, but also expressed in the human inner ear (Xiao et al., 2001,

54

Zhang et al., 2001). In their study, Söllner et al. (2003) reported the existence of a protein that, when knocked out, caused the formation of calcite (rather than aragonite) sagittae that were both star-like in appearance and lacking in daily rings, instead having a non-radial, disorganised organic matrix. Importantly, both morphant and wild-type otoliths were found to be normal during the seeding and anchoring stages, and that later, during early incrementation, the crystal polymorph type switched in morphants. In a separate study, Söllner et al. (2004) determined that Stm localisation is widespread throughout the inner ear epithelial cells, and that proper secretion is likely due to a transmembrane delivery protein, Otopetrin 1, discussed further below. With regards to the properties of Stm: it is highly hydrophilic, abundant (35%) in acidic residues, and predicted to be 20% phosphorylated. Stm has been found to be expressed by the entire epithelium of the otic placode within the first 20 hpf, and secreted into the lumen concurrently with seeding particles, being bound to them (Nicolson, 2004). Considering this early stage of expression, and that increment deposition does not occur until 72 hpf in zebrafish, Stm is almost certainly a key part of the primordium. Further, it is an integral part of the organic matrix throughout otolith development, but its expression is limited to the sensory epithelium – the site of otolith accretion.

While IDPs can be globular in structure, a study conducted by Kaplon et al. (2008, 2009) indicated that Stm is likely to be rod-like and showing low compaction due to its lack of a hydrophobic core.

The rod-like structure was inferred from its molecular weight (estimated as being 68-75 kDa, depending on method of determination) and its unusually high frictional ratio of 2.6-3.1 (compact, globular proteins usually have a frictional ratio of ~1, whereas extended proteins have a ratio of

>2). With regard to its degree of compaction being directly affected by calcium ion concentration, it was found that a fiftyfold increase of calcium ion concentration resulted in a reduction of 9.5A in the Stokes radius of the protein – a reasonable amount of compaction. Finally, Kaplon et al.

(2009) found residual secondary structures within the protein, and concluded that Stm likely

55 contains short, locally ordered clusters (such as alpha helices and beta sheets) that crucially influence its ability to fold. At the primary sequence level, Stm is rich in aspartic and glutamic acid, as well as containing a sequence of alternating serine and aspartic acid residues. It has been suggested that for it to effectively act during nucleation (and subsequent polymorph selection), it must be able to coordinate both with collagenous proteins as well as calcium ions. Kaplon et al.

(2009) indirectly found evidence for calcium binding, supporting the notion that richness in acidic residues (Asp and Glu) directly relate to calcium coordination, whereas they indicate that the repetitive Ser-Asp motif is typical of collagen fibril binding proteins.

However, the biomineralizing properties of Stm are not solely due to its aspartyl or glutamyl residues. In a thorough in vitro study of kinase effects on Stm activity, Wojtas et al. (2015) found that Stm has the potential to be highly phosphorylated. Not only was Stm an in vitro target of casein kinase II (an enzyme able to non-specifically phosphorylate candidate proteins), but also that its biomineralization ability was directly enhanced not only by calcium ion concentration but also by the degree of phosphorylation. They hypothesised that Stm is multifunctional with regards to its involvement in aragonite growth. Specifically, the large amount of phosphorylation is thought to cause a stiffening and extending of the conformation of Stm. This extended structure would allow for regularly spaced carboxylate and phosphate side groups, serving as binding points for calcium, with the interval between side groups selecting for calcium carbonate polymorph.

Despite this, they found that recombinant Stm stabilized calcite in vitro, rather than aragonite, and noted that this discrepancy was no doubt due to other assisting macromolecules present in the endolymph as well as the amount of magnesium ions present in the microenvironment.

Additionally, the high level of phosphorylation (coupled with acidic residues) results in an increased negative charge for the Stm molecule. It has been suggested that this property stimulates nucleation, by encouraging the clustering of calcium ions at the start of increment

56 deposition. Finally, they noted that increasing degree of phosphorylation had a marked effect in reducing crystal growth, suggesting it plays a role as an inhibitor (Wojtas et al., 2012, Wojtas et al.,

2015). It is likely, then, that phosphorylation of Stm (and its homologs) has multiple effects. Firstly, the addition of phosphate groups to Stm increases its negative charge, and allows it to attract more Ca2+ ions, thereby stimulating nucleation. Secondly, phosphorylation changes the tertiary structure of Stm causing it to align those Ca2+ ions in an aragonite geometry. Finally, the compact shape of phosphorylated-Stm restricts the plane of crystal growth, regulating aragonite precipitation and allowing close packing of needles.

Otolith Matrix Macromolecule-64

In rainbow trout, the IDP responsible for polymorph selection, OMM-64, has been the focus of three studies (Tohse et al., 2008, Tohse et al., 2009, Poznar et al., 2017). This protein has been found to interact with matrix collagens and was identified through screening rainbow trout cDNA with an antiserum that had been raised against the entire organic matrix. Structurally, OMM-64 shares some similarities with Stm in terms of its amino acid proportions, being rich in both acidic residues as well as being predicted to be able to be ~15% phosphorylated (Tohse et al., 2008).

Furthermore, an analysis of the primary sequences of it and Stm reveal that they both contain the conserved domain common to members of the potassium dependent sodium/calcium exchanger superfamily (accessed on February 14 2018 https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi). Additionally, the gene encoding OMM-64 shares similarities with the Stm gene in zebrafish and DSPP in humans, but are highly diverged and indicate the conservation of aragonite-producing genes in vertebrates (Tohse et al., 2008).

In terms of its expression, OMM-64 has been found to be restricted to the sacculum, semicircular canals and ovaries (Tohse et al., 2008). Specifically, its mRNA is found strongly in macular

57 peripheral cells as well as non-mitochondria rich transitional epithelial cells. Histochemical staining further showed that it is expressed in the matrix producing cells and is deposited directly onto the otolith, and is consequently present in the increments (Tohse et al., 2008). It was suggested by

Tohse et al. (2008) that OMM-64 is expressed in the early stages of otolith formation due to their finding of a partial sequence corresponding to it in a cDNA library from an embryo neurogenesis study.

Structurally, OMM-64 is 64 kDa in size and contains several tandem repeats which are rich in acidic residues. Indeed, like Stm, it is highly enriched in glutamic and aspartic acids, containing 213 of these residues (Poznar et al., 2013). It interacts readily with Otolin-1, forming ring-like structures.

As well as Otolin-1, it has been suggested to potentially bind heparan sulfate glycosaminoglycans which can be released following deglycosylation with Heparitinase II. In a comprehensive structural study conducted by Poznar et al. (2017), it was demonstrated that not only is OMM-64 an IDP, but that it can bind calcium ions with high capacity (being able to bind 61 calcium ions per molecule) but with weak affinity. Also, like Stm, increasing calcium ion concentration results in an increased degree of compaction of OMM-64, suggesting that this quality is essential for proper aragonite growth in rainbow trout. Also, when compacting, OMM-64 did not gain secondary structure.

In an in vitro crystallisation study, Tohse et al. (2009) demonstrated that a high molecular weight aggregate containing OMM-64 and Otolin-1, was necessary for the formation of aragonite crystals.

They also found that the lone presence of OMM-64 resulted in a mix of polymorphs (including vaterite) and Otolin-1 by itself resulted in the production of pure calcite. The authors suggest that either Otolin-1 and OMM-64 act synergistically to produce aragonite, or that an additional unidentified aragonite inducer is present in the high molecular weight aggregate.

58

Starmaker-like

A third IDP thought to be functionally homologous to Stm was identified in Oryzias latipes

(medaka) (Bajoghli et al., 2009), termed Starmaker-like. Despite having a markedly lower molecular weight of 39.6 kDa, Stm-l still appears to play a similar role in aragonite precipitation

(Rozycka et al., 2014). Like OMM-64 and Stm, it is rich in both aspartic and glutamic acid, and is likely a phosphoprotein. In a blast alignment, it showed virtually no homology with either of those two proteins (accessed on February 14 2018 https://blast.ncbi.nlm.nih.gov/blast/Blast.cgi).

Structurally, Stm-l is remarkably extended, highly asymmetrical and not folded. It was found through several different methods to have properties indicating that is a monomeric coil-like IDP

(Rozycka et al., 2014). Like OMM-64, Stm-l compacts in the presence of calcium ions, with little change to its secondary structure (Rozycka et al., 2014). Considering the reported disparity between ionic concentrations of the proximal and distal endolymph (Payan et al., 2004), the local concentration of calcium may have a vital effect on the shape of Stm-l, and thus influence its ability to modulate biomineralization. Interestingly, Rozycka et al. (2014) also found that its structure, however, is strongly influenced by external physical factors such as temperature, indicating its propensity for folding and forming locally ordered structures. While being rich in disorder-promoting residues, Stm-l also has numerous methionine residues. These are normally in the hydrophobic core of a protein, and, as Stm-l lacks this, may be acting at the surface in some as- of-yet uncharacterised way. The authors hypothesise that, given the propensity for oxidation of methionine, these surface residues may be present as sulfoxides, allowing Stm-l to be more susceptible to proteolysis, and thus playing a role in its spatial and temporal regulation in vivo.

59

Finally, as with Stm, Stm-l is likely to act as a template for calcium carbonate crystal nucleation as well as an inhibitor of crystal growth. It was shown that by increasing concentration of Stm-l, the total number of crystals substantially increased whereas the size of crystals decreased becoming smaller and more rounded. The authors also found the presence of a small number of vaterite crystals, regardless of protein concentration, that lost integrity after 48h (Rozycka et al., 2014). In terms of crystal nucleation, one of two mechanisms may be at play: Firstly, the protein may be acting as a physical template, coordinating and controlling direction of crystal growth through specific ion-protein interactions. Secondly, as suggested by Rozycka et al. (2014), the flexible IDPs may in fact be following the mechanism proposed by Gower (2008), where they induce the formation and stabilisation of amorphous calcium carbonate before a reordering of the crystal occurs, creating aragonite.

Extracellular serine/threonine protein kinase FAM20C

Extracellular serine/threonine protein kinase FAM20C (FAM20C) has been identified in the otoliths

(but not endolymph) of black bream (Thomas et al., 2019). FAM20C specifically catalyses the addition of phosphates to biomineral extracellular matrix proteins such as DSPP (Tagliabracci et al.,

2012). It contains two Mn2+ ions as cofactors (Xiao et al., 2013). During LA-ICP-MS analyses of otolith cores, both manganese and phosphorous have been found to be concentrated in the primordium (Brophy et al., 2004, Shima and Swearer, 2009). Furthermore, Thomas et al. determined that phosphorous was exclusively found in the proteinaceous fraction of the endolymph (Thomas et al., 2017). Considering the relationship between phosphorylation of Stm and its function, it is reasonable to suppose that a kinase regulating the degree of phosphorylation would be found. The functional homolog of Stm (the IDP DSPP) is specifically phosphorylated by

60

FAM20C. Consequently, Thomas et al. (2019) speculated that FAM20C is likely to also have Stm

(and its homologs) as substrates.

In terms of structure and mode of action, FAM20C is a casein kinase that is thought to localize within the Golgi apparatus where it is then secreted into the extracellular space. It specifically targets a Ser-X-Glu/pSer motif on its substrate (where X is any amino acid and pSer is phosphoserine (Tagliabracci et al., 2012)). FAM20C’s structure (i.e. the position of its αC helix and a lack of an activation loop) indicates that it is highly efficient in terms of catalytic activity. FAM20C specifically acts to phosphorylate members of the small integrin-binding ligand, N-linked glycoprotein (SIBLING) family, which are found in the extracellular matrices of dentin and bone, and thus regulate their effectiveness in biomineralization (Tagliabracci et al., 2012). FAM20C activity is regulated upstream by another secreted Golgi kinase, FAM20A (Ohyama et al., 2016).

Interestingly, mutations in the FAM20C gene lead to Raine syndrome, a form of osteosclerotic bone dysplasia (Tagliabracci et al., 2012), whereas alterations to FAM20A cause disruption to enamel formation and unwanted calcification in the kidney (Ohyama et al., 2016).

Thomas et al. (2019) did not find FAM20C in endolymph harvested from adult black bream. This suggests that its expression is either circadian (and was absent at the time of capture) or is only expressed in the inner ear during early development. Thomas et al. (2019) hypothesised that

FAM20C acts during nucleation to phosphorylate IDPs (e.g. Stm, Stm-l, OMM-64). They additionally found both phosphorylated and unphosphorylated Stm-l residues present in the otolith. This points to a tight control of Stm-l activity depending on its degree of phosphorylation (i.e. the number of phosphates substituted per molecule of Stm-l). As described above, increased phosphorylation causes IDPs to be able to bind more calcium ions, compacting their tertiary structure, allowing closer packing and thus exerting greater regulation over aragonite growth

(Wojtas et al., 2012).

61

Otoconin-90-like

In vertebrates, the initial seeding of otoconia or otoliths is thought to occur via otoconins, proteins that share homology with Secretory phospholipase A2 (PLA2). Despite retaining the calcium- binding ability of PLA2, otoconins generally lack its enzymatic ability (Pote and Ross, 1991, Wang et al., 1998b). Otoconins tend to be the dominant protein in otoconia, accounting for 90% of their mass, and are thought to act as templates for crystal growth, setting the specific polymorph that is present. In mammals, the protein responsible for calcitic otoconia is Otoconin 90 (Wang et al.,

1998b, Lu et al., 2010, Thalmann et al., 2011), whereas Otoconin 22 is thought to cause aragonitic otoconia in amphibians (Pote et al., 1993). In mammals, Otoconin 90 has been found to both interact with Otolin as well as trigger its expression (Zhao et al., 2007, Deans et al., 2010).

In zebrafish, Petko et al. (2008) identified Otoc1, a gene encoding the Otoconin 90 ortholog. Otoc1 was found to be extremely acidic, with several potential glycosylation sites. Similarly to other otoconins, it can bind calcium, and is thought to act as a scaffold during aragonite precipitation. In contrast to Otoconin 22, Otoc1 contains two PLA2 domains rather than one. These PLA2 domains do not have any enzymatic ability, but do conserve a high number of cysteine residues, suggesting that disulfide crosslinking may be a key feature of its structure.

In terms of its expression, Petko et al. (2008) detected Otoc1 as early as 15 hpf in the otic vesicle, midbrain-hindbrain boundary, optic stalk and spinal cord of zebrafish. It continued to be expressed in the inner ear, albeit in decreasingly lower levels, until halting at about 5 dpf. Considering that zebrafish OMP-1 first appears in low levels in the inner ear slightly after 15 hpf, and then is expressed throughout an individual’s life, it is speculated that Otoc1 precedes OMP-1, and acts to initiate matrix formation. This pattern of expression – Otoconin 90-like preceding OMP-1 – has also been observed in larval Mozambique tilapia (Weigele et al., 2017).

62

Petko et al. (2008) also investigated the effects of knocking out both Otoc1 and OMP-1 in zebrafish. Knocking out the former resulted in smaller, absent and extra otoliths, with the phenotypes shown being far stronger than those produced by OMP-1 knockout. As knocking out

OMP-1 has been shown to result in smaller otoliths (Murayama et al., 2005), the authors also tested for any possible interactive effects by co-injecting embryos with both OMP-1 and Otoc1 morpholinos. They found that the phenotypes produced by Otoc1 knockout were exacerbated when OMP-1 was also knocked out, suggesting synergy between these two proteins. Finally, they also found that Otoc1 had no effect on development of kinocilia or cell bodies. Taken together, this suggests that Otoconin 90-like proteins are necessary for otolith seeding and recruitment of other matrix proteins, whereas OMP-1 variants are required for organic matrix growth.

Secreted protein acid rich in cysteine (Sparc)

Sparc (sometimes called Osteonectin) is a glycoprotein found in basement membranes of vertebrate endothelial and osteoblast cells that plays a role in tissue growth (Kelm et al., 1994).

Specifically, it acts to modulate the assembly and maintenance of the extracellular matrix of bone, teeth and cartilage. It was originally thought to solely exist in these tissues but has since been identified in the matrices of both otoconia and otoliths from a variety of fish species (Kang et al.,

2008, Xu et al., 2010, Weigele et al., 2016, Weigele et al., 2017). While Sparc does play a role in fish bone growth (Weigele et al., 2015a), its presence in the otolith is of interest due to its multifunctional nature, and likely plays more than one role in deposition of the discontinuous zone.

63

Sparc has been found to be expressed in the inner ear long before the formation of the primordium. In a study on development of Mozambique tilapia, Weigele et al. (2016a), found that it was expressed in the otic cavity as early as Developmental Stage 7, with its expression appearing to halt by Stage 21. In that study, however, only sub-adult and larval fish were examined, so it is unclear whether Sparc expression recommences after Stage 25 (the endpoint of the study).

Similarly, Kang et al. (2008) found that, in zebrafish, Sparc was strongly expressed in the otic placode by 14 hpf whereas Nemoto et al. (2008) reported its expression by Stage 22 in the otic vesicle of medaka. Although Sparc is clearly expressed early in inner ear development, it is unlikely to be involved in seeding, as its knockout in zebrafish, whilst causing aberrance in otoliths, does not occur until 24 hpf, well after the formation of the primordium (Kang et al., 2008). Sparc has also been detected in otoliths from adult zebrafish, rainbow trout, and channel catfish (Kang et al.,

2008) as well as black bream (Thomas et al., 2019). However, its presence in adult otoliths is not necessarily an indication of Sparc being expressed after early inner ear development. Proteomic analyses of otoliths allow for detection of all proteins expressed throughout an individual’s lifespan, but currently lack the ability to resolve these proteins temporally. However, Thomas et al.

(2019) identified Sparc in deep sequence proteomic data gathered from endolymph harvested from adult black bream, indicating that its expression may continue throughout an individual’s life.

Consequently, Sparc may solely be involved in early otolith development or alternatively play a more vital role in increment deposition.

In terms of its structure, Sparc contains domains that allow it to bind to several different substrates. It can bind calcium ions readily, both at its C-terminus EF-hand domain and its acidic N- terminus domain. In addition to calcium binding, the EF-domain readily binds collagens. Finally, it has been found to also interact with plasminogen, playing a key role in regulating matrix

64 metalloproteinases, enzymes that specifically catalyse the proteolysis of collagens (Kelm et al.,

1994).

It is therefore speculated that Sparc acts firstly as a linking molecule, coordinating calcium ions to collagen, and thus promoting biomineralization. As well as this, it is thought to regulate organic matrix assembly through anchoring plasminogen to otolin-1, as well as enhancing the activity of plasmin, its active form. Finally, it may also be acting antagonistically with neuroserpin, as described below (Kang et al., 2008, Thomas et al., 2019).

Neuroserpin and its downstream targets

Neuroserpin (also called SerpinI1), has been found in the otoliths of three separate fish species; adult rainbow trout (Kang et al., 2008), juvenile Mozambique tilapia (Weigele et al., 2016), and adult black bream (Thomas et al., 2019). As a member of the Serpin superfamily, it is an inhibitor of serine proteases (Silverman et al., 2001). Specifically, it has been shown to regulate proteolytic cascades through direct inhibition of tissue type Plasminogen activator (tPA) (Lee et al., 2015). tPA is a serine protease whose main function is to convert other proteases into their active form. It, along with urokinase Plasminogen activator (uPA), can catalyse the conversion of Plasminogen to

Plasmin. Plasmin, then, can act to activate a broad range of collagenases, the matrix metalloproteinases, a family that play key roles in development and resorption of biomineralized tissues (Somerville et al., 2003, Visse and Nagase, 2003, Page-McCaw et al., 2007). Thomas et al.

(2019) identified tPA, Plasminogen, and Matrix metalloproteinase 2 (MMP 2, sometimes called

Type IV Collagenase) in the inner ear of black bream. They also detected Tissue Inhibitor of

Metalloproteinase 2 (TIMP 2), which regulates MMP 2. The authors speculated that Neuroserpin, along with TIMP 2, regulate the proteolysis of Otolin-1 C1q domain hubs, its presence allowing continued deposition of the discontinuous zone. As of time of writing, it is unclear whether

65

Neuroserpin’s expression follows a circadian rhythm, nor whether its absence results in malformation of the otolith. An investigation into the timing of Neuroserpin expression, both in terms of ontogeny and diel rhythms, as well as knockout studies, are fruitful lines of future research.

Otopetrin 1

Otopetrin 1, a multitransmembrane domain protein, is highly conserved across vertebrates

(Hughes et al., 2008, Hurle et al., 2011). It is thought to be involved in intracellular calcium regulation and protein trafficking during the critical seeding period of both otoconia and otoliths

(Hurle et al., 2003, Hughes et al., 2004). In mammals, it has been found to be expressed in complement with Otoconin 90/95, with mutations in OTOP1 (the relevant gene) leading to aberrant otoconia phenotypes and severe balance disorders (Hurle et al., 2003). In mice, Otopetrin

1 is expressed in the supporting cells of the sensory macula (Kim et al., 2010), whereas in fish it is expressed predominantly in hair cells (Hughes et al., 2004, Söllner et al., 2004). It is speculated that Otopetrin 1 acts, in mammals, by sensing calcium concentration adjacent to epithelial supporting cells as well as responding to ATP levels in the endolymph, and correspondingly regulates intracellular calcium levels during otoconia formation (Hughes et al., 2007, Kim et al.,

2010).

In contrast to this, in zebrafish, otopetrin 1 is first expressed at 18 hpf, with its localisation being restricted to the sensory epithelial hairs by 24 hpf. This timeframe represents the critical period of otolith seeding in zebrafish (Riley et al., 1997). As the inner ear develops, Otopetrin 1 continues to be expressed, showing a pattern consistent with the precursor hair cell formation. It greatly reduces in level of expression by 5 dpf, halting completely by 7 dpf (Hughes et al., 2004). The

66 development of the zebrafish inner ear is complete by 5 dpf, with the maculae and semicircular canals being completely formed (Whitfield et al., 2002). This expression pattern suggests that

Otopetrin 1 plays a role in otolith nucleation, but not increment formation.

Otopetrin 1 knockout studies identified the backstroke mutant, a phenotype in which otoliths are absent (Söllner et al., 2004). The backstroke mutation specifically results in the change of a single residue – the hydrophilic glutamic acid being changed to the hydrophobic valine. It is hypothesised that this mutation results in a subsequent misfolding of Otopetrin 1, rendering the protein ineffective. Backstroke mutants, in addition to being deficient in otoliths, showed expression patterns indicating that Stm was not secreted into the extracellular space, being instead restricted to the macular epithelial cells. Although backstroke mutants lack otoliths initially, they have been found to form late, albeit malformed. Interestingly, although otoliths were affected by OTOP1 mutation, no other inner ear structures were adversely affected (Hughes et al., 2004). Considering this, and the postulated ten transmembrane domains in Otopetrin 1, it is thought that it serves an additional function to intracellular calcium regulation, that of trafficking Stm to the otolith during seeding (Söllner et al., 2004).

Otolith membrane proteins

Otogelin, otogelin-like

Alpha- and Beta-tectorin, Otogelin and Otogelin-like are components of the tectorial membrane, a gel-matrix covering the tips of bundles of stereocilia in the organ of Corti of mammals (Goodyear and Richardson, 2002, Jovine et al., 2005). Fish, however, do not have a tectorial membrane.

Otogelin and otogelin-like are essential for anchoring the otolith to the macula (Cohen-Salmon et al., 1997, Wang et al., 2011, Yariz et al., 2012, Stooke-Vaughan et al., 2015, Weigele et al., 2016).

67

Structurally, like other tectorial membrane proteins (e.g Alpha-tectorin), they contain multiple von

Willebrand factor type D domains (VWD), which are important to cell adhesion (Lundberg et al.,

2006). In zebrafish early development, otogelin is expressed in regions corresponding to the two sides of the otic vesicle. As the larva develops, its expression slowly shifts to the ventral transitional epithelium. By 2-4 dpf, Otogelin is expressed in the cristae of the inner ear, as well as in supporting cells at the dorsal edge of the macula (Wang et al., 2011, Stooke-Vaughan et al.,

2015). Otogelin-like is expressed similarly to Otogelin, being found in the saccular hair cells as well as supporting cells of the macula. Knockdown of Otogelin-like in zebrafish leads to morphants with sensorineural hearing loss (Yariz et al., 2012).

In addition to being part of the otolith membrane, they may also be a component of otolith precursor particles (OPPs), fundamental to the seeding step of otolith formation. During seeding, mRNA expression of otogelin suggests that, like in the inner ear of mammals, it is expressed into the lumen of the otic vesicle, where it is then free to interact with tether-cell specific binding factors. Stooke-Vaughan et al. (2015) found that knockout of the gene that codes otogelin (OTOG) caused the disruption of seeding, leading to the einstein (single otolith) phenotype. They speculated that otogelin either acts as an OPP component, causing tether cilia recognition or forms a tethering complex with an integrin, α8β1. In this latter scenario, the tethering complex then interacts with the tips of tether cilia. They further suggested that otogelin secretion or processing was carried out by GP96 (also called Hsp90β1).

GP96

GP96, as a chaperonin, protects against heat-stress. However, in addition to this, it also facilitates the folding of other proteins, including integrins. Knockout of GP96 resulted in a single large, untethered otolith alongside either absent or substantially smaller paired otoliths (Sumanas et al.,

68

2003). In morphants, the OPPs did not adhere properly to kinocilia, instead becoming large,

“clumped” masses that remained untethered. As no other changes were observed to the architecture of the inner ear, it is thought that GP96 affects only otolith adhesion. It is likely that

GP96 processes integrins required for otolith seeding, as GP96-deficiency in mice causes the disruption of integrin secretion (Randow and Seed, 2001). Alternatively, as suggested by Lundberg et al. (2006), GP96 may also be processing otogelin and other OPPs.

Tectorins

Both Alpha and Beta tectorin have been identified in the teleost inner ear – either through knockout studies (Whitfield et al., 1996, Wang et al., 2011, Yang et al., 2011, Stooke-Vaughan et al., 2015), immunohistochemistry (Yang et al., 2011, Wang et al., 2011, Weigele et al., 2017) or proteomics (Weigele et al., 2016, Thomas et al., 2019). Alpha tectorin is thought to be essential to maintaining the interface between the otolith and the macula following seeding (Stooke-Vaughan et al., 2015). Beta tectorin is likely not a constituent of the otolith membrane, as it is expressed in the sensory patch of the macula (Yang et al., 2011).

Structurally, both tectorins contain a Zona Pellucida (ZP) domain. The ZP domain was initially characterised as being responsible for spermatozoa adhesion to the oocyte membrane of mammals, but has since been found to facilitate membrane adhesion in a number of different tissues (Yang et al., 2011). Beta-tectorin largely consists of this ZP domain (256 of its 336 amino acids), whereas Alpha-tectorin also has several VWD domains as well as an N-terminus Nidogen domain. This latter domain of Alpha-tectorin has been proposed to interact with the collagen domain of Otolin-1 allowing for adhesion of the otolith to kinocilia (Lundberg et al., 2006). Alpha- tectorin is considerably larger than Beta-tectorin, being over 2000 residues in length. Furthermore,

BLAST analyses of both Alpha- and Beta-tectorin reveal high homology (>70%, accessed on

69

February 14 2018 https://blast.ncbi.nlm.nih.gov/blast/Blast.cgi) between bony fish variants suggesting that they are highly conserved proteins. Both Alpha- and Beta-tectorin have been found in the otolith matrices of Mozambique tilapia (Weigele et al., 2016) and black bream

(Thomas et al., 2019). However, in both cases, their presence may not be indicative of being a

“true” matrix component but could be the result of constituents of either the otolith membrane or kinociliae being inadvertently incorporated during increment formation (Weigele et al., 2016,

Thomas et al., 2019).

Alpha-tectorin is expressed in the macula of zebrafish, initially appearing at 1 dpf, and becoming restricted to the edges of the maculae by 3 dpf. Its expression substantially decreases by 5 dpf and is restricted to the sensory cells (Wang et al., 2011, Stooke-Vaughan et al., 2015). A similar expression pattern for Alpha-tectorin has been observed in O. mossambicus (Weigele et al., 2017).

Unlike Otogelin, Alpha-tectorin is never expressed in the ventral epithelium of the otic vesicle.

Knockout of Alpha-tectorin leads to the rolling stones (rst) mutant, where the otolith is not tethered to the macula during development and is instead loose within the ear (Whitfield et al.,

1996, Stooke-Vaughan et al., 2015).

In zebrafish, Beta-tectorin shows a similar expression pattern to Stm, being expressed exclusively in the posterior and anterior cells of the macula (Söllner et al., 2003, Yang et al., 2011). Yang et al.

(2011) also showed that knockout of Beta-tectorin resulted in zebrafish mutants that had either solitary, or fused otoliths and were unable to balance or float. Clearly, Beta-tectorin, plays a role in inner ear development, although the specifics of this is still unclear. The authors speculated that

Beta-tectorin (as it is non-collagenous but knockout still produces fused otoliths) might polymerise with Otolin-1 and/or Stm into the otolith matrix during increment formation.

70

Myosins

Myosins are calcium-binding proteins that are typically associated with motor action. Myosins have been detected in teleost otoliths and endolymph, as well as in the otoconia of mammals

(Thalmann et al., 2006). Weigele et al. (2016a,b) identified the presence of Myosin Light Chain IX in the otoliths of O. mossambicus and Coffin et al. (2002) found Myosin VI and VIIa throughout the epithelia of the sacculum, the lagena and the utricle of D. rerio. Thomas et al. (2019) also found a range of myosins (Types I, II, III, VI, VII, IX and X), albeit their presence was restricted to the endolymph of A. butcheri. They did, however, find evidence of Myosin XV in A. butcheri otoliths.

Myosin XV has been previously identified in the stereocilia of the inner ear of humans (Wang et al.,

1998a). Although myosins cannot be ruled out as potential otolith matrix proteins, it is more likely that they are the result of hair cell proteins being inadvertently incorporated into the otolith through endolymph being trapped during increment formation. This is likely considering their motor function in other systems, coupled with their prevalence in the endolymph (Thomas et al.,

2019).

Proteoglycans

Inner ear proteoglycans have received relatively little attention, despite being fundamental to coordination of extracellular matrix collagens during biomineralization (Arias et al., 2004). In terms of structure, a proteoglycan typically has a protein core which is bonded to unbranched mucopolysaccharides called glycosaminoglycans (GAGs). These GAGs are usually chondroitin sulfate (CS), dermatan sulfate (DS), heparin, heparan sulfate (HS), keratan sulfate (KS), or hyaluronan. Hyaluronan is the only non-sulfated GAG (Arias et al., 2004). The fundamental difference between proteoglycans and glycoproteins, then, is the length and branching of the

71 bonded sugar chain. While proteoglycans typically have an attached GAG, glycoproteins typically have short and/or branched substituents.

The presence of proteoglycans in the otolith organic matrix was first suggested by Mugiya, who used histochemistry to identify both sulfated and unsulfated polysaccharides in the endolymph

(Mugiya, 1966) and otoliths (Mugiya, 1968) of rainbow trout. Later, other studies confirmed the presence of proteoglycans in the fish inner ear, but without the specifics of which proteoglycans were present (Asano and Mugiya, 1993, Borelli et al., 2001, Borelli et al., 2003b). Thomas et al.

(2019), however, detected the presence of several proteoglycan cores in the otoliths of adult black bream. In that study, they identified the presence of biglycan, decorin, osteoglycin (mimecan), epiphycan and aggrecan. The role that proteoglycans play during biomineralization of the otolith, however, remains unclear. In other systems (i.e. teeth and bone), these proteoglycans bind to collagen fibres in the extracellular matrix (Iozzo, 1998, Wang et al., 2005, Haruyama et al., 2009,

Ye et al., 2012, Zvackova et al., 2017). As well as Otolin-1, a short chain collagen, other collagens

(particularly types I and II) have been identified in the inner ear of fish (Nemoto et al., 2008,

Thomas et al., 2019). Considering this, it is likely that the functions of proteoglycans in the otolith matrix are similar, and act to stabilise collagenous fibrils within the matrix.

MODELS OF OTOLITH GROWTH

The biological mechanisms of otolith formation and subsequent growth are of great interest, particularly as these mechanisms underpin any inferences made from otoliths into the biology and ecology of fishes and their management. While we are beginning to be able to identify the various biomolecule players involved, there are still relatively few conceptual biochemical models that satisfactorily explain their specific functions and roles in otolith biomineralization. As discussed

72 above, much of the earliest work on otolith biomineralization was focussed on the mechanisms of calcium and bicarbonate ion supply. These early studies contributed greatly to our understanding of what happens during otolith formation, but not how it happens. I summarise below three different important mechanisms in otolith biomineralization: (1) Otolith tethering, the earliest stage of otolith nucleation – so called because of the attachment of the nascent otolith to sensory kinocilia in the inner ear, and fundamental to its correct anatomical function; (2) Regulation of crystallisation – the control over the direction, size and geometry of calcium carbonate during precipitation and; (3) Diel-entrained increment deposition – the apparent circadian growth of mineral-rich and protein-rich bands.

Otolith tethering

The first stage of otolith biomineralization is tethering, the expression of its first constituents from epithelial cells and their subsequent attachment to specialised hair cells in the inner ear (Riley et al., 1997, Yu et al., 2011, Stooke-Vaughan et al., 2012). Stooke-Vaughan et al. (2015) have suggested that tethering is a two-step process. Firstly, otolith precursor particles (OPPs) bind to kinocilia (i.e. “seeding”) in the macula. The OPPs are secreted by sensory and transitional epithelial cells, are glycoprotein-rich and may also contain glycogen as well as Cadherin 11, an osteoblastic cell adhesion protein (Pisam et al., 2002, Clendenon et al., 2009). There are several proteins that are likely to be OPP candidates, based on expression profiles and biochemistry. These are, Stm,

Beta-tectorin, OMP-1 and Otoconin 90-like. As tethering occurs, one or more of these OPP candidates likely sequester Ca2+ ions for crystal formation. Otoconin 90, for example, is thought to form free-floating “seeds” with calcium-carbonate during otoconia formation in mammals

(Lundberg et al., 2006). In the fish sacculum, the seeds are thought to be attracted to kinocilia

73 either through surface-charge interactions or kinetically, from endolymph movement produced from the gentle beating of cilia (Riley et al., 1997).

In zebrafish, OPPs appear within 18 hpf, and solely bind to the ends of the kinocilia of nascent hair cells (sometimes called tether cells) (Riley et al., 1997, Tanimoto et al., 2011). Otogelin is thought to be essential for this first step. This seeding process has been suggested to be also mediated by otolith precursor binding factors, though the specifics of these factors are not yet known (Riley et al., 1997, Yu et al., 2011, Stooke-Vaughan et al., 2012). The second step of tethering is the adhesion (and continued adhesion) of the otolith to the sensory patch of the macula via the otolith membrane (Stooke-Vaughan et al., 2015). This has been suggested to occur through the interaction of the nidogen domain of Alpha-tectorin in kinocilia with the collagen domain of

Otolin-1 in the otolith membrane (Lundberg et al., 2006). Following tethering, biomineralization can commence, with the first increment forming. In order for increment formation to occur, epithelial cells must increase secretion of both calcium and bicarbonate into the endolymph following the mechanism outlined earlier. It is thought that the organic matrix proteins

(specifically OMP-1, Stm (or its homolog), and Otolin-1) act synergistically to control formation of the crystal lattice. It has been suggested that Otolin-1 acts, with Pcbl, as a “hub” for coordination of other organic components of the matrix whereas OMP-1 and Stm provide a structural scaffold and regulation of crystallization respectively (Thomas et al., 2019).

Regulation of crystallisation

An inhibitor of crystallisation has been identified in otolith matrices in several studies (Wright,

1991, Borelli et al., 2001, Guibbolini et al., 2006). The method for detection employed by these studies is through observation of pH, following the method of Wheeler et al. (Wheeler et al.,

74

1981). In vitro precipitation of calcium carbonate is accompanied by a drop in pH due to the evolution of protons (Eq.2).

In these studies, the presence of the inner ear proteins in vitro is associated with a maintenance or slight increase in pH followed by a lower rate of pH decrease when compared with controls. From this it is inferred that the presence of a crystallisation factor in the inner ear stops unrestricted precipitation of calcium carbonate on the proximal face of the otolith, allowing for proper growth

(Guibbolini et al., 2006). Furthermore, in a study conducted on turbot (Psetta maxima) by

Guibbolini et al. (2006), this organic inhibitor was detected in the low molecular weight fraction of the endolymph. They determined the inhibitor to be water-soluble and polyanionic, with an approximate molecular weight of 20 kDa (Borelli et al., 2003b, Guibbolini et al., 2006). Although the authors speculate that it may be a proteoglycan, its identity and sequence remain unknown.

As the homolog of Stm has not been identified in turbot, two possible interpretations of these results exist. Firstly, that the Stm homolog in turbot is roughly 20 kDa, which is significantly smaller than all other previously identified Stm homologs. Alternatively, as small molecule proteoglycans have been positively identified in the otolith matrix (Mugiya, 1968, Borelli et al., 2003b, Thomas et al., 2019), it is possible that one or more of these is also involved in regulation of crystal growth.

As discussed earlier, Stm (and its homologs) has been found to both encourage calcium carbonate nucleation while simultaneously inhibiting its crystallisation. Stm’s effectiveness as a biomineralizing protein is thought to be a function of its degree of phosphorylation (Wojtas et al.,

2012). As the number of phosphates added to the IDP increases, the overall surface charge of the protein becomes more negative. This facilitates increased binding of calcium ions to the IDP, and as a result, both concentrates and orders calcium ions at the growing increment. Additionally, as calcium binding has been shown to cause the compaction of the tertiary structure of otolith IDPs, this subsequently allows for tighter packing of the IDP within the organic lattice. This allows for

75 both greater control over direction as well as size of crystal growth. An automatic consequence of this is also the selection of the crystal polymorph, aragonite, through the regulation of the ordering of the crystal geometric structure. This model agrees with similar biomineralization systems in other species examining the role of phosphorylation in crystallisation inhibition (Borbas et al., 1991). Considering that both phosphorylated and unphosphorylated forms of IDP peptides were found present in the otoliths of black bream (Thomas et al., 2019), it is likely that a kinase such as FAM20C is regulating and modifying IDP function through modification, allowing greater flexibility of otolith growth depending on life stage.

Diel-entrained increment deposition

The rhythmic deposition of protein-rich and mineral-rich bands has been observed in a wide number of otolith studies (Brothers, 1978, Mugiya et al., 1981, Mugiya, 1984, Mugiya, 1987,

Saitoh, 1993, Edeyer et al., 2000, Tohse et al., 2006, Dauphin and Dufour, 2008). Despite this, only one explanation for the regulation of the apparent circadian rhythm of increment deposition has been suggested. Thomas et al. (2019) suggested that, as Otolin-1 likely acts as an interaction hub for the organic matrix, lysis of this key protein would lead to the cessation of growth of the discontinuous zone. As described above, they suggest that a collagenase found enriched in the otolith, MMP 2, acts to cleave Otolin-1 hubs, halting further coordination of organic molecules.

They further speculated that MMP 2 activity was jointly regulated; from above, by the action of

Neuroserpin, a well characterised regulator of proteolytic cascades (Silverman et al., 2001), and from below, through its deactivation by TIMP 2 (Thomas et al., 2019).

76

IMPLICATIONS AND FUTURE DIRECTIONS

Otolith biomineralization mechanisms underpin the assumptions made when interpreting data from otoliths. As an example, a critical assumption in the use of otoliths in the reconstruction physico-chemical histories is that they are metabolically inert and do not undergo remodelling

(Campana, 1999, Campana and Thorrold, 2001). However, while this has been observed, the reasons for this are largely unclear. It is imperative that we deepen our understanding of these mechanisms. Below, I outline the implications of biochemistry to the three main areas of otolith studies, and suggest future directions of research.

Hearing

The anatomical function of otoliths within the inner ear of fish is as a sensor of both gravity and sound (Schulz-Mirbach et al., 2018). It accomplishes this through the shearing force produced as an individual otolith oscillates against the macular cilia (Popper et al., 2003). Oxman et al. (2007) suggested that correct oscillation of the otolith was directly a result of its crystal polymorph.

Reimer et al. (2016) demonstrated that substitution of one polymorph for another (e.g. vaterite for aragonite) resulted in up to a 50% reduction in hearing sensitivity in individual fish. Vaterite

(2.54 g cm-3) (Tomas and Geffen, 2003) is slightly less dense than aragonite (2.93 g cm-3) (Campana and Thorrold, 2001), and will consequently have a smaller amount of inertia, showing an undesired change in the otolith’s ability to oscillate. As discussed above, it is likely that the

“selection” of one polymorph of calcium carbonate over another is protein-mediated. It is therefore vital that we understand the mechanisms of polymorph selection. One fruitful line of research is into the phosphorylation of otolith matrix proteins, given it is likely key to crystal regulation and nucleation (including polymorph selection). While the otolith proteome has been

77 elucidated for the sagittae of black bream (Thomas et al., 2019), much can be gained from a comparison of the occurrence, relative abundance and post-translational modification of sagittae

(aragonite) matrix proteins with those of asterisci (vaterite). Such a study would be well complemented by genetic manipulation studies (i.e. knock-out), allowing for the clarification of individual protein roles in polymorph selection.

Otolith growth

Although the mechanisms of early otolith formation are beginning to be understood, the regulation of subsequent growth (i.e. increment deposition) is poorly described, with only untested hypotheses being offered as explanations for observed circadian and seasonal growth.

The measurement of daily growth increments is a fundamental tool in effective fisheries management. As increment widths are likely determined by relative levels of inner ear protein expression, a comprehensive understanding of biomineralization mechanisms would greatly aid in the interpretation of otolith microstructure. This, and other hypotheses suggested by Thomas et al. (2019) could be tested through immunohistochemical labelling studies of the piscine inner ear, with the view to determining the timing and relative abundance of expression of identified key proteins.

Otolith chemistry

The presence of metal-binding proteins (metalloproteins) in otoliths can be problematic in environmental reconstructions, as they may lead researchers to infer physico-chemical history from an endogenous event. To date, there have been only two studies examining the presence of

78 metalloproteins in otoliths (Miller et al., 2006, Izzo et al., 2016a) and one study determining metal- protein interactions in the endolymph (Thomas et al., 2017). In the study conducted by Miller et al.

(2006), they found that Zn and Cu were bound to soluble otolith proteins, whereas Mn was either not bound, or only weakly bound. Izzo et al. (2016a) found similar results for Zn and Cu but also found that Mn (as well as 6 other elements) were bound to otolith proteins. Through size exclusion chromatography-inductively coupled plasma-mass spectrometry (SEC-ICP-MS) of endolymph of black bream, Thomas et al. (2017) identified 15 trace elements interacting with protein components (Table 2.2).

Table 2.2. Summary of trace elements found in otoliths from cod (Gadus morhua, Miller et al., 2006) or saddletail snapper (Lutjanus malabaricus, Izzo et al., 2016) and endolymph from black bream

(Acanthopagrus butcheri, Thomas et al., 2017). Trace elements are categorized as being either solely bound to proteins (“protein-bound”), as free ions (“unbound”) or present in both forms. Please note that while Izzo et al. (2016) found nearly all assessed trace elements both as free ions and interacting with proteins, the four transition elements (Cu, Zn, Fe and Mn) occurred in high percentages in the protein fraction. LOD = limit of detection.

Tissue Otolith1 Otolith2 Endolymph3

Protein-bound elements Cu, Zn Fe, Co, Ni, Cu, Zn, Cr, V,

Hg, Cd, Sn, Pb, P

Unbound elements Mn (Below LOD) Al, Pb, Cd Li, K, Rb, Mg, Mn, I

Both Rb, Li, Mn, Cu, Zn, Fe, K, Sr, Ba Sr, Ba, As

1Miller et al., 2006; 2Izzo et al., 2016; 3Thomas et al., 2017

The presence of metalloproteins in the inner ear has implications for the use of otoliths in reconstructing physico-chemical histories. Indeed, it is likely that certain elements (particularly those of the transition series, which are well-characterised as protein factors) are indicative of

79 endogenous processes rather than exogenous ones (Otake et al., 1997, Funamoto and Mugiya,

1998, Brophy et al., 2004, Loewen et al., 2016). The link between physiological processes and otolith chemistry has been explored by several studies (Limburg and Elfman, 2010, Sturrock et al.,

2012, Sturrock et al., 2014, Sturrock et al., 2015, Limburg et al., 2018). Zinc, particularly, is an element which should be treated with caution. The otolith contains multiple Zn2+-binding proteins at high abundances, including both MMP 2 and CA 1 and 6. Considering the abundance of these otolith proteins, coupled with the absence of free-ion Zn2+ in the endolymph, it is far more likely to be an indicator of endogenous processes (such as growth) than environmental change. Thomas et al. (Chapter 5) compared cluster analyses of otolith microchemistry that determined clusters depending on three different sets of elements: (a) All elements assayed (7Li, 11B, 24Mg, 31P, 34S, 39K,

55Mn, 88Sr, 63Cu, 66Zn, 138Ba, 208Pb); (b) Elements primarily in the salt-fraction of the endolymph (7Li,

11B, 24Mg, 39K, 55Mn, 88Sr, 138Ba); and (c) Elements primarily bound to proteins in the endolymph

(31P, 34S, 63Cu, 66Zn, 208Pb). In that study, they found that salt-fraction elements alone were better able to resolve distinct spawning and larval nursery locations among cohorts and recruit locations

(Chapter Five). Future studies aimed at localisation of metalloproteins in both annuli and daily increments may prove invaluable and will allow for not only better physico-chemical reconstructions, but also for a greater understanding of the temporal dynamics of endogenous processes.

CONCLUSION

The otolith still has a great deal more to tell us. Thomas et al. (2019) found that although over 380 proteins could be found in black bream otoliths, the majority of these were likely to be uninvolved in biomineralization mechanisms. These “non-functional” proteins are trapped, much like trace

80 element ions, during the basic process of increment formation. For example, Thomas et al. (2019) found evidence in otoliths of Vitellogenin, an oestrogen-stimulated protein involved in egg yolk formation (Funamoto and Mugiya, 1998, Reading et al., 2017). An extremely fruitful line of research is the reconstruction of endogenous events from the otolith, provided one can determine the specifics of protein location in increments. There are currently difficulties in attempting this. In a normal proteomic analysis, proteins are uncoupled from their location in the otolith through dissolution techniques. If the mineral phase is kept intact, and the otolith polished to expose increments, the extremely high level of Ca interferes with assessments of protein location that might be undertaken through approaches traditionally used for tissue analysis (e.g. MALDI imaging). To deal with these two issues, then, the otolith mineral phase must somehow be removed, whilst keeping the proteins fixed in place. If this could be done, however, the otolith could be ‘unlocked’ to reveal unprecedented insight into the life-history of fish.

ACKNOWLEDGEMENTS

Funding was provided by the Australian Research Council project LP140100087. ORBT was supported by an Australian Postgraduate Award. I thank K. Limburg, M. Wojtas, D. Green, and J.

Lueders-Dumont for their invaluable advice and comments.

(Anken et al., 2004, Coffin et al., 2002, Davis et al., 1997, Degens et al., 1969, Gower, 2008, Holubowicz et al., 2017, Ibsch et al., 2004)

REFERENCES

Andersen, O., De Rosa, M.C., Pirolli, D., Tooming-Klunderud, A., Petersen, P.E., Andre, C. (2011)

Polymorphism, selection and tandem duplication of transferrin genes in Atlantic cod

(Gadus morhua)--conserved synteny between fish monolobal and tetrapod bilobal

transferrin loci. BMC Genetics 12, 51. [In English].

81

Anken, R.H., Beier, M., Rahmann, H. (2004) Hypergravity decreases carbonic anhydrase-reactivity

in inner ear maculae of fish. Journal of Experimental Zoology Part a-Comparative

Experimental Biology 301A, 815-819.

Arias, J.L., Neira-Carrillo, A., Arias, J.I., et al. (2004) Sulfated polymers in biological mineralization: a

plausible source for bio-inspired engineering. Journal of Materials Chemistry 14, 2154-

2160.

Aristotle, Balme, D.M., Peck, A.L. (ca. 340BCE (transl. 1965)) History of . [electronic

resource], (Loeb Classical Library: 437-439, Vol., Cambridge, MA : Harvard University

Press, 2014.

Asano, M., Mugiya, Y. (1993) Biochemical and calcium-binding properties of water-soluble proteins

isolated from otoliths of the tilapia, Oreochromis niloticus. Comparative Biochemistry and

Physiology Part B: Comparative Biochemistry 104, 201-205.

Baba, K., Shimizu, M., Mugiya, Y., Yamada, J. (1991) Otolith matrix proteins of walleye pollock;

biochemical properties and immunohistochemical localization in the saccular tissue. In:

Mechanisms and Phylogeny of Mineralization in Biological Systems: Biomineralization ′90.

(Eds. S. Suga, H. Nakahara), Springer Japan, Tokyo, pp. 57-61.

Bajoghli, B., Ramialison, M., Aghaallaei, N., Czerny, T., Wittbrodt, J. (2009) Identification of

Starmaker-like in medaka as a putative target gene of Pax2 in the otic vesicle.

Developmental Dynamics 238, 2860-2866.

Beier, M., Anken, R. (2006) On the role of carbonic anhydrase in the early phase of fish otolith

mineralization. Advances in Space Research 38, 1119-1122.

82

Beier, M., Anken, R.H., Rahmann, H. (2002) Susceptibility to abnormal (kinetotic) swimming fish

correlates with inner ear carbonic anhydrase-reactivity. Neuroscience Letters 335, 17-20.

Benser, M.E., Issa, N.P., Hudspeth, A.J. (1993) Hair-bundle stiffness dominates the elastic

reactance to otolithic-membrane shear. Hearing Research 68, 243-252.

Blasiole, B., Canfield, V., Thisse, B., Thisse, C., Mohideen, M.A.P.K., Levenson, R. (2004) The role of

Na,K-ATPase in zebrafish inner ear development. FASEB Journal 18, A34-A34. [In English].

Blasiole, B., Canfield, V.A., Vollrath, M.A., et al. (2006) Separate Na,K-ATPase genes are required

for otolith formation and semicircular canal development in zebrafish. Developmental

Biology 294, 148-160.

Borbas, J.E., Wheeler, A.P., Sikes, C.S. (1991) Molluscan shell matrix phosphoproteins - correlation

of degree of phosphorylation to shell mineral microstructure and to invitro regulation of

mineralization. Journal of Experimental Zoology 258, 1-13. [In English].

Borelli, G., Guibbohni, M.E., Mayer-Gostan, N., et al. (2003a) Daily variations of endolymph

composition: relationship with the otolith calcification process in trout. Journal of

Experimental Biology 206, 2685-2692.

Borelli, G., Mayer-Gostan, N., De Pontual, H., Boeuf, G., Payan, P. (2001) Biochemical relationships

between endolymph and otolith matrix in the trout (Oncorhynchus mykiss) and turbot

(Psetta maxima). Calcified Tissue International 69, 356-364.

Borelli, G., Mayer-Gostan, N., Merle, P.L., et al. (2003b) Composition of biomineral organic

matrices with special emphasis on turbot (Psetta maxima) otolith and endolymph.

Calcified Tissue International 72, 717-725. [In English].

83

Brophy, D., Jeffries, T.E., Danilowicz, B.S. (2004) Elevated manganese concentrations at the cores

of clupeid otoliths: possible environmental, physiological, or structural origins. Marine

Biology 144, 779-786.

Brothers, E.B. (1978) Exogenous factors and formation of daily and sub-daily growth increments in

fish otoliths. American Zoologist 18, 631-631.

Campana, S.E. (1999) Chemistry and composition of fish otoliths: pathways, mechanisms and

applications. Marine Ecology Progress Series 188, 263-297.

Campana, S.E. (2001) Accuracy, precision and quality control in age determination, including a

review of the use and abuse of age validation methods. Journal of Fish Biology 59, 197-

242. [In English].

Campana, S.E., Fowler, A.J., Jones, C.M. (1994) Otolith elemental fingerprinting for stock

identification of Atlantic cod (Gadus morhua) using Laser-Ablation ICP-MS. Canadian

Journal of Fisheries and Aquatic Sciences 51, 1942-1950. [In English].

Campana, S.E., Neilson, J.D. (1985) Microstructure of fish otoliths. Canadian Journal of Fisheries

and Aquatic Sciences 42, 1014-1032. [In English].

Campana, S.E., Thorrold, S.R. (2001) Otoliths, increments, and elements: keys to a comprehensive

understanding of fish populations? Canadian Journal of Fisheries and Aquatic Sciences 58,

30-38.

Clendenon, S.G., Shah, B., Miller, C.A., et al. (2009) Cadherin-11 controls otolith assembly:

Evidence for extracellular cadherin activity. Developmental Dynamics 238, 1909-1922.

Coffin, A.B., Higgs, D.M., Presson, J.C., Popper, A.N. (2002) Distribution of unconventional myosins

in the zebrafish ear. Bioacoustics 12, 140-142.

84

Coffin, A.B., Raine, J.C., Hawryshyn, C.W. (2012) Exposure to thyroid hormone in ovo affects otolith

crystallization in rainbow trout Oncorhynchus mykiss. Environmental Biology of Fishes 95,

347-354.

Cohen-Salmon, M., El-Amraoui, A., Leibovici, M., Petit, C. (1997) Otogelin: A glycoprotein specific

to the acellular membranes of the inner ear. Proceedings of the National Academy of

Sciences of the United States of America 94, 14450-14455. [In English].

Cruz, S., Shiao, J.-C., Liao, B.-K., Huang, C.-J., Hwang, P.-P. (2009) Plasma membrane calcium

ATPase required for semicircular canal formation and otolith growth in the zebrafish inner

ear. Journal of Experimental Biology 212, 639-647.

Dannevig, E.H. (1956) Chemical composition of the zones in cod otoliths. Jour Conseil Perm

Internatl Explor Mer 21, 156-159.

Dauphin, Y., Dufour, E. (2003) Composition and properties of the soluble organic matrix of the

otolith of a marine fish: Gadus morhua Linne, 1758 (Teleostei, Gadidae). Comparative

Biochemistry and Physiology. Part A: Molecular and Integrative Physiology 134, 551-561.

Dauphin, Y., Dufour, E. (2008) Nanostructures of the aragonitic otolith of cod (Gadus morhua).

Micron 39, 891-896.

Davis, J.G., Burns, F.R., Navaratnam, D., et al. (1997) Identification of a structural constituent and

one possible site of postembryonic formation of a teleost otolithic membrane.

Proceedings of the National Academy of Sciences of the United States of America 94, 707-

712.

Davis, J.G., Oberholtzer, J.C., Burns, F.R., Greene, M.I. (1995a) Molecular cloning and

characterization of an inner ear-specific structural protein. Science 267, 1031-1034.

85

Davis, J.G., Oberholtzer, J.C., Burns, F.R., et al. (1995b) Use of the teleost saccule to identify genes

involved in inner ear function. DNA and Cell Biology 14, 833-839. [In English].

Deans, M.R., Peterson, J.M., Wong, G.W. (2010) Mammalian Otolin: A multimeric glycoprotein

specific to the inner ear that interacts with otoconial matrix protein Otoconin-90 and

Cerebellin-1. PloS One 5.

Degens, E.T., Deuser, W.G., Haedrich, R.L. (1969) Molecular structure and composition of fish

otoliths. Marine Biology 2, 105-113.

DeLaurier, A., Eames, B.F., Blanco-Sánchez, B., et al. (2010) Zebrafish sp7:EGFP: A transgenic for

studying otic vesicle formation, skeletogenesis, and bone regeneration. Genesis 48, 505-

511.

Devries, D.R., Frie, R.V. (1996) Determination of age and growth, (Fisheries techniques, Second

edition, Vol., American Fisheries Society {a}, Suite 110, 5410 Grosvenor Lane, Bethesda,

Maryland 20814-2199, USA.

Dorval, E., Piner, K., Robertson, L., Reiss, C.S., Javor, B., Vetter, R. (2011) Temperature record in the

oxygen stable isotopes of Pacific sardine otoliths: Experimental vs. wild stocks from the

Southern California Bight. Journal of Experimental Marine Biology and Ecology 397, 136-

143.

Doubleday, Z.A., Harris, H.H., Izzo, C., Gillanders, B.M. (2014) Strontium randomly substituting for

calcium in fish otolith aragonite. Analytical Chemistry 86, 865-869.

Drenckhahn, D., Schluter, K., Allen, D., Bennett, V. (1985) Colocalization of band 3 with ankyrin and

spectrin at the basal membrane of intercalated cells in the rat kidney. Science 230, 1287-

1289. [In English].

86

Dunkelberger, D.G., Dean, J.M., Watabe, N. (1980) The ultrastructure of the otolithic membrane

and otolith in the juvenile mummichog, Fundulus heteroclitus. Journal of Morphology 163,

367-377. [In English].

Dyson, H.J., Wright, P.E. (2002) Coupling of folding and binding for unstructured proteins. Current

Opinion in Structural Biology 12, 54-60.

Edeyer, A., de Pontual, H., Payan, P., Troadec, H., Severe, A., Mayer-Gostan, N. (2000) Daily

variations of the saccular endolymph and plasma compositions in the turbot Psetta

maxima: relationship with the diurnal rhythm in otolith formation. Marine Ecology

Progress Series 192, 287-294.

Elsdon, T.S., Wells, B.K., Campana, S.E., et al. (2008) Otolith chemistry to describe movements and

life-history parameters of fishes: Hypotheses, assumptions, limitations and inferences. In:

Oceanography and Marine Biology: An Annual Review, Vol 46. Vol. 46. Oceanography and

Marine Biology (Eds. R.N. Gibson, R.J.A. Atkinson, J.D.M. Gordon), pp. 297-+.

Enger, P.S. (1964) Ionic composition of the cranial and labyrinthine fluids and saccular D.C.

potentials in fish. Comparative Biochemistry and Physiology 11, 131-137. [In English].

Falini, G., Fermani, S., Vanzo, S., Miletic, M., Zaffino, G. (2005) Influence on the formation of

aragonite or vaterite by otolith macromolecules. European Journal of Inorganic Chemistry,

162-167.

Fange, R., Lidman, U., Larsson, A. (1972) Fluids and jellies of acusticolateralis system in relation to

body-fluids in Coryphaenoides rupestris and other fishes. Marine Biology 17, 180-&. [In

English].

Fink, A.L. (2005) Natively unfolded proteins. Current Opinion in Structural Biology 15, 35-41.

87

Funamoto, T., Mugiya, Y. (1998) Binding of strontium vs calcium to 17 beta-estradiol-induced

proteins in the plasma of the goldfish Carassius auratus. Fisheries Science 64, 325-328.

Gaboriaud, C., Frachet, P., Thielens, N.M., Arlaud, G.J. (2012) The human C1q globular domain:

structure and recognition of non-immune self ligands. Frontiers in Immunology 2, 8. [In

English].

Gaffney, J.P., Valentine, A.M. (2012) Beyond bilobal: transferrin homologs having unusual domain

architectures. Biochimica et Biophysica Acta 1820, 212-217. [In English].

Gauldie, R.W., Davies, N.M., Coote, G., Vickridge, I. (1990) The relationship between organic

material and check rings in fish otoliths. Comparative Biochemistry and Physiology a-

Physiology 97, 461-474. [In English].

Gibert, Y., Sassi-Messai, S., Fini, J.-B., et al. (2011) Bisphenol A induces otolith malformations

during vertebrate embryogenesis. BMC Developmental Biology 11.

Goodyear, R.J., Richardson, G.P. (2002) Extracellular matrices associated with the apical surfaces of

sensory epithelia in the inner ear: Molecular and structural diversity. Journal of

Neurobiology 53, 212-227.

Gower, L.B. (2008) Biomimetic Model Systems for Investigating the Amorphous Precursor Pathway

and Its Role in Biomineralization. Chemical Reviews 108, 4551-4627. [In English].

Guibbolini, M., Borelli, G., Mayer-Gostan, N., et al. (2006) Characterization and variations of

organic parameters in teleost fish endolymph during day-night cycle, starvation and stress

conditions. Comparative Biochemistry and Physiology a-Molecular & Integrative

Physiology 145, 99-107.

88

Hale, R., Swearer, S.E. (2008) Otolith microstructural and microchemical changes associated with

settlement in the diadromous fish Galaxias maculatus. Marine Ecology Progress Series

354, 229-234.

Haruyama, N., Sreenath, T.L., Suzuki, S., et al. (2009) Genetic evidence for key roles of decorin and

biglycan in dentin mineralization. Matrix Biology 28, 129-136.

Hicks, A.S., Closs, G.P., Swearer, S.E. (2010) Otolith microchemistry of two amphidromous galaxiids

across an experimental salinity gradient: A multi-element approach for tracking

diadromous migrations. Journal of Experimental Marine Biology and Ecology 394, 86-97.

Hoie, H., Otterlei, E., Folkvord, A. (2004) Temperature-dependent fractionation of stable oxygen

isotopes in otoliths of juvenile cod (Gadus morhua L.). ICES Journal of Marine Science 61,

243-251.

Holubowicz, R., Wojtas, M., Taube, M., Kozak, M., Ozyhar, A., Dobryszycki, P. (2017) Effect of

calcium ions on structure and stability of the C1q-like domain of Otolin-1 from human and

zebrafish. FEBS Journal 284, 4278-4297.

Hughes, I., Binkley, J., Hurle, B., et al. (2008) Identification of the Otopetrin Domain, a conserved

domain in vertebrate otopetrins and invertebrate otopetrin-like family members. BMC

Evolutionary Biology 8.

Hughes, I., Blasiole, L., Huss, D., et al. (2004) Otopetrin 1 is required for otolith formation in the

zebrafish Danio rerio. Developmental Biology 276, 391-402.

Hughes, I., Saito, M., Schlesinger, P.H., Ornitz, D.M. (2007) Otopetrin 1 activation by purinergic

nucleotides regulates intracellular calcium. Proceedings of the National Academy of

Sciences of the United States of America 104, 12023-12028. [In English].

89

Hurle, B., Ignatova, E., Massironi, S.M., et al. (2003) Non-syndromic vestibular disorder with

otoconial agenesis in tilted/mergulhador mice caused by mutations in Otopetrin 1. Human

Molecular Genetics 12, 777-789. [In English].

Hurle, B., Marques-Bonet, T., Antonacci, F., et al. (2011) Lineage-specific evolution of the

vertebrate Otopetrin gene family revealed by comparative genomic analyses. BMC

Evolutionary Biology 11, 18. [In English].

Ibsch, M., Anken, R., Beier, M., Rahmann, H. (2004) Endolymphatic calcium supply for fish otolith

growth takes place via the proximal portion of the otocyst. Cell and Tissue Research 317,

333-336.

Iozzo, R.V. (1998) Matrix proteoglycans: From molecular design to cellular function. Annual Review

of Biochemistry 67, 609-652. [In English].

Irie, T. (1960) The growth of the fish otolith. Jour Fac Fish and Animal Husbandry Hiroshima Univ 3,

203-229.

Izzo, C., Doubleday, Z.A., Gillanders, B.M. (2016a) Where do elements bind within the otoliths of

fish? Marine and Freshwater Research 67, 1072-1076.

Izzo, C., Doubleday, Z.A., Grammer, G.L., et al. (2016b) Fish as proxies of ecological and

environmental change. Reviews in Fish Biology and Fisheries 26, 265-286. [In English].

Izzo, C., Doubleday, Z.A., Schultz, A.G., Woodcock, S.H., Gillanders, B.M. (2015) Contribution of

water chemistry and fish condition to otolith chemistry: comparisons across salinity

environments. Journal of Fish Biology 86, 1680-1698.

Jolivet, A., Bardeau, J.-F., Fablet, R., Paulet, Y.-M., de Pontual, H. (2008) Understanding otolith

biomineralization processes: new insights into microscale spatial distribution of organic

90

and mineral fractions from Raman microspectrometry. Analytical and Bioanalytical

Chemistry 392, 551-560.

Jovine, L., Darie, C.C., Litscher, E.S., Wassarman, P.M. (2005) Zona Pellucida Domain proteins. In:

Annual Review of Biochemistry. Vol. 74. Annual Review of Biochemistry, Annual Reviews,

Palo Alto, pp. 83-114.

Kalish, J.M. (1989) Otolith microchemistry - validation of the effects of physiology, age and

environment on otolith composition. Journal of Experimental Marine Biology and Ecology

132, 151-178. [In English].

Kalish, J.M. (1991) Determinants of otolith chemistry - seasonal-variation in the composition of

blood-plasma, endolymph and otoliths of bearded rock cod Pseudophycis barbatus.

Marine Ecology Progress Series 74, 137-159. [In English].

Kang, Y.J., Stevenson, A.K., Yau, P.M., Kollmar, R. (2008) Sparc protein is required for normal

growth of zebrafish otoliths. J Assoc Res Otolaryngol 9, 436-451.

Kaplon, T.M., Michnik, A., Drzazga, Z., Richter, K., Kochman, M., Ozyhar, A. (2009) The rod-shaped

conformation of Starmaker. Biochimica Et Biophysica Acta-Proteins and Proteomics 1794,

1616-1624.

Kaplon, T.M., Rymarczyk, G., Nocula-Lugowska, M., et al. (2008) Starmaker exhibits properties of

an intrinsically disordered protein. Biomacromolecules 9, 2118-2125.

Kelm, R.J., Jr., Swords, N.A., Orfeo, T., Mann, K.G. (1994) Osteonectin in matrix remodeling. A

plasminogen-osteonectin-collagen complex. Journal of Biological Chemistry 269, 30147-

30153. [In English].

91

Khan, K.M., Drescher, D.G. (1990) Proteins of the gelatinous layer of the trout saccular otolithic

membrane. Hearing Research 43, 149-158.

Kim, E., Hyrc, K.L., Speck, J., et al. (2010) Regulation of cellular calcium in vestibular supporting

cells by Otopetrin 1. Journal of Neurophysiology 104, 3439-3450. [In English].

Kwan, A.P.L., Cummings, C.E., Chapman, J.A., Grant, M.E. (1991) Macromolecular organization of

chicken Type-X collagen invitro. Journal of Cell Biology 114, 597-604. [In English].

Lacruz, R.S., Smith, C.E., Bringas, P., et al. (2012) Identification of novel candidate genes involved in

mineralization of dental enamel by genome-wide transcript profiling. Journal of Cellular

Physiology 227, 2264-2275.

Lagerstromfermer, M., Nilsson, M., Backman, B., et al. (1995) Amelogenin signal peptide mutation

- correlation between mutations in the amelogenin gene (Amgx) and manifestations of x-

linked amelogenesis imperfecta. Genomics 26, 159-162. [In English].

Lambert, L.A. (2012) Molecular evolution of the transferrin family and associated receptors.

Biochimica Et Biophysica Acta-General Subjects 1820, 244-255.

Lee, T.W., Tsang, V.W.K., Birch, N.P. (2015) Physiological and pathological roles of tissue

plasminogen activator and its inhibitor neuroserpin in the nervous system. Frontiers in

Cellular Neuroscience 9, 9. [In English].

Limburg, K.E. (1995) Otolith strontium traces environmental history of subyearling american shad

Alosa sapidissima. Marine Ecology Progress Series 119, 25-35. [In English].

Limburg, K.E., Elfman, M. (2010) Patterns and magnitude of Zn:Ca in otoliths support the recent

phylogenetic typology of Salmoniformes and their sister groups. Canadian Journal of

Fisheries and Aquatic Sciences 67, 597-604.

92

Limburg, K.E., Wuenschel, M.J., Hüssy, K., Heimbrand, Y., Samson, M. (2018) Making the otolith

magnesium chemical calendar-clock tick: Plausible mechanism and empirical evidence.

Reviews in Fisheries Science & Aquaculture 26, 479-493. [In English].

Loewen, T.N., Carriere, B., Reist, J.D., Halden, N.M., Anderson, W.G. (2016) Linking physiology and

biomineralization processes to ecological inferences on the life history of fishes.

Comparative Biochemistry and Physiology a-Molecular & Integrative Physiology 202, 123-

140. [In English].

Lowenstein, O. (1936) The equilibrium function of the vertebrate labyrinth. Biological Reviews of

the Cambridge Philosophical Society 11, 113-145.

Lu, W., Zhou, D., Freeman, J.J., Thalmann, I., Ornitz, D.M., Thalmann, R. (2010) In vitro effects of

recombinant otoconin 90 upon calcite crystal growth. Significance of tertiary structure.

Hearing Research 268, 172-183.

Lundberg, Y.W., Zhao, X., Yamoah, E.N. (2006) Assembly of the otoconia complex to the macular

sensory epithelium of the vestibule. Brain Research 1091, 47-57.

Malicki, J., Schier, A.F., SolnicaKrezel, L., et al. (1996) Mutations affecting development of the

zebrafish ear. Development 123, 275-283. [In English].

Mann, S., Parker, S.B., Ross, M.D., Skarnulis, A.J., Williams, R.J.P. (1983) The ultrastructure of the

calcium carbonate balance organs of the inner-ear - an ultrahigh resolution electron-

microscopy study. Proceedings of the Royal Society Series B-Biological Sciences 218, 415-+.

[In English].

93

Mayer-Gostan, N., Kossmann, H., Watrin, A., Payan, P., Boeuf, G. (1997) Distribution of ionocytes

in the saccular epithelium of the inner ear of two teleosts (Oncorhynchus mykiss and

Scophthalmus maximus). Cell & Tissue Research 289, 53-61. [In English].

Melancon, S., Fryer, B.J., Markham, J.L. (2009) Chemical analysis of endolymph and the growing

otolith: fractionation of metals in freshwater fish species. Environmental Toxicology and

Chemistry 28, 1279-1287.

Miller, M.B., Clough, A.M., Batson, J.N., Vachet, R.W. (2006) Transition metal binding to cod otolith

proteins. Journal of Experimental Marine Biology and Ecology 329, 135-143.

Miyamoto, H. (1978) [Amino acids and electrolytes of the endolymph of the fish (Prussian carp)

and the transfer of kanamycin to the endolymph (author's transl)]. Nihon Jibiinkoka Gakkai

Kaiho 81, 358-371. [In Japanese].

Morales-Nin, B. (1986a) Chemical composition of the otoliths of the sea bass Dicentrarchus labrax

pisces serranidae. Cybium 10, 115-120.

Morales-Nin, B.Y.O. (1986b) Structure and composition of otoliths of cape hake Merluccius

capensis. South African Journal of Marine Science, 3-10.

Motta, C.M., Avallone, B., Balassone, G., et al. (2009) Morphological and biochemical analyses of

otoliths of the ice-fish Chionodraco hamatus confirm a common origin with red-blooded

species. Journal of Anatomy 214, 153-162.

Mugiya, Y. (1966) Calcification in fish and shell-fish-V. Nippon Suisan Gakkaishi 32, 117-129.

Mugiya, Y. (1968) Calcification in fish and shellfish: Histochemical similarities between the otolith

and the macula region of sacculus in juvenile rainbow trout with special reference to the

94

otolith formation of fish. Bulletin of the Japanese Society of Scientific Fisheries 34, 1096-

1106.

Mugiya, Y. (1974) Calcium-45 behavior at the level of the otolithic organs of rainbow trout. Nippon

Suisan Gakkaishi 40, 457-463.

Mugiya, Y. (1977) Effect of acetazolamide on the otolith growth of goldfish. Nippon Suisan

Gakkaishi 43, 1053-1058.

Mugiya, Y. (1984) Diurnal rhythm in otolith formation in the rainbow-trout, Salmo gairdneri -

seasonal reversal of the rhythm in relation to plasma calcium concentrations. Comparative

Biochemistry and Physiology a-Physiology 78, 289-293. [In English].

Mugiya, Y. (1986) Effects of calmodulin inhibitors and other metabolic modulators on invitro

otolith formation in the rainbow trout, Salmo gairdnerii. Comparative Biochemistry and

Physiology a-Physiology 84, 57-60. [In English].

Mugiya, Y. (1987) Phase difference between calcification and organic matrix formation in the

diurnal growth of otoliths in the rainbow-trout, Salmo gairdneri. Fishery Bulletin 85, 395-

401. [In English].

Mugiya, Y., Kawamura, H., Aratsu, S. (1979) Carbonic anhydrase and otolith formation in the

rainbow trout Salmo gairdneri enzyme activity of the sacculus and calcium uptake by the

otolith in vitro. Nippon Suisan Gakkaishi 45, 879-882.

Mugiya, Y., Takahashi, K. (1985) Chemical properties of the saccular endolymph in the rainbow

trout Salmo gairdneri. Bulletin of the Faculty of Fisheries Hokkaido University 36, 57-63. [In

English].

95

Mugiya, Y., Tanaka, S. (1995) Incorporation of water-borne strontium into otoliths and its turnover

in the goldfish Carassius auratus - effects of strontium concentrations, temperature, and

17-Beta-Estradiol. Fisheries Science 61, 29-35. [In English].

Mugiya, Y., Watabe, N., Yamada, J., Dean, J.M., Dunkelberger, D.G., Shimizu, M. (1981) Diurnal

rhythm in otolith formation in the goldfish, Carassius auratus. Comparative Biochemistry

and Physiology a-Physiology 68, 659-662. [In English].

Mugiya, Y., Yoshida, M. (1995) Effects of calcium antagonists and other metabolic modulators on

in vitro calcium deposition on otoliths in the rainbow trout Oncorhynchus mykiss. Fisheries

Science 61, 1026-1030. [In English].

Murayama, E., Herbomel, P., Kawakami, A., Takeda, H., Nagasawa, H. (2005) Otolith matrix

proteins OMP-1 and Otolin-1 are necessary for normal otolith growth and their correct

anchoring onto the sensory maculae. Mechanisms of Development 122, 791-803.

Murayama, E., Okuno, A., Ohira, T., Takagi, Y., Nagasawa, H. (2000) Molecular cloning and

expression of an otolith matrix protein cDNA from the rainbow trout, Oncorhynchus

mykiss. Comparative Biochemistry and Physiology B-Biochemistry & Molecular Biology 126,

511-520.

Murayama, E., Takagi, Y., Nagasawa, H. (2004) Immunohistochemical localization of two otolith

matrix proteins in the otolith and inner ear of the rainbow trout, Oncorhynchus mykiss:

comparative aspects between the adult inner ear and embryonic otocysts. Histochemistry

and Cell Biology 121, 155-166.

96

Murayama, E., Takagi, Y., Ohira, T., Davis, J.G., Greene, M.I., Nagasawa, H. (2002) Fish otolith

contains a unique structural protein, Otolin-1. European Journal of Biochemistry 269, 688-

696.

Nemoto, Y., Chatani, M., Inohaya, K., Hiraki, Y., Kudo, A. (2008) Expression of marker genes during

otolith development in medaka. Gene Expression Patterns 8, 92-95.

Nicolson, T. (2004) Control of crystal growth in biology: A molecular biological approach using

zebrafish. Crystal Growth & Design 4, 667-669.

Ohyama, Y., Lin, J.H., Govitvattana, N., et al. (2016) FAM20A binds to and regulates FAM20C

localization. Scientific Reports 6.

Otake, T., Ishii, T., Ishii, T., Nakahara, M., Nakamura, R. (1997) Changes in otolith strontium:

Calcium ratios in metamorphosing Conger myriaster leptocephali. Marine Biology 128,

565-572.

Oxman, D.S., Barnett-Johnson, R., Smith, M.E., et al. (2007) The effect of vaterite deposition on

sound reception, otolith morphology, and inner ear sensory epithelia in hatchery-reared

Chinook salmon (Oncorhynchus tshawytscha). Canadian Journal of Fisheries and Aquatic

Sciences 64, 1469-1478.

Page-McCaw, A., Ewald, A.J., Werb, Z. (2007) Matrix metalloproteinases and the regulation of

tissue remodelling. Nature Reviews: Molecular Cell Biology 8, 221-233. [In English].

Pannella, G. (1971) Fish otoliths: daily growth layers and periodical patterns. Science (New York,

N.Y.) 173, 1124-1127.

Parmentier, E., Cloots, R., Warin, R., Henrist, C. (2007) Otolith crystals (in Carapidae): Growth and

habit. Journal of Structural Biology 159, 462-473.

97

Payan, P., Borelli, G., Boeuf, G., Mayer-Gostan, N. (1998) Relationship between otolith and somatic

growth: consequence of starvation on acid-base balance in plasma and endolymph in the

rainbow trout Oncorhynchus mykiss. Fish Physiology and Biochemistry 19, 35-41.

Payan, P., Borelli, G., Priouzeau, F., De Pontual, H., Boeuf, G., Mayer-Gostan, N. (2002) Otolith

growth in trout Oncorhynchus mykiss: supply of Ca2+ and Sr2+ to the saccular endolymph.

Journal of Experimental Biology 205, 2687-2695.

Payan, P., De Pontual, H., Boeuf, G., Mayer-Gostan, N. (2004) Endolymph chemistry and otolith

growth in fish. Comptes Rendus Palevol 3, 535-547.

Payan, P., Edeyer, A., De Pontual, H., Borelli, G., Boeuf, G., Mayer-Gostan, N. (1999) Chemical

composition of saccular endolymph and otolith in fish inner ear: lack of spatial uniformity.

American Journal of Physiology-Regulatory Integrative and Comparative Physiology 277,

R123-R131.

Payan, P., Kossmann, H., Watrin, A., MayerGostan, N., Boeuf, G. (1997) Ionic composition of

endolymph in teleosts: Origin and importance of endolymph alkalinity. Journal of

Experimental Biology 200, 1905-1912.

Petko, J.A., Millimaki, B.B., Canfield, V.A., Riley, B.B., Levenson, R. (2008) Otoc1: A novel Otoconin-

90 ortholog required for otolith mineralization in zebrafish. Developmental Neurobiology

68, 209-222.

Pisam, M., Jammet, C., Laurent, D. (2002) First steps of otolith formation of the zebrafish: role of

glycogen? Cell and Tissue Research 310, 163-168.

98

Pisam, M., Payan, P., LeMoal, C., Edeyer, A., Boeuf, G., Mayer-Gostan, N. (1998) Ultrastructural

study of the saccular epithelium of the inner ear of two teleosts, Oncorhynchus mykiss and

Psetta maxima. Cell and Tissue Research 294, 261-270.

Popper, A.N., Fay, R.R., Platt, C., Sand, O. (2003) Sound detection mechanisms and capabilities of

teleost fishes, (Sensory processing in aquatic environments, Vol.

Popper, A.N., Platt, C., Saidel, W.M. (1982) Acoustic functions in the fish ear. Trends in

Neurosciences 5, 276-280. [In English].

Popper, A.N., Ramcharitar, J., Campana, S.E. (2005) Why otoliths? Insights from inner ear

physiology and fisheries biology. Marine and Freshwater Research 56, 497-504.

Pote, K.G., Hauer, C.R., Michel, H., Shabanowitz, J., Hunt, D.F., Kretsinger, R.H. (1993) Otoconin-22,

the major protein of aragonitic frog otoconia, is a homolog of Phospholipase-A2.

Biochemistry 32, 5017-5024. [In English].

Pote, K.G., Ross, M.D. (1991) Each otoconia polymorph has a protein unique to that polymorph.

Comparative Biochemistry and Physiology B-Biochemistry & Molecular Biology 98, 287-

295. [In English].

Poznar, M., Dudek, J., Dobryszycki, P. (2013) Expression and purification of the intrinsically

disordered Otolith Matrix Macromolecule-64. FEBS Journal 280, 121-121.

Poznar, M., Hołubowicz, R., Wojtas, M., et al. (2017) Structural properties of the intrinsically

disordered, multiple calcium ion-binding Otolith Matrix Macromolecule-64 (OMM-64).

Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics 1865, 1358-1371.

99

Radtke, R.L. (1985) Recruitment parameters resolved from structural and chemical components of

juvenile Dascyllus albisella otoliths. Proceedings of the 6th International Coral Reef

Congress 5, 397-401.

Randow, F., Seed, B. (2001) Endoplasmic reticulum chaperone GP96 is required for innate

immunity but not cell viability. Nature Cell Biology 3, 891.

Reading, B., Sullivan, C., Schilling, J. (2017) Vitellogenesis in Fishes, Vol.

Reibisch, J. (1899) Über die eizahl bei Pleuronectes platessa und die altersbestimmung dieser form

aus den otolithen. Wissenschaftliche Meeresuntersuchungen 4, 232-248.

Reimer, T., Dempster, T., Wargelius, A., et al. (2017) Rapid growth causes abnormal vaterite

formation in farmed fish otoliths. Journal of Experimental Biology 220, 2965-2969. [In

English].

Reimer, T., Dempster, T., Warren-Myers, F., Jensen, A.J., Swearer, S.E. (2016) High prevalence of

vaterite in sagittal otoliths causes hearing impairment in farmed fish. Scientific Reports 6,

25249. [In English].

Renn, J., Winkler, C. (2014) Osterix/Sp7 regulates biomineralization of otoliths and bone in medaka

(Oryzias latipes). Matrix Biology 34, 193-204.

Riley, B.B., Zhu, C.W., Janetopoulos, C., Aufderheide, K.J. (1997) A critical period of ear

development controlled by distinct populations of ciliated cells in the zebrafish.

Developmental Biology 191, 191-201. [In English].

Romanek, C.S., Gauldie, R.W. (1996) A predictive model of otolith growth in fish based on the

chemistry of the endolymph. Comparative Biochemistry and Physiology a-Physiology 114,

71-79.

100

Rozycka, M., Wojtas, M., Jakob, M., et al. (2014) Intrinsically disordered and pliable Starmaker-like

protein from medaka (Oryzias latipes) controls the formation of calcium carbonate

crystals. PloS One 9, e114308.

Saitoh, S. (1993) Studies on the anatomy and microstructure of the inner ear of tilapia

(Oreochromis niloticus), with special reference to the otolith growth in the otolithic organ.

Scientific Reports of Hokkaido Fisheries Experimental Station 0, 1-87.

Sasagawa, T., Mugiya, Y. (1996) Biochemical properties of water-soluble otolith proteins and the

immunobiochemical detection of the proteins in serum and various tissues in the tilapia

Oreochromis niloticus. Fisheries Science 62, 970-976.

Schreiber, A.M., Wang, X., Tan, Y., et al. (2010) Thyroid hormone mediates otolith growth and

development during flatfish metamorphosis. General and Comparative Endocrinology 169,

130-137.

Schulz-Mirbach, T., Ladich, F., Plath, M., Hess, M. (2018) Enigmatic ear stones: what we know

about the functional role and evolution of fish otoliths. Biological Reviews of the

Cambridge Philosophical Society 0.

Shiao, J.-C., Wu, S.-M., Hwang, Y.-P., Wu, D.-P., Hwang, P.-P. (2008) Evaluation of thyroid-mediated

otolith growth of larval and juvenile tilapia. Journal of Experimental Biology 211, 1919-

1926.

Shiao, J.C., Hwang, P.P. (2004) Thyroid hormones are necessary for teleostean otolith growth.

Marine Ecology Progress Series 278, 271-278.

101

Shiao, J.C., Lin, L.Y., Horng, J.L., Hwang, P.P., Kaneko, T. (2005) How can teleostean inner ear hair

cells maintain the proper association with the accreting otolith? Journal of Comparative

Neurology 488, 331-341.

Shima, J.S., Swearer, S.E. (2009) Larval quality is shaped by matrix effects: implications for

connectivity in a marine metapopulation. Ecology 90, 1255-1267.

Silverman, G.A., Bird, P.I., Carrell, R.W., et al. (2001) The serpins are an expanding superfamily of

structurally similar but functionally diverse proteins - Evolution, mechanism of inhibition,

novel functions, and a revised nomenclature. Journal of Biological Chemistry 276, 33293-

33296. [In English].

Smith, C.E., Nanci, A., Moffatt, P. (2006) Evidence by signal peptide trap technology for the

expression of carbonic anhydrase 6 in rat incisor enamel organs. European Journal of Oral

Sciences 114, 147-153.

Sokolowski, B.H. (1986) Development of the otolith in embryonic fishes with special reference to

the toadfish, Opsanus tau. Scanning Electron Microscopy, 1635-1648.

Söllner, C., Burghammer, M., Busch-Nentwich, E., et al. (2003) Control of crystal size and lattice

formation by Starmaker in otolith biomineralization. Science 302, 282-286.

Söllner, C., Schwarz, H., Geisler, R., Nicolson, T. (2004) Mutated Otopetrin 1 affects the genesis of

otoliths and the localization of Starmaker in zebrafish. Development Genes and Evolution

214, 582-590.

Somerville, R.P.T., Oblander, S.A., Apte, S.S. (2003) Matrix metalloproteinases: old dogs with new

tricks. Genome Biology 4, 11. [In English].

102

Stooke-Vaughan, G.A., Huang, P., Hammond, K.L., Schier, A.F., Whitfield, T.T. (2012) The role of

hair cells, cilia and ciliary motility in otolith formation in the zebrafish otic vesicle.

Development 139, 1777-1787.

Stooke-Vaughan, G.A., Obholzer, N.D., Baxendale, S., Megason, S.G., Whitfield, T.T. (2015) Otolith

tethering in the zebrafish otic vesicle requires Otogelin and Alpha-Tectorin. Development

142, 1137-1145.

Sturrock, A.M., Hunter, E., Milton, J.A., et al. (2015) Quantifying physiological influences on otolith

microchemistry. Methods in Ecology and Evolution 6, 806-816. [In English].

Sturrock, A.M., Trueman, C.N., Darnaude, A.M., Hunter, E. (2012) Can otolith elemental chemistry

retrospectively track migrations in fully marine fishes? Journal of Fish Biology 81, 766-795.

Sturrock, A.M., Trueman, C.N., Milton, J.A., Waring, C.P., Cooper, M.J., Hunter, E. (2014)

Physiological influences can outweigh environmental signals in otolith microchemistry

research. Marine Ecology Progress Series 500, 245-264.

Sumanas, S., Larson, J.D., Bever, M.M. (2003) Zebrafish chaperone protein GP96 is required for

otolith formation during ear development. Developmental Biology 261, 443-455.

Sutmuller, M., Bruijn, J.A., deHeer, E. (1997) Collagen types VIII and X, two non-fibrillar, short-

chain collagens. Structure homologies, functions and involvement in pathology. Histology

and Histopathology 12, 557-566. [In English].

Swearer, S.E., Shima, J.S. (2010) Regional variation in larval retention and dispersal drives

recruitment patterns in a temperate reef fish. Marine Ecology Progress Series 417, 229-

236.

103

Tagliabracci, V.S., Engel, J.L., Wen, J.Z., et al. (2012) Secreted kinase phosphorylates extracellular

proteins that regulate biomineralization. Science 336, 1150-1153. [In English].

Takagi, Y. (1997) Meshwork arrangement of mitochondria-rich, Na+,K+-ATPase-rich cells in the

saccular epithelium of rainbow trout (Oncorhynchus mykiss) inner ear. Anatomical Record

248, 483-489.

Takagi, Y. (2000) Ultrastructural immunolocalization of the otolith water-soluble-matrix in the

inner ear of rainbow trout just-hatched fry. Fisheries Science 66, 71-77.

Takagi, Y. (2002) Otolith formation and endolymph chemistry: a strong correlation between the

aragonite saturation state and pH in the endolymph of the trout otolith organ. Marine

Ecology Progress Series 231, 237-245.

Takagi, Y., Ishida, K., Mugiya, Y. (2000) Carbohydrates of the otolith organ in the rainbow trout

Oncorhynchus mykiss detected by lectins. Fisheries Science 66, 933-939.

Takagi, Y., Takahashi, A. (1999) Characterization of otolith soluble-matrix producing cells in the

saccular epithelium of rainbow trout (Oncorhynchus mykiss) inner ear. Anatomical Record

254, 322-329.

Takagi, Y., Tohse, H., Murayama, E., Ohira, T., Nagasawa, H. (2005) Diel changes in endolymph

aragonite saturation rate and mRNA expression of otolith matrix proteins in the trout

otolith organ. Marine Ecology Progress Series 294, 249-256.

Tanimoto, M., Ota, Y., Inoue, M., Oda, Y. (2011) Origin of inner ear hair cells: Morphological and

functional differentiation from ciliary cells into hair cells in zebrafish inner ear. Journal of

Neuroscience 31, 3784-3794.

104

Thalmann, I., Hughes, I., Tong, B.D., Ornitz, D.M., Thalmann, R. (2006) Microscale analysis of

proteins in inner ear tissues and fluids with emphasis on endolymphatic sac, otoconia, and

organ of Corti. Electrophoresis 27, 1598-1608.

Thalmann, R., Thalmann, I., Lu, W. (2011) Significance of tertiary conformation of otoconial matrix

proteins - clinical implications. Acta Oto-Laryngologica 131, 382-390.

Thomas, O.R.B., Ganio, K., Roberts, B.R., Swearer, S.E. (2017) Trace element-protein interactions in

endolymph from the inner ear of fish: implications for environmental reconstructions

using fish otolith chemistry. Metallomics 9, 239-249.

Thomas, O.R.B., Swearer, S.E., Kapp, E.A., et al. (2019) The inner ear proteome of fish. FEBS Journal

286, 66-81.

Tohse, H., Ando, H., Mugiya, Y. (2004) Biochemical properties and immunohistochemical

localization of carbonic anhydrase in the sacculus of the inner ear in the salmon

Oncorhynchus masou. Comparative Biochemistry and Physiology a-Molecular & Integrative

Physiology 137, 87-94.

Tohse, H., Mugiya, Y. (2001) Effects of enzyme and anion transport inhibitors on in vitro

incorporation of inorganic carbon and calcium into endolymph and otoliths in salmon

Oncorhynchus masou. Comparative Biochemistry and Physiology a-Molecular and

Integrative Physiology 128, 177-184.

Tohse, H., Mugiya, Y. (2004) Effects of acidity and a metabolic inhibitor on incorporation of calcium

and inorganic carbon into endolymph and otoliths in salmon Oncorhynchus masou.

Fisheries Science 70, 595-600.

105

Tohse, H., Murayama, E., Ohira, T., Takagi, Y., Nagasawa, H. (2006) Localization and diurnal

variations of carbonic anhydrase rnRNA expression in the inner ear of the rainbow trout

Oncorhynchus mykiss. Comparative Biochemistry and Physiology B-Biochemistry &

Molecular Biology 145, 257-264.

Tohse, H., Saruwatari, K., Kogure, T., Nagasawa, H., Takagi, Y. (2009) Control of polymorphism and

morphology of calcium carbonate crystals by a matrix protein aggregate in fish otoliths.

Crystal Growth & Design 9, 4897-4901.

Tohse, H., Takagi, Y., Nagasawa, H. (2008) Identification of a novel matrix protein contained in a

protein aggregate associated with collagen in fish otoliths. FEBS Journal 275, 2512-2523.

Tomas, J., Geffen, A.J. (2003) Morphometry and composition of aragonite and vaterite otoliths of

deformed laboratory reared juvenile herring from two populations. Journal of Fish Biology

63, 1383-1401.

Uversky, V.N., Dunker, A.K. (2010) Understanding protein non-folding. Biochimica Et Biophysica

Acta-Proteins and Proteomics 1804, 1231-1264.

Visse, R., Nagase, H. (2003) Matrix metalloproteinases and tissue inhibitors of metalloproteinases -

Structure, function, and biochemistry. Circulation Research 92, 827-839. [In English].

Wang, A.H., Liang, Y., Fridell, R.A., et al. (1998a) Association of unconventional myosin MYO15

mutations with human nonsyndromic deafness DFNB3. Science 280, 1447-1451.

Wang, H.J., Tannukit, S., Zhu, D.H., Snead, M.L., Paine, M.L. (2005) Enamel matrix protein

interactions. Journal of Bone and Mineral Research 20, 1032-1040. [In English].

106

Wang, X., Tan, Y., Schreiber, A.M. (2006) A dramatic influence of thyroid hormone on larval

southern flounder otolith development. Integrative and Comparative Biology 46, E263-

E263.

Wang, X., Tan, Y., Sievers, Q., et al. (2011) Thyroid hormone-responsive genes mediate otolith

growth and development during flatfish metamorphosis. Comparative Biochemistry and

Physiology a-Molecular & Integrative Physiology 158, 163-168.

Wang, Y.X., Kowalski, P.E., Thalmann, I., Ornitz, D.M., Mager, D.L., Thalmann, R. (1998b) Otoconin-

90, the mammalian otoconial matrix protein, contains two domains of homology to

secretory Phospholipase A2. Proceedings of the National Academy of Sciences of the

United States of America 95, 15345-15350.

Warren-Myers, F., Dempster, T., Fjelldal, P.G., Hansen, T., Jensen, A.J., Swearer, S.E. (2014) Stable

isotope marking of otoliths during vaccination: a novel method for mass-marking fish.

Aquaculture Environment Interactions 5, 143-154.

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2015a) Expression of SPARC and the Osteopontin-like

protein during skeletal development in the cichlid fish Oreochromis mossambicus.

Developmental Dynamics 244, 955-972. [In English].

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2015b) Spatial expression of Otolith Matrix Protein-1

and Otolin-1 in normally and kinetotically swimming fish. Anatomical Record-Advances in

Integrative Anatomy and Evolutionary Biology 298, 1765-1773. [In English].

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2016) Not all inner ears are the same: Otolith matrix

proteins in the inner ear of sub-adult cichlid fish, Oreochromis mossambicus, reveal

107

insights into the biomineralization process. Anatomical Record-Advances in Integrative

Anatomy and Evolutionary Biology 299, 234-245.

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2017) Formation of the inner ear during embryonic

and larval development of the cichlid fish (Oreochromis mossambicus). Connective Tissue

Research 58, 172-195.

Weiner, S., Addadi, L. (1997) Design strategies in mineralized biological materials. Journal of

Materials Chemistry 7, 689-702. [In English].

Wheeler, A.P., George, J.W., Evans, C.A. (1981) Control of calcium carbonate nucleation and crystal

growth by soluble matrix of oyster shell. Science 212, 1397-1398. [In English].

Whitfield, T.T., Granato, M., vanEeden, F.J.M., et al. (1996) Mutations affecting development of

the zebrafish inner ear and lateral line. Development 123, 241-254. [In English].

Whitfield, T.T., Riley, B.B., Chiang, M.Y., Phillips, B. (2002) Development of the zebrafish inner ear.

Developmental Dynamics 223, 427-458. [In English].

Wojtas, M., Holubowicz, R., Poznar, M., Maciejewska, M., Ozyhar, A., Dobryszycki, P. (2015)

Calcium ion binding properties and the effect of phosphorylation on the intrinsically

disordered Starmaker protein. Biochemistry 54, 6525-6534.

Wojtas, M., Wolcyrz, M., Ozyhar, A., Dobryszycki, P. (2012) Phosphorylation of intrinsically

disordered Starmaker protein increases its ability to control the formation of calcium

carbonate crystals. Crystal Growth & Design 12, 158-168.

Wright, P.J. (1991) Calcium binding by soluble matrix of the otoliths of Atlantic salmon Salmo salar

Journal of Fish Biology 38, 625-628.

108

Xiao, J.Y., Tagliabracci, V.S., Wen, J.Z., Kim, S.A., Dixon, J.E. (2013) Crystal structure of the Golgi

Casein Kinase. Proceedings of the National Academy of Sciences of the United States of

America 110, 10574-10579.

Xiao, S.X., Yu, C., Chou, X.M., et al. (2001) Dentinogenesis imperfecta 1 with or without progressive

hearing loss is associated with distinct mutations in DSPP. Nature Genetics 27, 201-204. [In

English].

Xu, Y., Zhang, H., Yang, H., Zhao, X., Lovas, S., Lundberg, Y.W. (2010) Expression, functional, and

structural analysis of proteins critical for otoconia development. Developmental Dynamics

239, 2659-2673.

Yang, C.-H., Cheng, C.-H., Chen, G.-D., et al. (2011) Zona Pellucida Domain-containing protein Beta-

Tectorin is crucial for zebrafish proper inner ear development. PloS One 6.

Yariz, K.O., Duman, D., Zazo Seco, C., et al. (2012) Mutations in OTOGL, encoding the inner ear

protein Otogelin-like, cause moderate sensorineural hearing loss. American Journal of

Human Genetics 91, 872-882.

Ye, Y.P., Hu, W.H., Guo, F.J., Zhang, W.K., Wang, J., Chen, A.M. (2012) Glycosaminoglycan chains of

Biglycan promote Bone Morphogenetic Protein-4-induced osteoblast differentiation.

International Journal of Molecular Medicine 30, 1075-1080. [In English].

Yin, K., Paine, M.L. (2017) Bicarbonate transport during enamel maturation. Calcified Tissue

International 101, 457-464.

Yu, X., Lau, D., Ng, C.P., Roy, S. (2011) Cilia-driven fluid flow as an epigenetic cue for otolith

biomineralization on sensory hair cells of the inner ear. Development 138, 487-494.

109

Zhang, X.A., Bontrager, A.L., Hemler, M.E. (2001) Transmembrane-4 superfamily proteins associate

with activated Protein Kinase C (PKC) and link PHC to specific beta(1) integrins. Journal of

Biological Chemistry 276, 25005-25013. [In English].

Zhang, Z. (1992) Ultrastructure of otolith increments and checks in the teleost fish Oreochromis

niloticus. Journal of Morphology 211, 213-220. [In English].

Zhang, Z., Runham, N.W. (1992) Temporal deposition of incremental and discontinuous zones in

the otoliths of Oreochromis niloticus (L.). Journal of Fish Biology 40, 333-339.

Zhao, X., Yang, H., Yamoah, E.N., Lundberg, Y.W. (2007) Gene targeting reveals the role of Oc90 as

the essential organizer of the otoconial organic matrix. Developmental Biology 304, 508-

524. [In English].

Zvackova, I., Matalova, E., Lesot, H. (2017) Regulators of collagen fibrillogenesis during molar

development in the mouse. Frontiers in Physiology 8, 14. [In English].

110

CHAPTER THREE: TRACE ELEMENT–PROTEIN INTERACTIONS IN ENDOLYMPH FROM THE INNER EAR OF FISH: IMPLICATIONS FOR ENVIRONMENTAL RECONSTRUCTIONS USING FISH OTOLITH CHEMISTRY

(Appendix 3.1; Originally published in Metallomics, 2017, 9, 239)

ABSTRACT

Otoliths, the biomineralised hearing “earstones” from the inner ear of fishes, grow throughout the lifespan of an individual, with deposition of alternating calciferous and proteinaceous bands occurring daily. Trace element:calcium ratios within daily increments measured by laser ablation- inductively coupled plasma-mass spectrometry (LA-ICP-MS) are often used in fisheries science to reconstruct environmental histories. There is, however, considerable uncertainty as to which elements are interacting with either the proteinaceous or calciferous zones of the otolith, and thus their utility as indicators of environmental change. To answer this, we used size exclusion chromatography-inductively coupled plasma-mass spectrometry (SEC-ICP-MS) of endolymph, the otolith growth medium, to determine the binding interactions for a range of elements. In addition, we used solution ICP-MS to quantify element concentrations in paired otolith and endolymph samples and determined relative enrichment factors for each. We found 12 elements that are only present in the proteinaceous fraction, 6 that are only present in the salt fraction, and 4 that are present in both. These findings have important implications for the reconstruction of environmental histories based on changes in otolith elemental composition: (1) elements occurring only in the salt fraction are most likely to reflect changes in the physico-chemical environment experienced during life; (2) elements occurring only in the proteinaceous fraction are more likely to reflect physiological rather than environmental events; and (3) elements occurring in both the salt and proteinaceous fractions are likely to be informative of both endogenous and exogenous processes, potentially reducing their utility in environmental reconstructions.

111

INTRODUCTION

Otoliths, or ear stones, are calciferous and proteinaceous structures found in the inner ear of all bony fishes, which aid in the detection of changes in ambient water pressure due to gravity or sound (Popper and Fay, 2011). Growing continually throughout an individual’s life, otoliths are chiefly calcium carbonate (90-97% by mass) with the remainder being proteins, glycoproteins, lipoproteins, glycosaminoglycans and polysaccharides (Campana, 1999, Dauphin and Dufour,

2003). The fish inner ear contains three pairs of otoliths, termed sagittae, lapilli and asteriscii, with sagittal otoliths being the most widely studied (Campana, 1999). The three otolith types differ in size (sagittae often being the largest), location in the ear labyrinth, and calcium carbonate polymorph. Sagittae and lapilli tend to be aragonite (although calcite has been observed in select species, such as Oreochromis mossambicus(Weigele et al., 2016)), whereas asteriscii are usually composed of vaterite (Campana, 1999). In this study, the term “otolith” solely refers to sagittal otoliths.

Otoliths form in embryo as primordia that are subsequently calcified within the saccula (Lundberg et al., 2015). These sac-like sections of the inner ear labyrinth are filled with endolymph, a fluid rich in both structural materials (such as calcium and hydrogen carbonate ions), as well as the proteins and other macromolecules that are believed to facilitate growth (Asano and Mugiya,

1993a, Payan et al., 1999, Payan et al., 2004). From embryogenesis on, the otoliths continue to grow, with aragonite being precipitated onto the organic matrix daily, producing a double-banded increment composed of an aragonite-rich area and a protein-rich area, respectively termed the incremental zone and the discontinuous zone (Campana and Neilson, 1985, Zhang, 1992).

Otoliths are an important tool in fish and fisheries research, as they are largely metabolically inert, recording the specifics of the physicochemical environment experienced by a fish at any given

112 point in its life (Campana, 1999, Campana and Thorrold, 2001). Elemental incorporation in the otolith has been shown to be influenced by pH, salinity, temperature and concentration gradients of trace elements (Mugiya and Tanaka, 1995, Elsdon et al., 2008, Sturrock et al., 2012, Izzo et al.,

2015). Consequently, otoliths are increasingly being used for life-history reconstructions of past environments by determining otolith increment trace element-to-calcium ratios via laser ablation- inductively coupled plasma-mass spectrometry (LA-ICP-MS).

A suite of elements are routinely used in otolith microchemical analyses. A few elements are used as proxies for change in the physico-chemical properties of an individual’s environment (e.g. Sr and Ba for salinity and temperature (Elsdon and Gillanders, 2002), Mn for hypoxia (Limburg et al.,

2011)), whilst the entire suite of elements can be used as a natural tag for stock discrimination.

Whether these elements are present due to incorporation resulting from environmental change

(either by substituting for calcium in the inorganic phase, being trapped in the increment or adhering to organic macromolecules) or because they occur ‘naturally’ as biochemical co-factors, being physiologically regulated by the organism (Mugiya et al., 1991, Campana, 1999, Doubleday et al., 2014), is largely unknown. To date, only two studies have examined the interaction of trace elements with the proteinaceous portions of otoliths. In both cases, however, the methods employed could only indicate whether a given element was present or absent in the organic fraction of the otolith and thus did not reveal whether it was bound to any macromolecules, or existed as a free ion (Miller et al., 2006, Izzo et al., 2016). Currently, the methods used to prepare otoliths for microchemical analysis are not consistent. Thorrold and Swearer found that the majority of microchemical studies simply rinsed the otolith in ultrapure water, with few studies thoroughly removing organic material prior to analysis (Thorrold and Swearer, 2009). As a consequence, the presence of proteinaceous material during LA-ICP-MS analyses has great potential to blur and confound otolith elemental data, potentially reducing confidence in using

113 these data to reconstruct environmental histories. For example, ontogenic variation in otolith chemistry has been observed in some species (e.g. Zn in Esox lucius(Friedrich and Halden, 2010)).

It has been suggested that age related variation in element incorporation may be due to changes in aragonite-to-protein ratios within and between annuli (Thorrold and Swearer, 2009).

Although the great majority of otolith research has focussed on the ecological and fisheries management applications of otolith microstructure and microchemistry, there have been some important studies on the structural anatomy and biochemistry of otoliths and endolymph. These studies have provided insights into the nature of the proteinaceous matrix of otoliths (Degens et al., 1969, King, 1978, Söllner et al., 2003, Miller et al., 2006, Tohse et al., 2008), the composition of endolymph (Mugiya and Takahashi, 1985, Kalish, 1991, Payan et al., 1999, Borelli et al., 2001,

Tohse and Mugiya, 2001), and the factors necessary for biogenic crystallisation of aragonite

(Takagi, 2002, Jolivet et al., 2008, Melancon et al., 2008).

To date, only a small number of fish otolith proteins have been identified and characterised.

Amongst these are: seeding proteins such as the otoconin-90 ortholog, Otoc1 (Petko et al., 2008) and several matrix proteins, Otolith Matrix Protein-1 (OMP-1) (Murayama et al., 2000a), Otolin-1

(Murayama et al., 2004), Otolith Matrix Molecule-64 (OMM-64) (Tohse et al., 2008), and

Starmaker(Söllner et al., 2003). In addition, a host of glycoproteins (such as sparc) (Kang et al.,

2008), otopetrins (Hughes et al., 2004, Wang et al., 2011) and other assisting macromolecules that are believed to be involved in tethering, scaffolding and growth of otoliths have been identified, both in the otolith and the endolymph. The emerging picture of the endolymph is one of a “factory floor” for otolith development. Although general information about endolymph composition is available for a variety of species (Kalish, 1991, Mugiya and Yoshida, 1995, Romanek and Gauldie,

1996, Payan et al., 1999, Borelli et al., 2001, Takagi, 2002, Melancon et al., 2008), specific

114 information on the biochemical composition of otoliths is only known for a handful of species, such as Danio rerio and Oncorhynchus mykiss (Murayama et al., 2000b, Tohse et al., 2008).

There are considerable difficulties associated with determining metal partitioning with regards to the organic component of an otolith. Obtaining the organic material is simple enough, with several methods available that involve either the dissolution of the inorganic matrix (e.g. dissolution in

Ethylenediaminetetraacetic acid (EDTA) (Asano and Mugiya, 1993b), or other acids (Davies et al.,

2011)) or the denaturing of the organic phase (e.g. by using nonspecific protein cutters such as

Proteinase-K (Shiao et al., 1999), or inorganic reagents such as peroxides or sodium hypochlorite

(Thorrold and Swearer, 2009)). Unfortunately, all of these methods result in the potential decoupling of trace elements from the proteinaceous portion of the otolith.

Due to the difficulty in successfully extracting otolith protein without disrupting its structure, we set out to determine the trace element-protein interactions in the endolymph of adult

Acanthopagrus butcheri. A. butcheri is an estuarine fish native to southern Australia and an excellent candidate for this study given its history of use as a model species in numerous previous otolith microchemistry studies (Elsdon and Gillanders, 2002, Elsdon and Gillanders, 2003, Elsdon and Gillanders, 2004, de Vries et al., 2005). As otoliths mineralise within the endolymph, we reasoned that it would contain all the necessary organic precursors for otolith growth. Therefore, any trace element-protein interactions occurring within the endolymph might reasonably be expected to also occur within the otolith.

To this end, we employed native separation of endolymph, by using size exclusion chromatography linked to inductively coupled plasma mass spectrometry (SEC-ICP-MS) and assessed samples for a suite of elements. SEC is a low resolution separation technique, but suitable for native separation of proteinaceous and salt fractions. In SEC-ICP-MS, a biological fluid

115 is separated into different molecular mass fractions, which are then subsequently determined for the presence of trace elements. Through this approach, we were able to determine the presence of trace element-protein interactions as specific molecular mass fractions eluted from the SEC column, indicating which elements are present as a result of playing a role as a co-factor, and thus physiologically regulated, and those that are solely present as free ions in endolymph, and therefore potential candidates for either substitution for calcium or through random trapping in the aragonite lattice. In addition to this, we used solution ICP-MS to perform bulk metal analysis on both otoliths as well as their associated endolymph, providing an indication of relative levels of enrichment of each of the trace elements. The results from these two sets of analyses were then used to evaluate which elements are most likely to reflect changes in environmental conditions experienced by fish during their lifetime.

In this study, we use the term “proteinaceous fraction” to refer to elution volumes that are less than the total volume (Vt) of the column (4 mL) where elements are bound to high molecular weight species (such as proteins), and the term “salt fraction” to refer to elution volumes equivalent to the Vt where elements are unbound, present as free ions in solution.

METHODS

Animals

Adult Acanthopagrus butcheri (n=4) were caught, via gill net, in the Werribee Estuary, Victoria,

Australia (37°57'45.6"S, 144°40'02.9"E) in August 2015. Fish were measured (29-32 cm fork length) and then euthanised. Immediately following death, fish were decapitated. A transverse incision was made, from the snout above the eyes to above the gill arches, and the cranium cut open. The brain was removed, taking care not to disrupt the sacculum during dissection and thus avoid

116 contamination of the endolymph with blood. A needle and syringe was inserted into both the left and right saccula and used to carefully draw out the endolymph, which was placed in labelled cryotubes and stored on dry ice. After extraction of endolymph, both the left and right sagittal otoliths were gently removed with forceps, placed into individually labelled cryotubes and stored on dry ice. Otolith and endolymph samples were transferred to a -80 ⁰C freezer for storage until analysis. Fish were collected under Arthur Rylah Institute Animal Ethics Committee approval 14/12 and the State of Victoria, Australia Department of Environment and Primary Industries’ Fisheries research permit #1135.

Chemicals and Standards

All standards, samples and buffers were prepared using ultra-pure Milli-Q H2O (18.2 MΩ; Merck

Millipore, Australia). Superoxide dismutase-1 (SOD-1, copper, zinc, Mw = 32 kDa (dimer); Sigma

Aldrich, Castle Hill, Australia) and ferritin (iron, Mw = 440 kDa; Sigma Aldrich, Castle Hill, Australia) were employed as protein standards for quantification of metals. The buffer, 0.2 M pH 7.7 ammonium nitrate (Sigma Aldrich, Castle Hill, Australia), was prepared daily. Size exclusion chromatography (SEC) columns were calibrated using ferritin, ceruloplasmin (Mw = 132 kDa), SOD-

1, conalbumin (Mw = 75 kDa) and Vitamin B12 (Mw = 1.36 kDa), following the method described previously by Lothian and Roberts (Lothian and Roberts, 2016).

For bulk metal analyses, nitric acid (HNO3 65% v/v, analytical grade Merck Millipore, Australia) was used to acidify samples and standards, and diluted with either Milli-Q water or a 3 mM calcium diluent to a final concentration of 1% v/v. Otolith samples were diluted to an approximate concentration of 3 mM Ca2+ (assuming 97% w/w calcium carbonate) and analysed using multi- element certified reference standards ICP-MS-CAL2-1, ICP-MS-CAL3-1, ICP-MS-CAL4-1

117

(AccuStandard, USA) at levels of 0, 0.1, 0.5, 1.0, 5.0, 10.0, 50.0, 100.0, 500.0 and 1000.0 μg L-1, dissolved in an acidified calcium diluent. The calcium diluent was made from a stock calcium

-1 2+ 2+ standard (1000 mg L Ca , Agilent Technologies, Australia) diluted to 3 mM Ca and 1% v/v HNO3.

This was done to ensure matrix-matching of otoliths with standards. Endolymph samples were analysed with standards of the same levels, made from the same multi- element certified reference standards, dissolved in 1% v/v acid diluent. A reference standard, yttrium (200 µg L-1,

ICP-MS-Internal Standard Solution-1, AccuStandard), was used to normalise all bulk metal analyses, and introduced to the ICP-MS via a T-piece positioned after the peristaltic pump.

Sample Preparation- size exclusion chromatography-inductively coupled plasma mass spectrometry

Prior to SEC-ICP-MS, extracted endolymph samples were thawed for 10 minutes on wet ice. They were then centrifuged at 16000 G for 5 minutes. Estimation of protein concentration was undertaken by diluting samples to 1 in 10 in Tris-buffered saline, loading 2 μL on a NanoDrop spectrophotometer (ThermoFisher Scientific, Victoria, Australia) and reading UV-absorbance at

280nm. A volume of undiluted endolymph was then loaded on the SEC-ICP-MS at volumes (ca. 5-

15 µL) such that approximately 100 µg of protein was present in each sample. The remaining endolymph was refrozen and retained for subsequent bulk metal analysis.

We used an Agilent Technologies 1200 liquid chromatography system (Mulgrave, Australia). The column used was an Agilent Technologies Bio-SEC 3 (3 μm particle size; 150 Å pore structure; 4.6 mm i.d.) using a 200 mM ammonium nitrate buffer of pH 7.7–7.8. As fractions eluted, they were introduced into a MicroMist nebuliser (Glass Expansion, Australia) fitted to a Agilent Technologies

7700x ICP-MS. Eluents were carried to the nebuliser via polyethyl ether ketone (PEEK) tubing (o.d.

118

1.58 mm, i.d. 0.13 mm, ca. 50 cm). Helium was used as the collision gas to minimise polyatomic interferences for all elements except lithium, boron, magnesium, phosphorous and chromium, for which no collision gas was used. We employed cesium and antimony (133Cs, 121Sb; 10 µg L-1 each) as internal standards in our buffer. The following elements were analysed: 7Li, 11B, 24Mg, 31P, 39K,

44Ca, 51V, 52Cr, 55Mn, 56Fe, 59Co, 60Ni, 63Cu, 66Zn, 75As, 85Rb, 88Sr, 111Cd, 118Sn, 127I, 137Ba, 201Hg and

208Pb. The ICP-MS was tuned daily through solution nebulisation of a 1 µg L-1 lithium, cobalt, yttrium, cerium and thallium standard in 2% v/v HNO3 (Agilent Technologies, Australia). A more detailed method of both SEC-ICP-MS and solution ICP-MS bulk metal analysis is presented in

Lothian and Roberts(Lothian and Roberts, 2016). A summary of the ICP-MS operating parameters employed for all analyses is provided in Table 3.1.

Table 3.1. Typical operating parameters of SEC-ICP-MS Agilent 1200 LC Mobile phase 0.2M NH4NO3, pH 7.7-7.8 Column type Agilent Bio-SEC 3 Particle size 3 μm Pore size 150 Å Flow rate 0.4 mL/min Injection volume 5-15 μL (~100 μg total protein) Column dimensions 4.6 mm (i.d.) x 300 mm

Agilent 7700x ICP-MS RF Power 1550W Sample depth 8.0 mm Carrier gas 0.95 L/min Makeup gas 0.20 L/min Spray chamber temperature 7 ⁰C Extracts 1, 2 -12, -200 V Omega bias, lens -95, 8.3 V Deflect, plate bias 2.2, -60 V Cell entrance, exit -38, -68 V Octopole bias, RF -18.0, 200 V Collision gas He, 3.4 mL/min (for all except Li, Mg, P and Cr)

Sample preparation- bulk metal analysis

Extracted otoliths were transferred to 1.5 mL microcentrifuge tubes (TechnoPlas, Australia), and immersed in 500 µL of a cleaning solution composed of 15% v/v ultra-pure H2O2 (VLSI Selectipur,

119

BASF) buffered with 0.1 M NaOH (Sigma Aldrich, Castle Hill, Australia). After immersion, tubes were ultra-sonicated for 5 min (Sonic Clean 250HT), and then left for 24 h in the cleaning solution.

The cleaning solution was then aspirated off, and samples underwent three rinses in Milli-Q water.

Each rinse was accompanied by 5 min of ultra-sonication and 30 min resting time. Finally, otoliths were air-dried in a laminar flow bench for 24 h. Otoliths were then weighed, digested for 24 h in

2+ 10% v/v HNO3 and then diluted to give a final concentration of 3 mM Ca and 1% v/v HNO3.

Thawed endolymph samples were reconstituted in 0.1 M PBS and diluted to give sufficient volume for analysis (20-40x dilution). 2 mL aliquots of dissolved otoliths, endolymph and standards were transferred to polypropylene tubes, and introduced into the ICP-MS via an integrated automation system (IAS) autosampler (Agilent Technologies, Australia) employing a peristaltic pump. The tubing used was 400 mm polypropylene tubing (0.15 mm i.d.) and Tygon PeriPump tubing (0.25 mm i.d.). Time taken for sample uptake was 52 s, with a stabilisation time of 40 s. The integration time for each isotope monitored was 0.1 s. We assessed for the presence of calcium and 22 trace elements, listed in Table 3.2, below.

Comparison of endolymph and otolith samples

We determined the relative percentage enrichment of each element in otoliths as follows:

[푒푙푒푚푒푛푡 푐표푛푐푒푛푡푟푎푡𝑖표푛 𝑖푛 표푡표푙𝑖푡ℎ] 푃푒푟푐푒푛푡푎𝑔푒 푒푛푟𝑖푐ℎ푚푒푛푡 = 100 × [푒푙푒푚푒푛푡 푐표푛푐푒푛푡푟푎푡𝑖표푛 𝑖푛 푒푛푑표푙푦푚푝ℎ]

120

These enrichment factors should only be considered qualitative estimates to allow for comparison among elements as they assume that the elemental concentrations in the endolymph are representative of the average concentration throughout the life of the fish.

RESULTS AND DISCUSSION

Analytical performance

Through SEC-ICP-MS, we produced elution profiles for 24 elements present in 8 endolymph samples (n=4, samples from the left and right saccula of each fish). These were averaged, and showed little variation in either intensity or presence of peaks among individuals (Appendix 3.2).

Broadly, when considering the various elution profiles, a peak in the salt fraction indicates that the element in question may be influenced by exogenous processes and a potential candidate for use in environmental reconstructions, whereas a peak in the proteinaceous fraction indicates that the element is bound to a protein species and is likely to be physiologically regulated.

Due to a limited volume of endolymph material, we were only able to perform bulk metal analyses on paired otolith and endolymph samples from 3 fish, assessing for calcium and 22 trace elements.

In these samples, all trace element concentrations other than calcium were found to be present in considerably higher concentrations in the endolymph than in the otolith. Limits of bulk analyses were determined using 3σ for the limit of detection (LOD) and 10σ for the limit of quantification

(LOQ), with only 1 element (boron) falling beneath the LOQ for endolymph, and 9 elements falling beneath the LOQ or LOD for otoliths (Table 3.2).

121

Table 3.2. Average concentration, standard deviation, range, coefficient of variation, and percentage enrichment of 22 trace elements from paired otolith and endolymph samples from adult Acanthopagrus butcheri (n=3), determined by inductively coupled plasma-mass spectrometry. All concentration values are expressed as µg of trace element per g of calcium.

Element Average concentration, Otolith coefficient of Otolith concentration Average concentration, Endolymph coefficient of Endolymph concentration % Enrichment Factor otolith (µg g-1), SD variation min-max endolymph variation min-max (100x[otolith]/ -1 -1 -1 (µg g ) (µg g ), SD (µg g ) [endolymph]) 7Li 8.5 (1.1) 13.4 7.2-10.3 261 (241) 92.4 108.1-539.6 3.25 11Ba 87.8 (41.3) 47.0 35.1-160.8 1954 (735) 37.6 1182.9-2646.4 4.49 24Mg 86.0 (9.5) 11.1 76.1-101.8 348564 (110199) 31.6 256831.2-470802.2 0.02 31P 294.1 (49.3) 16.8 216.4-357.3 2750186 (487376) 17.7 2370700.4-3299830.4 0.01 39K 2148.9 (281.7) 13.1 1804-2537.8 37529243 (13387597) 35.7 22400266.7-47843877.7 0.01 51V 0.1 (0.0) 43.9 BDL-0.1 101 (39) 38.6 64.2-141.4 0.08 52Cr 1.7 (0.2) 14.3 1.4-2.1 309 (99) 32.2 211.3-410.1 0.55 55Mn 3.3 (2.1) 63.7 1.8-6.5 73 (46) 63.2 32.5-123.1 4.51 56Fe 10.7 (1.9) 17.3 8.9-13.8 32584 (13257) 40.7 17319.9-41216.6 0.03 59Co 0.2 (0.0) 14.7 0.2-0.3 330 (20) 6.0 310.3-350.3 0.06 60Nib 1.9 (0.8) 40.8 0.6-2.7 288 (68) 23.5 212.7-343.7 0.66 63Cub 1.2 (0.7) 63.0 0.2-2.1 1357 (542) 39.9 880.1-1946.7 0.09 66Znb 11.6 (7.2) 61.8 BDL-20.2 52625 (8378) 15.9 43957.1-60679.6 0.02 75Asc 0.2 (0.2) 86.0 BDL-0.4 1247 (263) 21.1 1070.7-1549.2 0.02 85Rbb 0.8 (0.2) 27.6 0.5-1.1 2274 (958) 42.1 1420.1-3310.5 0.04 88Sr 10644 (458) 4.3 9990.3-11126.9 7754 (621) 8.0 7037.6-8135.3 137.27 111Cdd 0.1 (0.1) 128.8 BDL-0.2 9.9 (1.5) 15.1 8.3-11.2 0.57 118Snd 0.4 (0.4) 102.9 BDL-1.1 92 (74) 79.7 19.8-166.9 0.45 121Sbd 0.1 (0.1) 48.7 0.1-0.2 5.2 (3.3) 63.8 2.6-8.9 2.31 133Csd 0.2 (0.0) 12.9 0.2-0.2 29 (16) 53.6 16-46.6 0.62 137Ba 17.5 (12.8) 73.2 8.1-42.6 77 (44) 56.9 43.6-127.1 22.59 208Pb 0.8 (0.5) 55.1 0.4-1.6 35 (17) 50.3 16.1-50.5 2.41 aEndolymph concentration falls below limit of quantification, guideline only. bOtolith concentration value falls below limit of quantification, guideline only. cOtolith only assessed. dOtolith concentration value falls below limit of detection, guideline only.

122

In ecological applications of otolith trace element data (e.g. in life-history reconstructions), it is usual to present values with respect to their ratio to the total calcium concentration of the otolith

(i.e. µg of element per g of Ca). Consequently, after adjusting all values for dilution, we then converted all concentrations to µg g-1 to reflect their ratio to calcium in either otolith or endolymph. In endolymph, total calcium concentration ranged from 89-161 mg L-1, whereas otolith concentrations ranged from 1081-1210 mg L-1.

The individual results of both SEC-ICP-MS and bulk analysis are presented below, organised in the following categories; Group 1 elements, Group 2 elements, Transition metals, Elements of environmental concern, and Non-metals.

Group 1 elements

Lithium, potassium and rubidium all showed peaks at elution volumes greater than Vt, suggesting that these elements were only present in the salt fraction of the endolymph (Fig. 3.1a). Lithium, potassium and rubidium showed high concentrations in the endolymph accompanied by relatively low otolith concentrations (Table 3.2). An absence of a peak in the <4 mL region indicates that these elements are not interacting with any high molecular weight molecules (such as proteins).

Furthermore, as they are present solely in the salt fraction, and are univalent in terms of charge, they are highly unlikely to be candidates for substitution of Ca2+ in the aragonite lattice of the otolith. As a result, it is reasonable to suppose that the presence of these metals in otoliths is due to them being randomly trapped in interstitial spaces in the aragonite crystal structure during daily increment formation. This is further supported by their low percentage enrichment scores (Table

3.2). Lithium, having the highest enrichment (3.25%) is likely to be the most frequently incorporated into the otolith due to it having the smallest ionic radius of the Group 1 metals, and

123 therefore the most likely to fit in the interstitial spaces of the crystal lattice. Potassium, being present in extremely high concentrations in the endolymph (37.5 g g-1 Ca), has a relatively low enrichment (0.01%), suggesting a purely physiological role, particularly as it is a common biological analyte.

Our results suggest that the presence of K+ in the otolith is due to random trapping, an inevitable consequence of its high endolymph concentration. In the past, one or more of the alkali metals have been utilised in stock discrimination or life history reconstructions (Edmonds et al., 1992,

Hoff and Fuiman, 1993, Gillanders, 2002, Hicks et al., 2010, Miyan et al., 2016). As changes in otolith growth are likely to affect the probability of random trapping, caution should be exercised when making environmental predictions based on Group 1 metals.

Group 2 elements

Barium, strontium and calcium all showed a peak at 2.30 mL, a volume corresponding to the proteinaceous fraction (ca. 105 kDa, Fig. 3.1b) as well as a strong peak in the volume corresponding to the salt fraction. Magnesium was only present in the salt fraction.

Superimposition of traces suggests that calcium, barium and strontium may all be interacting with the same protein complexes (Fig. 3.1c). Interestingly, an examination of the elution profile for lead

(Fig. 3.1f), although not a Group 2 element, is interacting with similar macromolecules, as it shows a peak at the same volume. Considering the similar ionic radii of these elements and their identical charge (+2), it is reasonable to suppose that they are competing for binding sites within any macromolecules present. Of the known otolith proteins, Sparc (34kDa) in O. mykiss and Starmaker

(66 kDa) in D. rerio and Starmaker-like protein (39kDa) in Oryzias latipes have been found to have a calcium binding role. A similar protein would therefore be a candidate for binding elements

124 similar to calcium in both size and charge (Kang et al., 2008, Söllner et al., 2003, Rozycka et al.,

2014).

Strontium and barium both show high levels of enrichment in the otolith (137.27% and 22.59% respectively). Doubleday et al. used extended X-ray absorption fine structure spectroscopy to assay otoliths from a variety of species from a range of environments, and found that strontium directly substitutes for calcium within the aragonite lattice (Doubleday et al., 2014). Aragonite is orthorhombic in terms of crystal structure, with each calcium ion being surrounded by nine oxygen atoms (from carbonate ions). Strontium, barium and lead similarly form orthorhombic crystals (strontianite, witherite and cerussite, respectively (Antao and Hassan, 2009)). Aragonite preferentially incorporates species of similar, but larger, ionic radii than calcium at the growing edge of the otolith (Melancon et al., 2005). As all three metals are slightly larger than calcium (in the order Sr

Magnesium, on the other hand, has a smaller ionic radius than calcium and, from our results, is present solely in the salt fraction of the endolymph. As a consequence, it is unlikely to be preferentially replacing calcium within the growing crystal lattice. We speculate that magnesium’s incorporation within an otolith is due to random trapping during increment formation.

Furthermore, despite it having a comparable ionic radius to lithium (Shannon, 1976), otoliths show weak magnesium enrichment (0.02%), suggesting that its utility as an environmental marker is limited. Our results, however, confirm the use of both barium and strontium in life-history reconstructions derived from otolith microchemistry.

125

a) b)

c) d)

e) f)

g)

Figure 3.1. Size exclusion ICP-MS chromatograms of endolymph. (a) 7Li, 39K and 85Rb, (b) 24Mg, 44Ca, 88Sr and 137Ba, (c) 44Ca, 88Sr and 137Ba, (d) 56Fe, 59Co, 60Ni, 63Cu and 66Zn, (e) 51V, 52Cr and 201Hg, (f) 75As, 111Cd, 118Sn and 208Pb, (g) 31P, 60Ni and 63Cu chromatograms produced via SEC-ICP-MS from endolymph of adult Acanthopagrus butcheri (n=4). Markers indicate elution volumes for Ferritin (MW = 440 kDa), Ceruloplasmin (MW = 132 kDa), Conalbumin (MW = 75 kDa), Superoxide dismutase-1 (MW = 32 kDa dimer) and total volume (Vt).

126

Transition metals

As iron, zinc, cobalt, copper and nickel are excellent candidates for bioinorganic co-factors to a variety of vertebrate metalloproteins(Szpunar, 2005), it is unsurprising that these elements showed peaks in fractions corresponding to volumes of less than 4 mL (Fig. 3.1d). Copper presented 4 distinct peaks at 2.00, 2.25, 2.35 and 2.90 mL (ca. 235, 120, 90 and 20 kDa; Fig. 3.1d,

Appendix 3.2), suggesting its role as a cofactor in at least four or more distinct macromolecules.

Iron showed 2 peaks at 2.00 and 2.50 mL (ca. 235, 60 kDa), whereas nickel, cobalt and zinc each presented a single peak at 2.25, 2.40 and 2.40 mL, respectively (ca. 120, 80, 80 kDa).

Superimposition of all five traces reveals possible multi-elemental interactions. We suggest that copper and iron are interacting with the same large molecular weight macromolecule. Similarly, other volume pairs (i.e. copper and nickel, cobalt and zinc) may also potentially be interacting with the same macromolecule. However, further metalloproteomic investigation is required to determine if this is the case. In Oncorhynchus mykiss, for example, the identified 40 kDa protein

Otolith Matrix Protein 1, has been found to be a member of the transferrin family, and consequently has a Fe3+ binding domain (Murayama et al., 2000a, Murayama et al., 2005). We speculate that a similar protein to this, in A. butcheri, could potentially be that corresponding to iron’s second peak at 2.50 mL.

Considering that these elements show no peaks at volumes corresponding to the salt fraction, it is highly unlikely that they are incorporating in the otolith through substitution or random trapping to any great degree. Organic material has been found to account for 3-10% of the mass of an otolith (Campana, 1999, Dauphin and Dufour, 2003), this is reflected in the low percentage enrichment of the transition metals. We suggest that the presence of copper, zinc, iron, cobalt and nickel, in the otolith is primarily through their physiological roles as biomolecule co-factors, rather than as a result of the ambient environment, and would therefore caution against their

127 employment as environmental markers. The physiological role of some of these metals has begun to be elucidated within fish. For example, the timing of spawning in Pleuronectes platessa correlates with otolith zinc/calcium ratios in females (Sturrock et al., 2015). In light of this, we suggest that the analysis of otoliths for these metals may shed light on the timing of this, and other, life-history events.

Vanadium, mercury and chromium all produced peaks at approximately 2.40 mL (ca. 80 kDa), suggesting that they may be interacting with the same high molecular weight macromolecule as cobalt and zinc. Chromium also produced a second peak at 3.5 mL (ca. 5 kDa). This suggests that chromium is interacting with at least two distinct biomolecules. Both vanadium and chromium were found to have both low concentrations in endolymph and otolith (Table 3.2) as well as correspondingly low percentage enrichment factors of 0.08% and 0.55%. This indicates that the incorporation of these trace elements into the otolith is largely due to small interactions with biomolecules.

Manganese produced a single peak in the salt fraction (Appendix 3.2), with a relatively high percentage enrichment factor (4.51%, Table 3.2). These results suggest that the mechanism of manganese incorporation is substitution for calcium, a reasonable assumption given its ability to form a 2+ ion, and the relative similarity of its ionic radius (Shannon, 1976). There is evidence, however, of manganese occurring in high concentrations within the core of the otolith, and as such, is often used as an indication that the core has been reached during LA-ICP-MS (Brophy et al., 2004). A possible explanation for this is that manganese may play a role in embryo, facilitating the assembly of otolith primordia. Considering that our samples came from adult A. butcheri, it is reasonable to suppose that any developmental stage-dependent macromolecules would not be present.

128

Elements of environmental interest

Some of the elements assessed are considered toxicants to either humans or aquatic life (Florea and Busselberg, 2006). Tin, interestingly, showed a peak in its profile at 2.35 mL (ca. 90 kDa; Fig.

3.1f). As it is not traditionally a biomolecule cofactor, we propose that this peak may correspond to the presence of a large molecular weight organotin complex. This is a possible concern, considering tin’s toxicity to aquatic organisms (Rouleau et al., 2003, Burgin and Hardiman, 2011).

Further investigation is required to understand the nature of the tin species present in endolymph.

Cadmium showed peaks in the region corresponding to high molecular weight, with peaks at 2.32 and 3.05 mL (ca. 100 and 15 kDa). Neither cadmium nor tin had particularly high percentage enrichment factors (0.57% and 0.45%) or endolymph concentrations (Table 3.2). The relatively low concentrations and associated enrichment of these two elements suggests that they are poor candidates as environmental markers, except perhaps in heavily polluted waters.

The element profile for arsenic revealed a peak in both the high molecular weight region (2.18 mL, ca. 150 kDa) and the salt fraction. As arsenic is an element of potential toxicological concern, data relating to its concentration in the Werribee Estuary is readily available. It ranges from 0.5 to 2.0

µg L-1 in this system (Melbourne Water, 2014). Although this range is well below concentrations which were detected in endolymph samples (Table 3.2), arsenic is known to bio-accumulate

(Florea and Busselberg, 2006). However, arsenic had a relatively low percentage enrichment

(0.02%, Table 3.2). This, coupled with its relatively high endolymph concentration (Table 3.2), indicates that although it is unlikely to be a reasonable indicator of environmental change over time as recorded by the otolith, it may be a potential avenue of investigation with regards to its bioaccumulation effects.

129

As mentioned above, lead produced a single peak in the proteinaceous fraction of the endolymph

(Fig. 3.1f), eluting at a similar time as strontium, barium and calcium (i.e. 2.30 mL, ca. 105 kDa).

Lead has been used in life history reconstructions (Campana, 1999), as the similarity of its ionic radius to calcium’s makes it an excellent candidate for potential substitutions (Shannon, 1976).

However, considering its absence from the salt fraction, and its moderate percentage enrichment

(2.41%, Table 3.2), its viability as an environmental marker is uncertain. Lead may either be biochemically regulated and acting as an essential element, acting antagonistically by competing with calcium for protein binding-sites, or interacting with a protein, away from the active site.

Whether lead signatures provide useful environmental data will depend on how lead is incorporating into the otolith.

Non-metals

The elution profile for boron did not show any distinct peaks (Appendix 3.2). However, solution

ICP-MS of boron returned reasonably high endolymph concentrations, suggesting that boron concentrations used to create the elution profile may have been below the limit of detection of the SEC-ICP-MS. Iodine was not assessed for total concentration in either endolymph or otolith.

The elution profile for iodine showed a single, strong peak in the salt fraction (Appendix 3.2).

Phosphorous, however, produced two peaks in its elution profile, at 2.20 mL (possibly attaching to the same biomolecule as copper and nickel, ca. 140 kDa; Fig. 3.1g) and at 3.50 mL (ca. 5 kDa).

These results suggest either the presence of phosphorylation of endolymph macromolecules, and/or the presence of a high molecular weight inorganic phosphate. In a study conducted by

Söllner et al, the Starmaker protein was predicted, using NetPhos, as having up to 20% of its residues phosphorylated (Blom et al., 1999, Söllner et al., 2003). The high endolymph phosphorous

130 concentration, accompanied by a relatively low percentage enrichment factor (0.01%) suggests that phosphorous is physiologically regulated. Phosphorous content in the otolith may be due to its presence in a larger macromolecule, or through small deposits of calcium phosphate.

CONCLUSION

The present work reports the presence or absence of biomolecule interactions for a range of elements in fish endolymph as determined by SEC-ICP-MS, in conjunction with total metal quantification through solution ICP-MS. Consequently, we are able to make certain recommendations regarding the use of a variety of minor and trace elements for life-history reconstructions based on LA-ICP-MS of otoliths. Broadly, we have three main recommendations:

Firstly, for those elements that occur solely within the salt fraction, we believe that they are appropriate markers of ambient environmental change, but that caution must be exercised, particularly when using Group 1 or other singly-charged elements. These elements are unlikely to directly substitute for calcium in the aragonite lattice, and are likely to be incorporated through random trapping, the degree of which will be influenced by both thermokinetics and the ionic radius of the trapped species, smaller atoms being more easily trapped than those of large size.

Secondly, elements that occur only in the proteinaceous fractions of endolymph are clearly adhering to biomolecules, either by acting as a co-factor, or through intermolecular interactions.

Due to the biochemically regulated nature of the former role, it is wise to discount such elements from analyses, as an elevated level on a given increment may reveal physiological rather than environmental events. This does, however, suggest that a fruitful line of research is to determine what aspects of physiology, development or other endogenous factors influence incorporation of these elements. Finally, for those elements (e.g., Sr and Ba) that show peaks in both the salt and

131 proteinaceous fractions, this is an indication that both exogenous and endogenous processes may be contributing to the presence of these elements in the otolith, potentially reducing confidence in their utility in environmental reconstructions. Although thorough cleaning of otoliths may not resolve this issue, an investigation into the efficacy of cleaning to remove protein-bound elements is a potential avenue of future research.

The work described here is the first step in evaluating the mechanisms of elemental incorporation into otoliths. At this stage, we cannot be certain of the various element-macromolecule interactions until more is known about the tertiary structure of otolith macromolecules. To this end, a rewarding line of research would be to use SEC in conjunction with tandem mass spectrometry to perform proteomics on both otolith and endolymph from A. butcheri. In addition to this, the inclusion of additional native separation techniques (such as ion exchange) will result in finer resolution of proteins. A thorough proteomic analysis would allow identification of those proteins present in both endolymph and otoliths (Lothian et al., 2013) and their tertiary structures, and thus the trace element interactions occurring with them.

ACKNOWLEDGEMENTS

We thank G. Jenkins, D. Chamberlain, and L. Barrett for their assistance in the capture of adult fish.

We are deeply grateful to D. Chamberlain for photography and image development, and to I.

Volitakis for assistance with chemical analyses. All equipment images are © Agilent Technologies,

Inc. <26/8/16> Reproduced with Permission, Courtesy of Agilent Technologies, Inc. Funding was provided by the Australian Research Council project LP140100087 and the Australian Postgraduate

Award. The present study was conducted under Arthur Rylah Institute Animal Ethics Committee approval 14/12 and the State of Victoria Fisheries research permit #1135.

132

REFERENCES

Antao, S.M., Hassan, I. (2009) The orthorhombic structure of CaCO3, SrCO3, PbCO3 and BaCO3:

Linear structural trends. Canadian Mineralogist 47, 1245-1255. [In English].

Asano, M., Mugiya, Y. (1993a) Biochemical and Calcium-Binding Properties of Water-Soluble

Proteins Isolated from Otoliths of the Tilapia, Orechromis-Niloticus. Comparative

Biochemistry and Physiology B-Biochemistry & Molecular Biology 104, 201-205. [In

English].

Asano, M., Mugiya, Y. (1993b) Biochemical and calcium-binding properties of water-soluble

proteins isolated from otoliths of the tilapia, Oreochromis niloticus. Comparative

Biochemistry and Physiology Part B: Comparative Biochemistry 104, 201-205.

Blom, N., Gammeltoft, S., Brunak, S. (1999) Sequence and structure-based prediction of eukaryotic

protein phosphorylation sites. Journal of Molecular Biology 294, 1351-1362. [In English].

Borelli, G., Mayer-Gostan, N., De Pontual, H., Boeuf, G., Payan, P. (2001) Biochemical relationships

between endolymph and otolith matrix in the trout (Oncorhynchus mykiss) and turbot

(Psetta maxima). Calcified Tissue International 69, 356-364.

Brophy, D., Jeffries, T.E., Danilowicz, B.S. (2004) Elevated manganese concentrations at the cores

of clupeid otoliths: possible environmental, physiological, or structural origins. Marine

Biology 144, 779-786.

Burgin, S., Hardiman, N. (2011) The direct physical, chemical and biotic impacts on Australian

coastal waters due to recreational boating. Biodiversity and Conservation 20, 683-701. [In

English].

133

Campana, S.E. (1999) Chemistry and composition of fish otoliths: pathways, mechanisms and

applications. Marine Ecology Progress Series 188, 263-297.

Campana, S.E., Neilson, J.D. (1985) Microstructure of fish otoliths. Canadian Journal of Fisheries

and Aquatic Sciences 42, 1014-1032. [In English].

Campana, S.E., Thorrold, S.R. (2001) Otoliths, increments, and elements: keys to a comprehensive

understanding of fish populations? Canadian Journal of Fisheries and Aquatic Sciences 58,

30-38.

Dauphin, Y., Dufour, E. (2003) Composition and properties of the soluble organic matrix of the

otolith of a marine fish: Gadus morhua Linne, 1758 (Teleostei, Gadidae). Comparative

Biochemistry and Physiology. Part A: Molecular and Integrative Physiology 134, 551-561.

Davies, C.A., Brophy, D., Jeffries, T., Gosling, E. (2011) Trace elements in the otoliths and dorsal

spines of albacore tuna (Thunnus alalunga, Bonnaterre, 1788): An assessment of the

effectiveness of cleaning procedures at removing postmortem contamination. Journal of

Experimental Marine Biology and Ecology 396, 162-170. de Vries, M.C., Gillanders, B.M., Elsdon, T.S. (2005) Facilitation of barium uptake into fish otoliths:

Influence of strontium concentration and salinity. Geochimica et Cosmochimica Acta 69,

4061-4072.

Degens, E.T., Deuser, W.G., Haedrich, R.L. (1969) Molecular structure and composition of fish

otoliths. Marine Biology 2, 105-113.

Doubleday, Z.A., Harris, H.H., Izzo, C., Gillanders, B.M. (2014) Strontium randomly substituting for

calcium in fish otolith aragonite. Analytical Chemistry 86, 865-869.

134

Edmonds, J.S., Lenanton, R.C.J., Caputi, N., Morita, M. (1992) Trace-elements in the otoliths of

yellow-eye mullet (Aldrichetta forsteri) as an aid to stock identification. Fisheries Research

13, 39-51. [In English].

Elsdon, T.S., Gillanders, B.M. (2002) Interactive effects of temperature and salinity on otolith

chemistry: challenges for determining environmental histories of fish. Canadian Journal of

Fisheries and Aquatic Sciences 59, 1796-1808.

Elsdon, T.S., Gillanders, B.M. (2003) Relationship between water and otolith elemental

concentrations in juvenile black bream Acanthopagrus butcheri. Marine Ecology Progress

Series 260, 263-272.

Elsdon, T.S., Gillanders, B.M. (2004) Fish otolith chemistry influenced by exposure to multiple

environmental variables. Journal of Experimental Marine Biology and Ecology 313, 269-

284.

Elsdon, T.S., Wells, B.K., Campana, S.E., et al. (2008) Otolith chemistry to describe movements and

life-history parameters of fishes: Hypotheses, assumptions, limitations and inferences.

Oceanography and Marine Biology: An Annual Review, Vol 46 46, 297-+. [In English].

Florea, A.M., Busselberg, D. (2006) Occurrence, use and potential toxic effects of metals and metal

compounds. BioMetals 19, 419-427.

Friedrich, L.A., Halden, N.M. (2010) Determining exposure history of northern pike and walleye to

tailings effluence using trace metal uptake in otoliths. Environmental Science & Technology

44, 1551-1558.

135

Gillanders, B.M. (2002) Temporal and spatial variability in elemental composition of otoliths:

implications for determining stock identity and connectivity of populations. Canadian

Journal of Fisheries and Aquatic Sciences 59, 669-679.

Hicks, A.S., Closs, G.P., Swearer, S.E. (2010) Otolith microchemistry of two amphidromous galaxiids

across an experimental salinity gradient: A multi-element approach for tracking

diadromous migrations. Journal of Experimental Marine Biology and Ecology 394, 86-97.

Hoff, G.R., Fuiman, L.A. (1993) Morphometry and composition of red drum otoliths - changes

associated with temperature, somatic growth-rate, and age. Comparative Biochemistry

and Physiology a-Physiology 106, 209-219. [In English].

Hughes, I., Blasiole, L., Huss, D., et al. (2004) Otopetrin 1 is required for otolith formation in the

zebrafish Danio rerio. Developmental Biology 276, 391-402.

Izzo, C., Doubleday, Z.A., Gillanders, B.M. (2016) Where do elements bind within the otoliths of

fish? Marine and Freshwater Research 67, 1072-1076.

Izzo, C., Doubleday, Z.A., Schultz, A.G., Woodcock, S.H., Gillanders, B.M. (2015) Contribution of

water chemistry and fish condition to otolith chemistry: comparisons across salinity

environments. Journal of Fish Biology 86, 1680-1698.

Jolivet, A., Bardeau, J.-F., Fablet, R., Paulet, Y.-M., de Pontual, H. (2008) Understanding otolith

biomineralization processes: new insights into microscale spatial distribution of organic

and mineral fractions from Raman microspectrometry. Analytical and Bioanalytical

Chemistry 392, 551-560.

136

Kalish, J.M. (1991) Determinants of otolith chemistry - seasonal-variation in the composition of

blood-plasma, endolymph and otoliths of bearded rock cod Pseudophycis barbatus.

Marine Ecology Progress Series 74, 137-159. [In English].

Kang, Y.J., Stevenson, A.K., Yau, P.M., Kollmar, R. (2008) Sparc protein is required for normal

growth of zebrafish otoliths. J Assoc Res Otolaryngol 9, 436-451.

King, K., Jr. (1978) Distribution of gamma-carboxyglutamic acid in calcified tissues. Biochimica et

Biophysica Acta 542, 542-546.

Limburg, K.E., Olson, C., Walther, Y., Dale, D., Slomp, C.P., Hoie, H. (2011) Tracking Baltic hypoxia

and cod migration over millennia with natural tags. Proceedings of the National Academy

of Sciences of the United States of America 108, E177-E182.

Lothian, A., Hare, D.J., Grimm, R., Ryan, T.M., Masters, C.L., Roberts, B.R. (2013)

Metalloproteomics: principles, challenges and applications to neurodegeneration.

Frontiers in Aging Neuroscience 5, 35. [In English].

Lothian, A., Roberts, B.R. (2016) Standards for quantitative metalloproteomic analysis using size

exclusion ICP-MS. Journal of visualized experiments : JoVE. [In English].

Lundberg, Y.W., Xu, Y., Thiessen, K.D., Kramer, K.L. (2015) Mechanisms of otoconia and otolith

development. Developmental Dynamics 244, 239-253.

Melancon, S., Fryer, B.J., Gagnon, J.E., Ludsin, S.A. (2008) Mineralogical approaches to the study of

biomineralization in fish otoliths. Mineralogical Magazine 72, 627-637.

Melancon, S., Fryer, B.J., Ludsin, S.A., Gagnon, J.E., Yang, Z.P. (2005) Effects of crystal structure on

the uptake of metals by lake trout (Salvelinus namaycush) otoliths. Canadian Journal of

Fisheries and Aquatic Sciences 62, 2609-2619.

137

Melbourne Water (2014) Summary Waterway Water Quality Data 2014. Annual Water Quality

Factsheet, Long Term Water Quality Monitoring Sites, 1-90.

Miller, M.B., Clough, A.M., Batson, J.N., Vachet, R.W. (2006) Transition metal binding to cod otolith

proteins. Journal of Experimental Marine Biology and Ecology 329, 135-143.

Miyan, K., Khan, M.A., Patel, D.K., Khan, S., Prasad, S. (2016) Otolith fingerprints reveal stock

discrimination of Sperata seenghala inhabiting the Gangetic river system. Ichthyological

Research 63, 294-301. [In English].

Mugiya, Y., Hakomori, T., Hatsutori, K. (1991) Trace-metal incorporation into otoliths and scales in

the goldfish, Carassius auratus. Comparative Biochemistry and Physiology C-Pharmacology

Toxicology & Endocrinology 99, 327-331. [In English].

Mugiya, Y., Takahashi, K. (1985) Chemical properties of the saccular endolymph in the rainbow

trout Salmo gairdneri. Bulletin of the Faculty of Fisheries Hokkaido University 36, 57-63. [In

English].

Mugiya, Y., Tanaka, S. (1995) Incorporation of water-borne strontium into otoliths and its turnover

in the goldfish Carassius auratus - effects of strontium concentrations, temperature, and

17-Beta-Estradiol. Fisheries Science 61, 29-35. [In English].

Mugiya, Y., Yoshida, M. (1995) Effects of calcium antagonists and other metabolic modulators on

in vitro calcium deposition on otoliths in the rainbow trout Oncorhynchus mykiss. Fisheries

Science 61, 1026-1030. [In English].

Murayama, E., Herbomel, P., Kawakami, A., Takeda, H., Nagasawa, H. (2005) Otolith matrix

proteins OMP-1 and Otolin-1 are necessary for normal otolith growth and their correct

anchoring onto the sensory maculae. Mechanisms of Development 122, 791-803.

138

Murayama, E., Okuno, A., Ohira, T., Takagi, Y., Nagasawa, H. (2000a) Molecular cloning and

expression of an otolith matrix protein cDNA from the rainbow trout, Oncorhynchus

mykiss. Comparative Biochemistry and Physiology B-Biochemistry & Molecular Biology 126,

511-520.

Murayama, E., Takagi, Y., Nagasawa, H. (2000b) A study of the matrix protein in the fish otolith.

Zoological Science (Tokyo) 17, 52-52.

Murayama, E., Takagi, Y., Nagasawa, H. (2004) Immunohistochemical localization of two otolith

matrix proteins in the otolith and inner ear of the rainbow trout, Oncorhynchus mykiss:

comparative aspects between the adult inner ear and embryonic otocysts. Histochemistry

and Cell Biology 121, 155-166.

Payan, P., De Pontual, H., Boeuf, G., Mayer-Gostan, N. (2004) Endolymph chemistry and otolith

growth in fish. Comptes Rendus Palevol 3, 535-547.

Payan, P., Edeyer, A., De Pontual, H., Borelli, G., Boeuf, G., Mayer-Gostan, N. (1999) Chemical

composition of saccular endolymph and otolith in fish inner ear: lack of spatial uniformity.

American Journal of Physiology-Regulatory Integrative and Comparative Physiology 277,

R123-R131.

Petko, J.A., Millimaki, B.B., Canfield, V.A., Riley, B.B., Levenson, R. (2008) Otoc1: A novel Otoconin-

90 ortholog required for otolith mineralization in zebrafish. Developmental Neurobiology

68, 209-222.

Popper, A.N., Fay, R.R. (2011) Rethinking sound detection by fishes. Hearing Research 273, 25-36.

139

Romanek, C.S., Gauldie, R.W. (1996) A predictive model of otolith growth in fish based on the

chemistry of the endolymph. Comparative Biochemistry and Physiology a-Physiology 114,

71-79.

Rouleau, C., Xiong, Z.H., Pacepavicius, G., Huang, G.L. (2003) Uptake of waterborne tributyltin in

the brain of fish: Axonal transport as a proposed mechanism. Environmental Science &

Technology 37, 3298-3302. [In English].

Rozycka, M., Wojtas, M., Jakob, M., et al. (2014) Intrinsically disordered and pliable Starmaker-like

protein from medaka (Oryzias latipes) controls the formation of calcium carbonate

crystals. PloS One 9, e114308.

Shannon, R.D. (1976) Revised effective ionic-radii and systematic studies of interatomic distances

in halides and chalcogenides. Acta Crystallographica Section A 32, 751-767. [In English].

Shiao, J.C., Tzeng, C.S., Leu, C.L., Chen, F.C. (1999) Enhancing the contrast and visibility of daily

growth increments in fish otoliths etched by Proteinase K buffer. Journal of Fish Biology

54, 302-309.

Söllner, C., Burghammer, M., Busch-Nentwich, E., et al. (2003) Control of crystal size and lattice

formation by Starmaker in otolith biomineralization. Science 302, 282-286.

Sturrock, A.M., Hunter, E., Milton, J.A., et al. (2015) Quantifying physiological influences on otolith

microchemistry. Methods in Ecology and Evolution 6, 806-816. [In English].

Sturrock, A.M., Trueman, C.N., Darnaude, A.M., Hunter, E. (2012) Can otolith elemental chemistry

retrospectively track migrations in fully marine fishes? Journal of Fish Biology 81, 766-795.

140

Szpunar, J. (2005) Advances in analytical methodology for bioinorganic speciation analysis:

metallomics, metalloproteomics and heteroatom-tagged proteomics and metabolomics.

Analyst 130, 442-465. [In English].

Takagi, Y. (2002) Otolith formation and endolymph chemistry: a strong correlation between the

aragonite saturation state and pH in the endolymph of the trout otolith organ. Marine

Ecology Progress Series 231, 237-245.

Thorrold, S.R., Swearer, S.E. (2009) Otolith Chemistry. In: Tropical Fish Otoliths: Information for

Assessment, Management and Ecology. Vol. 11. Reviews-Methods and Technologies in Fish

Biology and Fisheries (Eds. B.S. Green, B.D. Mapstone, G. Carlos, G.A. Begg), pp. 249-295.

Tohse, H., Mugiya, Y. (2001) Effects of enzyme and anion transport inhibitors on in vitro

incorporation of inorganic carbon and calcium into endolymph and otoliths in salmon

Oncorhynchus masou. Comparative Biochemistry and Physiology a-Molecular and

Integrative Physiology 128, 177-184.

Tohse, H., Takagi, Y., Nagasawa, H. (2008) Identification of a novel matrix protein contained in a

protein aggregate associated with collagen in fish otoliths. FEBS Journal 275, 2512-2523.

Wang, X., Tan, Y., Sievers, Q., et al. (2011) Thyroid hormone-responsive genes mediate otolith

growth and development during flatfish metamorphosis. Comparative Biochemistry and

Physiology a-Molecular & Integrative Physiology 158, 163-168.

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2016) Not all inner ears are the same: Otolith matrix

proteins in the inner ear of sub-adult cichlid fish, Oreochromis mossambicus, reveal

insights into the biomineralization process. Anatomical Record-Advances in Integrative

Anatomy and Evolutionary Biology 299, 234-245.

141

Zhang, Z. (1992) Ultrastructure of otolith increments and checks in the teleost fish Oreochromis

niloticus. Journal of Morphology 211, 213-220. [In English].

142

CHAPTER FOUR: THE INNER EAR PROTEOME OF FISH (In press, FEBS Journal, Accepted December 2018)

ABSTRACT

The mechanisms that underpin the formation, growth and composition of otoliths, the biomineralized stones in the inner ear of fish, are largely unknown, as only a few fish inner ear proteins have been reported. Using a partial transcriptome for the inner ear of black bream

(Acanthopagrus butcheri), in conjunction with proteomic data, we discovered hundreds of previously unknown proteins in the otolith. This allowed us to develop hypotheses to explain the mechanisms of inorganic material supply and daily formation of growth bands. We further identified a likely protein mediator of crystal nucleation and an explanation for the apparent metabolic inertness of the otolith. Due to the formation of both daily and annual increments, otoliths are routinely employed as natural chronometers, being used for age and growth estimation, fisheries stock assessments, and the reconstruction of habitat use, movement, diet, and the impacts of climate change. Our findings provide an unprecedented view of otolith molecular machinery, aiding in the interpretation of these essential archived data.

INTRODUCTION

Otoliths are small, biomineralized earstones within the inner ear of fish that aid in gravity reception and balance (Lowenstein, 1936). Although understanding their functional anatomy and physiology is important to hearing and balance research, otoliths are especially valuable as biochronometers, archiving age and growth information essential for assessing fisheries stocks globally and predicting their responses to climate change, as well as elemental and stable isotope time series for reconstructing fish life histories of habitat use, movement, pollution exposure and diet (Izzo et al., 2016).

143

All bony fish have three pairs of otoliths, termed sagittae, lapilli and asterisci (Campana, 1999).

These differ in size (sagittae are usually the largest), location in the inner ear, and calcium carbonate polymorph. Sagittae and lapilli are typically aragonite (although calcite has been observed in some species (Weigele et al., 2016)), and asteriscii are often composed of vaterite

(Campana, 1999). In this study, the term “otolith” refers to sagittal otoliths. Otoliths grow continuously from embryo, being encased within fluid-filled sacs that are the end organs of the inner ear. The fluid growth medium, endolymph, is rich in both inorganic materials as well as proteins and other accessory macromolecules necessary for otolith formation and growth (Payan et al., 2004).

Several hypotheses have been suggested to explain the supply of inorganic constituents from the saccular epithelial cells to the endolymph. It has been proposed that calcium is supplied through either a paracellular route or actively through ATPase activity, whereas bicarbonate is transferred through exchange with chloride ions (Mugiya and Yoshida, 1995, Tohse and Mugiya, 2001, Payan et al., 2004, Cruz et al., 2009). However, the underpinning biochemical mechanisms for ion transfer, exchange and regulation within the endolymph have not been fully elucidated. Otoliths first form shortly after fertilization and grow incrementally, incorporating both inorganic and organic material from the endolymph. A consequence of this is that a physical record of endolymph change is created, allowing for a lifetime chronology to be established. The utility of otoliths as biochronometers is reliant upon a proper understanding of the protein-mediated mechanisms of both nucleation as well as increment formation. These mechanisms are currently poorly understood. Incremental otolith growth occurs via a daily cycle of deposition of alternating aragonite-rich (calcium carbonate) and protein-rich bands. Despite being first documented in 1971 by Panella (Panella, 1971), a suitable explanation for how this happens remains unresolved.

Finally, a fundamental assumption of the use of otoliths as biochronometers is that they are

144

metabolically inert and do not undergo any remodeling. Although a study has demonstrated that otoliths are only resorbed under extreme hypoxia (Mugiya and Uchimura, 1989), no mechanism has been identified for their apparent metabolic inertness. The reason that these fundamental questions remain unanswered is that fewer than a dozen otolith proteins have been identified. In comparison, the understanding of other biomineralizing systems is considerably more advanced: over 1300 proteins in mouse osteoblast extracellular matrix (Xiao et al., 2007) and more than 200 in dental tissues (i.e. dentin, enamel and cementum), which also grow incrementally, are known

(Charone et al., 2016, Salmon et al., 2017, Lima Leite et al., 2018).

Although the first constituent of the organic matrix of fish otoliths was described almost 50 years ago (Degens et al., 1969), to date only a handful of proteins have been successfully identified and characterized. These include Otolin-1 (sometimes called saccular or inner-ear specific collagen)

(Davis et al., 1995), Otolith Matrix Protein-1 (OMP-1) (Murayama et al., 2000), Starmaker (Stm)

(Söllner et al., 2003), Otolith Matrix Macromolecule-64 (OMM-64) (Tohse et al., 2008), Starmaker- like (Stm-l) (Bajoghli et al., 2009), an ortholog to mammalian otoconin 90 (Otoc1) (Petko et al.,

2008), Osteonectin (Sparc) (Kang et al., 2008), Neuroserpin (Kang et al., 2008), Tectorin (Weigele et al., 2016), Otogelin (Weigele et al., 2016), Transferrin (Weigele et al., 2016), and Myosin Light

Chain 9 (Weigele et al., 2016). In addition to these structural constituents, other proteins thought to be necessary for biomineralization have also been identified in the inner ear. These include the

Na+/K+ and Ca2+ ATPases (Blasiole et al., 2006), GP96 (Sumanas et al., 2003), and Carbonic anhydrase (CA) isozymes (Tohse et al., 2004). Other macromolecules, such as carbohydrates, proteoglycans, and lipoproteins, have also been detected in the inner ear (Takagi et al., 2000,

Borelli et al., 2001).

To understand in greater detail the protein machinery that composes the otolith we conducted a bottom-up proteomic approach to develop a comprehensive picture of the inner ear for black

145

bream (Acanthopagrus butcheri), a species which has been extensively studied in terms of otolith chemistry (Morison et al., 1998, Sarre and Potter, 2000, Elsdon and Gillanders, 2003, Elsdon and

Gillanders, 2004, de Vries et al., 2005, Elsdon and Gillanders, 2005, Ranaldi and Gagnon, 2008,

Partridge et al., 2009, Disspain et al., 2011, Sakabe et al., 2011, Webb et al., 2012, Williams et al.,

2018). As otoliths archive proteins from all life stages, we constructed a library composed of a mixture of both adult and juvenile peptides for use in label-free quantitative Mascot searches. The inclusion of adult and juvenile tissues provides a more comprehensive dataset of the proteins likely to be present in an otolith regardless of life stage. This is important for future studies that may compare the proteins in otoliths between various life stages. Briefly, we produced deep sequenced proteomic mass spectra through high pH reverse phase high performance liquid chromatography (RP-HPLC) fractionation (Gilar et al., 2005) followed by traditional low pH-LC tandem mass spectrometry (LC-MS/MS) of proteinaceous material from adult and juvenile otoliths as well as endolymph harvested from wild caught adults. The mass spectrometry peptide data were then matched with annotated cDNA libraries that had been created from homogenized adult brain tissue and juvenile heads and used to create a species-specific peptide library (Fig. 4.1,

Appendix 4.1; Table 4.1, 4.2). This was then used to identify peptides present in the otolith for adult biological replicates (from the same environment) following a traditional quantitative LC-

MS/MS approach. As otoliths are constantly bathed in endolymph, we also identified peptides in endolymph samples harvested at the same time as our otolith replicates. This allowed us to assess the likelihood of whether a given protein was present due to playing a functional role in biomineralization, or as an inescapable consequence of endolymph contamination – being either trapped during increment formation or adhering to the surface of the otolith.

146

Figure 4.1. Schematic of workflow for creation of a custom species-specific proteomics sequence database (Appendix 4.1).

147

RESULTS

High pH reverse phase fractionation of both otolith and endolymph samples for peptide library creation proved highly effective, yielding a roughly seven-fold increase in unique peptides when comparing fractions with starting materials (Table 4.1). Contig sequences from the cDNA FASTA files were successfully aligned with RefSeq proteins.

Table 4.1. Comparison of unique peptides occurring in starting material versus RP-HPLC fractions for pooled otolith and endolymph samples from Acanthopagrus butcheri. All peptides are Mascot Score ≥13 and E- value ≤ 0.01. Mass spectra were produced on a Thermo Scientific Orbitrap Fusion Lumos Tribrid Mass Spectrometer.

Peptides, starting Peptides, RP-HPLC Total material fractions Otoliths 459 3216 3261 Endolymph 2743 21080 21355

The percentage of significantly aligned proteins differed from file to file, however, with 45.6%,

47.7% and 38.9% for Adult Brain and each of the Juvenile Heads, respectively (Table 4.2). The top scoring RefSeq protein was selected for each alignment, and the redundancies removed.

Subsequently, the custom peptide library used for quantitative analysis contained 175,396 unique inner ear peptides from black bream, 944 proteins from Sparidae, 52,715 top hit proteins from

RefSeq, OMP-1 from L. crocea, and common contaminants (Appendix 4.1).

Table 4.2. Results of BLASTX v2.60+ alignment of Acanthopagrus butcheri cDNA FASTA with RefSeq proteins. RefSeq proteins were retrieved on May 28 2017, ftp://ftp.ncbi.nlm.nih.gov/refseq/release/, and totalled 84,756,971. An E-value of 1E-6 was employed as a threshold. All search results were sorted hierarchically based on Bitscore, E-value and Percentage Identity.

FASTA file No. of cDNA contigs No. of successful alignments No. of best scoring alignments Adult Brain 92273 17963561 42033 Juvenile Head 1 70334 14465474 33553 Juvenile Head 2 44402 6997060 17258

148

Mascot searches against the custom library for label-free quantitative assessment of inner ear protein abundances resulted in the identification of 1,065 otolith peptides and 2,685 endolymph peptides. There were 997 total protein types present in otolith and endolymph tissue, with 25.3% occurring in the otolith, 61.1% in the endolymph and only 13.6% of these occurring in both sample types (Fig. 4.2). Of those proteins occurring in the otolith, 54.4% were common to all three replicates.

Figure 4.2. Comparison of protein distributions, based on unique peptides, occurring in otolith and endolymph samples from adult Acanthopagrus butcheri (n=3). All samples underwent identical CNBr/Trypsin digestion protocols and were sequenced on a Thermo Scientific NanoLC-Q Exactive Plus Hybrid Quadrupole- Orbitrap Mass Spectrometer. All peptides are Mascot Score ≥13 and E-value ≤ 0.01. Venn diagram created with http://bioinfogp.cnb.csic.es/tools/venny/

We were able to identify the presence and relative abundance, at the peptide level, of over 380 proteins present in the otolith, including all previously identified otolith matrix proteins as well as

42 uncharacterized proteins (Fig. 4.3, Appendix 4.1). Please note here that feature intensity is a

149

guideline to sample protein abundance. This can be affected by several factors including how well a given protein is extracted from the otolith, the effectiveness of it undergoing trypsin digestion, and the readiness of individual peptides to ionize within the MS/MS.

Figure 4.3. All proteins, based on proteotypical peptides, found in the otoliths of adult black bream (Acanthopagrus butcheri, n=3). Previously identified inner ear proteins are highlighted in red and labelled. Heights of bars are indicative of protein abundance in the otolith, based on logarithm+1 transformed average feature intensity. All data were collected on a Thermo Scientific NanoLC-Q Exactive Plus Hybrid Quadrupole-Orbitrap Mass Spectrometer. Feature intensities were calculated using MaxQuant v1.5.8.3. The full list of peptides with associated feature intensities is available as Appendix 4.1.

A truncated list of inner ear proteins is presented, along with their ratios of otolith:endolymph abundance, in Fig. 4.4. Proteins presented are either known fish inner ear proteins or known to be involved in vertebrate biomineralization.

We identified several proteins from our quantitative proteomic data that are likely to be involved in inorganic material secretion; Sodium/potassium-transporting ATPase (Na+/K+-ATPase), V-type proton ATPase (V-ATPase), Band 3 anion exchange protein (B3AE), Sodium/calcium exchanger 2

(Na+/Ca2+-exchanger 2) and Calmodulin (CaM). We also found peptides aligning with Plasma membrane Ca2+-ATPase (PMCA) in our deep sequence data. We found two isozymes of CA; CA 1 and CA 6. Interestingly, the former was restricted to the endolymph and the latter to the otolith.

150

Figure 4.4. Heat map of selected proteins occurring in the inner ear of black bream (Acanthopagrus butcheri). Relative protein abundances are represented as the logarithm ratio of feature intensities (FI) of peptides in otolith relative to endolymph (n=3). All data were collected on a Thermo Scientific NanoLC-Q Exactive Plus Hybrid Quadrupole-Orbitrap Mass Spectrometer. Feature intensities were calculated using MaxQuant v1.5.8.3. Proteins presented have either been previously identified in the fish inner ear or known to be involved in biomineralization. The full list of peptides with associated feature intensities is available as Appendix 4.1.

We found, enriched in the otolith, two key proteins in nucleation and polymorph control: a phosphorylated putative Stm homolog as well as a kinase known to be involved in

151

biomineralization, Extracellular serine/threonine protein kinase FAM20C (FAM20C) (Tagliabracci et al., 2012). Stm and its homologs are intrinsically disordered proteins, and thus have low sequence homology, despite playing similar functional roles (Kaplon et al., 2008, Uversky and Dunker, 2010,

Rozycka et al., 2014, Poznar et al., 2017). Consequently, we identified the black bream Stm homolog through Ward’s minimum variance hierarchical clustering method (Fig. 4.5, Materials and

Methods). Our constellation plot revealed high levels of conservation between all otolith proteins

(except the Stm homologs) across taxa. Additionally, we undertook an error tolerant search of our proteomics data which identified that our putative Stm homolog has at least 4 phosphorylated serine residues (Appendix 4.1).

We, like others, also found considerable levels in the otolith of Otolin-1, OMP-1, and Neuroserpin, a member of the serine protease inhibitor family (Murayama et al., 2004, Takagi et al., 2005, Kang et al., 2008, Weigele et al., 2016). In addition to this, we identified Sparc, which plays a major role in the biomineralization of bone (Kang et al., 2008). Unlike previous studies, however, we also found the presence of other substrates to Neuroserpin and Sparc in the otolith and/or the endolymph: Plasminogen, Matrix metalloproteinase 2 (MMP 2), and Tissue inhibitor of metalloproteinase 2 (TIMP 2) (Mazzieri et al., 1997).

152

Figure 4.5. Constellation plot illustrating the clustering of inner ear proteins from a range of fish species; Acanthopagrus butcheri, Danio rerio, Lates calcarifer, Larimichthys crocea, Oryzias latipes, Oncorhynchus mykiss, Oreochromis niloticus, Salmo salar, Seriola lalandi, Stegastes partitus, Takifugu rubipes, as well as Homo sapiens. The putative Starmaker homolog from A. butcheri (“Unknown”) clusters closely with Otolith Matrix Macromolecule-64 (OMM-64), the Starmaker functional homolog in O. mykiss. The distribution of Dentin sialophosphoprotein (DSPP), Starmaker and their homologs is as expected for Intrinsically Disordered Proteins. In contrast, Transferrin, Neuroserpin, Otolin-1 (Otolin), Otolith Matrix Protein-1 (OMP-1), and Osteonectin (SPARC) show high levels of conservation among species. The following data were analysed, for each protein, using the protein parameter tool at ExPASy (https://web.expasy.org/protparam/); molecular weight, theoretical isoelectric point, percentage amino acid composition, instability index, aliphatic index and grand average of hydropathy (GRAVY). The plot was constructed, in JMP v 13 (SAS Institute Inc.) using Ward’s minimum variance hierarchical clustering method on standardised data. In this approach, the distance between two clusters (represented by the length of the line in the plot) is the ANOVA sum of squares between clusters summed over all variables. Clustering at each subsequent level of the hierarchy is achieved by merging clusters from the previous level, which minimizes the within-cluster sum of squares relative to all possible clusters.

153

In addition to the ‘usual’ otolith proteins, we found high abundances of bone and cartilage type proteoglycans (Mimecan (Osteoglycin), Epiphycan, Biglycan, Decorin, Aggrecan) , Collagen types I and II, and osteoblast regulating proteins (e.g. E3 ligase itchy-like) (Iozzo, 1998, Xiao et al., 2007,

Zhen et al., 2008, Ye et al., 2012, Liu et al., 2017). Importantly, we found no evidence of protein markers for osteoclastic function. Our data set also reveals considerable levels of Apolipoprotein

A1 (ApoA1) and Transferrin as well as five key proteoglycans, although the former two, being prime components of the endolymph, are probably the result of endolymph fluid trapping during increment deposition rather than their playing a functional role. It is possible, however, that

ApoA1 may be involved in initiating increment formation through binding a product of polyketide synthase, an enzyme recently implicated in proper otolith biomineralization (Hojo et al., 2015).

Similarly, it cannot be ruled out that Transferrin, an iron transporter, plays a role in otolith growth

(Lambert, 2012).

DISCUSSION

Supply of inorganic materials

The necessary organic and inorganic materials for otolith growth are secreted from saccular epithelial cells into the endolymph. For aragonite growth to occur, a sufficient proton gradient must be present along with bicarbonate and calcium ions (Payan et al., 2004). It is likely that, at the epithelial cell plasma membrane, bicarbonate ions are being supplied to the endolymph by

B3AE whereas calcium is transported by a combination of PMCA and Na+/Ca2+-exchanger 2. B3AE is a transmembrane protein that has been previously observed to regulate calcification in kidney duct cells (Drenckhahn et al., 1985). Carbonic anhydrase (CA) has been implicated in otolith growth, its expression following a diurnal pattern (Tohse et al., 2006). It is likely that CA 1 acts to produce bicarbonate ions from carbon dioxide in the epithelial cells, which is then supplied to the

154

endolymph. At the growing edge of the otolith, CA 6 is likely to be cooperating with V-ATPase in the plasma membrane of epithelial cells to maintain pH while also maintaining the observed constant supersaturation of bicarbonate in the endolymph (Takagi et al., 2005). These two proteins have been suggested to provide the same function during enamel maturation in teeth

(Yin and Paine, 2017). In Fig. 4.6, we present our hypothesis for the supply of inorganic components to the endolymph, expanding on previous models (Payan et al., 2004).

Figure 4.6. Schematic of hypothetical ion transport mechanisms in the inner ear of black bream (Acanthopagrus butcheri). The sagittal otolith pictured is in the sacculum, a fluid-filled sac in the inner ear. The saccula are located in the cranium – behind the brain and approximately level, vertically, with the eyes. In the epithelial cell, Carbonic anhydrase 1 (CA 1) converts carbon dioxide to bicarbonate ions and protons. The bicarbonate ions are transferred across the cell membrane by Band 3 anion exchange protein (B3AE), which exchanges them for chloride ions in the endolymph. Calmodulin (CaM) transfers intracellular calcium to Na+/Ca2+-exchanger 2 and Plasma membrane Ca2+-ATPase (PMCA). Na+/Ca2+-exchanger 2 and PMCA transfer calcium into the endolymph. At the growing surface of the otolith, bicarbonate ions precipitate with calcium as aragonite needles. The carbon dioxide created as a by-product of this reaction is converted to protons and bicarbonate in situ by Carbonic anhydrase 6 (CA 6). The protons are pumped out of the endolymph by V-type proton ATPase (V-ATPase). The combined action of V-ATPase and CA 6 maintain the necessary pH for aragonite formation while at the same time ensuring that the endolymph is supersaturated with bicarbonate ions. Sodium/potassium ATPase (Na+/K+ ATPase) generates the electrochemical potential necessary for all active transport processes.

155

The importance of phosphorylation in nucleation and polymorph control

The nucleus (or primordium) of the otolith can be described as a dense organic core surrounded by a rigid calcified layer. The very early expression patterns of Stm in the inner ear (Söllner et al.,

2003) suggests it is involved in initial otolith seeding. Stm, Stm-l, OMM-64, and the functionally similar human Dentin sialophosphoprotein (DSPP) are intrinsically disordered proteins (IDPs), having low sequence homology but comparable functional roles in biomineral growth. As Stm and

OMM-64 regulate crystal polymorph selection in the otoliths of zebrafish (Danio rerio) and rainbow trout (Oncorhynchus mykiss) respectively, it is likely that our Stm homolog performs the same function in black bream. Söllner (Söllner et al., 2003) demonstrated in the seminal knockout studies of Stm, that a lack of this protein invariably resulted in calcite and malformed “star- shaped” otoliths. From this it was inferred that Stm both coordinates calcium ions for proper precipitation as aragonite as well as restricts the plane of crystal growth. Similarly, an in vitro study of OMM-64 and Otolin-1 suggested a cooperative role where absence of the former led to calcite being formed, whereas absence of the latter resulted in vaterite (Tohse et al., 2009). Considering that calcite is the naturally occurring inorganic form of calcium carbonate, the fact that the protein scaffold facilitates the precipitation of aragonite, which is normally only stable at much higher temperatures, is of prime interest.

Wojtas et al. (Wojtas et al., 2012) showed that the ability of Stm to act as a crystal nucleation factor is directly related to its degree of phosphorylation. In that study, the authors phosphorylated Stm in vitro through action with Casein Kinase II (CK II) and examined its effectiveness as a crystal nucleator and regulator. The authors suggest that the increased negative charge due to phosphorylation allows greater binding of calcium, and later allows for inhibition and control of crystallization following nucleation. A similar regulation of nucleation and mineral growth through phosphorylation has been shown in amelogenin in enamel (Wiedemann-Bidlack et

156

al., 2011, Le Norcy et al., 2018). It has also been shown that, with IDPs, an increased degree of calcium binding leads to greater compaction of the protein, allowing for closer packing and thereby regulating the extent of crystal growth (Wojtas et al., 2012, Poznar et al., 2017).

In inorganic chemical analyses of otolith cores, the primordium has been found to be enriched in both manganese and phosphorus (Shima and Swearer, 2009). Until now, no suitable explanation has been given for this observation. In a previous study, we identified the presence of phosphorus exclusively in the high molecular weight fractions of endolymph (Thomas et al., 2017). The kinase

FAM20C has two manganese (II) ion cofactors and has been found to regulate biomineralization pathways through phosphorylation of the small integrin-binding ligand, N-linked glycoprotein

(SIBLING) family, which are found in the extracellular matrices of dentin and bone (Tagliabracci et al., 2012). The ortholog of Stm, DSPP, is a member of this family. FAM20C acts by targeting a

Serine-X-glutamic acid/phosphoserine (S-X-E/pS) motif (Tagliabracci et al., 2012).

We suggest, considering FAM20C’s presence in our data (as well as the absence of other suitable kinases such as CK II), coupled with the evidence of manganese and phosphorus enrichment at the otolith core, that it is acting to phosphorylate the black bream Stm homolog during the crucial early period of nucleation. Only with effective and complete phosphorylation of Stm can the proper polymorph, aragonite, be selected during seeding, thereby setting the template for crystal growth. Furthermore, Reimer et al. (Reimer et al., 2017) observed an increased prevalence of vaterite among otoliths from farmed salmon and hypothesized that this was due to a shift in the

Otolin-1/OMM-64 ratio and Gauldie (Gauldie, 1986) detected increased levels of phosphorus in aragonite otoliths when compared with vaterite otoliths. We hypothesise that it is not the ratio of these proteins that is of importance to polymorph selection, but rather the degree of phosphorylation of the Stm homolog that is present both during nucleation as well as later, during increment deposition.

157

These hypotheses (i.e. the roles of Stm and FAM20C in nucleation and polymorph selection) can be tested in two ways. Firstly, one can investigate the levels of expression of these proteins in embryonic fish. Secondly, to further understand their roles in polymorph selection, one could compare the proteins and associated degrees of phosphorylation between asterisci (vaterite) and sagittae (aragonite) otoliths.

Diel entrainment of increment deposition

Increment deposition is well characterized as occurring diurnally, with the main organic and mineral components of an increment being deposited during the crepuscular periods of the night:day cycle. As the endolymph is constantly supersaturated with respect to aragonite, Takagi et al. speculated that the timing of increment deposition is purely a function of organic molecules

(Takagi et al., 2005). The main structural proteins in the organic matrix are Otolin-1, which has sequence homology to Collagen X as well as a C-terminus C1q domain, and OMP-1, which contains a transferrin-like domain (Murayama et al., 2000, Murayama et al., 2002). Otolin has been found to interact with Otoconin 90 in the formation of otoconia in mammals (Deans et al., 2010). In fish it may act as a general interaction hub for otolith organic matrix proteins (such as OMP-1) through trimerization of its C1q domain (Holubowicz et al., 2017). Consequently, we suggest that it specifically acts as a stabilizing molecule for the organic scaffold (Murayama et al., 2005), providing a link between the various collagens and abundant OMP-1 molecules in the matrix. This hypothesis is consistent with previous studies demonstrating no effect from knockdown of OMP-1 or Otolin-1 on nucleation (Murayama et al., 2005), and that the timing of expression of both proteins coincided with increment deposition (Murayama et al., 2004, Weigele et al., 2017).

158

Considering this underpinning role for Otolin-1, any circadian regulation of the organic matrix must occur through interaction with this crucial molecule. We found both MMP 2 and TIMP 2, proteins whose expression likely follows a circadian pattern as knockout studies of the clock gene PER2 has resulted in a shift in their levels (Su et al., 2017). MMP 2 catalyzes the cleavage of several types of collagen, including type X. Considering that MMP 2 is also able to cleave those collagen types occurring in the macular epithelium, organic matrix, otolith membrane and kinocilia, we suggest that regulation of MMP 2 activity – specifically cleavage of Otolin-1 – is essential for the cyclical growth pattern of the organic matrix. As Otolin-1 is likely to be an interaction hub for the otolith matrix proteins, any cessation of organic growth can be achieved through its proteolysis.

Mineralization likely occurs at dawn, with Otolin-1 expression occurring throughout the night

(Mugiya, 1987, Zhang and Runham, 1992, Borelli et al., 2003, Tohse et al., 2006). Our adult fish were caught in the late afternoon. This is reflected in the relatively high abundance of Otolin-1 in our endolymph samples (Fig. 4). We theorize that at dawn, organic matrix growth is halted through a proteolytic cascade. Specifically, the inactive protease Plasminogen is coordinated and targeted to Otolin-1 by Sparc. Sparc is multi-functional, being able to bind to calcium as well as collagens, and is therefore likely to have roles in both chaperoning collagen to the organic matrix and aiding mineral precipitation (Kang et al., 2008). Sparc has been additionally shown to be able to enhance the activity of Plasmin (the active form of Plasminogen) (Kang et al., 2008).

Plasminogen is converted to its active form, the protease Plasmin, by either Tissue-type plasminogen activator (tPA) or Urokinase plasminogen activator (uPA). Indeed, examination of our deep-sequenced proteomic data indicates the presence of tPA in the otolith, suggesting that its expression may also be circadian in nature, with its absence from the endolymph likely because we collected fish in the late afternoon. Plasmin then activates MMP 2 (Mazzieri et al., 1997), which cleaves Otolin-1, halting growth of the organic matrix and allowing aragonite mineralization to

159

begin. At dusk, we suggest that this proteolytic cascade is inhibited by TIMP 2 and Neuroserpin, allowing growth of the organic matrix to recommence. TIMP 2 inhibits any active MMP 2 in the endolymph, whereas Neuroserpin halts any further activation of MMP 2. Neuroserpin is well established as a regulator of proteolytic cascades through its inhibition of tPA and uPA. A flowchart of our proposed mechanism for circadian control of organic matrix growth is presented in Fig. 4.7. Whilst our proposed mechanism is hypothetical, we believe that it can be effectively tested through an immunohistochemistry study focused on the timing of expression of these proteins (and their interactions) over a 24 h period, in combination with genetic manipulation, in a model organism such as D. rerio. This would be especially valuable as very little is known about the daily timing of otolith proteins other than Otolin-1 and OMP-1.

160

Figure 4.7. Proposed regulation of diel-entrained organic matrix deposition. It is believed that the organic phase is deposited at dusk and that mineralization occurs at dawn (Mugiya, 1987, Zhang and Runham, 1992, Borelli et al., 2003, Takagi et al., 2005). At dusk, Otolin-1 is expressed, forming interaction hubs with other structural proteins (Holubowicz et al., 2017). The organic matrix continues to be deposited until dawn, where the action of Matrix metalloproteinase 2 (MMP 2) cleaves Otolin-1 hubs, halting organic deposition, and allowing mineral precipitation to commence. The activation of MMP 2 is through a proteolytic cascade. At dusk, this cascade is inhibited from activating further MMP 2 by Neuroserpin. Any still active MMP 2 are inhibited by the action of Tissue Inhibitor of Metalloproteinase 2 (TIMP 2). MMP2 and TIMP 2 levels have been shown to be affected by the clock gene, PER2 (Su et al., 2017).

161

Otoliths are like bones but inert

We found the presence of many other proteins involved in the biomineralization of bone, cartilage or teeth in other vertebrates (Fig. 4.8) (Iozzo, 1998, Xiao et al., 2007, Zhen et al., 2008, Salmon et al., 2017).

Figure 4.8. Proteins common to bone and other biomineralization systems found in the inner ear of adult black bream (Acanthopagrus butcheri, n=3) (Iozzo, 1998, Xiao et al., 2007, Zhen et al., 2008, Salmon et al., 2017). In addition to structural collagens and proteoglycans, we also found proteins that are associated with osteoblast function or otherwise act as substrates for extracellular matrix formation (Kohfeldt et al., 1998, Siebers et al., 2005, Jeong et al., 2007, Liu et al., 2017).

This suggests that the macromolecules involved in biomineralization are highly conserved - a given protein may carry out the same function in separate systems, regardless of whether the mineral is hydroxyapatite (bone, cartilage or teeth) or calcium carbonate (otolith). Interestingly, while we found an abundance of osteoblast-associated proteins (e.g. E3 ligase itchy-like, Collagen triple helix repeat-containing protein 1 (Siebers et al., 2005, Jeong et al., 2007, Liu et al., 2017)), we found no typical markers of osteoclast activity such as Carbonic anhydrase 2, Cathepsin K,

Calcitonin receptor, Cell surface aminopeptidase N/CD13, or Integrin β3 (Castillo et al., 2017).

162

Cathepsin K, particularly, is abundantly expressed by osteoclasts (Drake et al., 1996). In contrast to this, evidence of osteoblast-like activity has been observed in previous studies of fish otoliths, notably with Osterix (a protein that differentiates osteoblasts) being identified in the piscine inner ear (Renn and Winkler, 2014). In a two-way biomineralizing system, which involves the alternating processes of growth and remodeling, the lack of one of these protein types would be unusual. The lack of osteoclastic pathways supports the long-held assumption that otolith formation is unidirectional and doesn’t undergo remodeling. This provides further support to the assumption that otoliths are ‘permanent’ archives of fish life histories.

CONCLUSION

In conclusion, we identified over 380 proteins in the otoliths of adult black bream. As the majority of these are unlikely to be directly involved in biomineralization, it suggests that a diversity of proteins present in the endolymph are being trapped in the otolith during increment formation.

This raises the intriguing possibility that the otolith is not only archiving elemental markers of environmental history, but also protein markers of development and physiological change over an individual’s lifetime. Depending on the lability and spatial resolution of these biomarkers, this has the potential to reconstruct the life histories of fishes to an unprecedented level, opening up new and exciting avenues of research in ecology and fisheries science.

163

MATERIALS AND METHODS

Animals

Juvenile Acanthopagrus butcheri were caught via seine net either on the Werribee River, Australia

(37°57'45.6"S, 144°40'02.9"E, cDNA library creation) in May 2016 or in the Gippsland Lakes

(37°51'28.3572''S, 147°44'39.0984''E, peptide library creation) in July 2017. Juveniles were measured (5-7 cm fork length) and then euthanised. Following this, they were immediately placed on dry ice until their return to the University of Melbourne, and stored at -80 °C. Adult A. butcheri were caught via gill net on the Werribee River, Australia over a range of dates; August 2015,

March 2016, February and June 2017. Adult fish were measured (30-38 cm fork length), euthanised and decapitated at the first cervical vertebra. Following decapitation, the otoliths, endolymph and brain tissue (which also included saccular tissue) were removed and individually stored on dry ice in labelled microcentrifuge tubes. Our method for careful removal of otoliths and endolymph has been described previously (Thomas et al., 2017). Upon return from the field, all samples were stored at -80 °C at the University of Melbourne. Fish were collected under Arthur

Rylah Institute Animal Ethics Committee approval 14/12 and the State of Victoria, Australia

Department of Environment and Primary Industries’ Fisheries research permits #1135 and #1204.

Chemicals and standards

All standards, samples and buffers were prepared using ultra-pure Milli-Q H2O (18.2MΩ; Merck

Millipore, Australia). Unless otherwise indicated, all chemicals were purchased from Sigma Aldrich,

Castle Hill, Australia and were of the highest available purity.

164

Transcriptome sequencing, assembly and annotation

Two juveniles were selected for transcriptome assembly. They were thawed and decapitated at the first cervical vertebra. Their heads were homogenised in RNALater™ and refrozen for subsequent RNA extraction. A sample of adult tissue (consisting of the brain, as well as saccular tissue) was similarly homogenised in RNALater™. Total RNA was extracted and purified from each sample using a QIAGEN RNeasy® Plus Mini Kit, following the supplied protocol. All samples were mRNA-purified, converted to cDNA and prepared for sequencing as per the Low Sample protocol,

Illumina® TruSeq® RNA Sample Preparation v2 Guide. cDNA libraries were sequenced on an

Illumina® NextSeq 500® at the Walter and Eliza Hall Institute of Medical Research (WEHI). cDNA sequences were then assembled using Trinity (v2.2.0). FastQC (v0.11.5) was initially employed to generate quality control summaries for the input FASTQ files. The results from this suggested that there was a small amount of adapter contamination at the beginning of the reads. Consequently, three similar assembly pipelines were run. The first used the default Trinity pipeline whereas the second and third trimmed the input reads by removing the first 5 bases and running Trimmomatic

(v 0.36) respectively. This ultimately produced three cDNA FASTA files – two for juvenile heads, and one for adult brain.

The three cDNA FASTA files were then annotated by using BLASTX (v2.6.0+) to compute alignments between the query cDNA contigs and the complete RefSeq protein database (retrieved on May 28 2017, ftp://ftp.ncbi.nlm.nih.gov/refseq/release/) with a threshold Expect value (E- value) of 1E-6, on a supercomputer (X99-PRO server, Xeon E7 processors, 32GB RAM, Plant

Genome and Molecular Evolution Lab, Department of Life Science, Tunghai University, Taiwan).

The output was parsed to a tabular format and the matches per query sequence were sorted hierarchically based on Bitscore, E-value, and Percentage Identity. The best matches for each query were extracted, compiled to a table, and imported into OpenRefine (v2.7).

165

Tissue preparation for peptide library creation

Otoliths from wild caught juveniles (n=10, 6-28 mg) and adults (n=6, 74-120 mg) were dissolved in an excess of 20% w/v trichloroacetic acid (TCA, Chem-Supply, Australia) for 2h. TCA at this concentration has been demonstrated to be optimal for otolith protein extraction, as it both dissolves the mineral phase as well as precipitates the entirety of the protein phase while minimizing protein denaturation (Weigele et al., 2016). The dissolved otoliths were then centrifuged for 30 min at 16000 x g (Eppendorf 5415 D), after which the supernatant was discarded, and the remaining pellets washed with 100 µL of -20 °C acetone and recentrifuged for 5 min at 16000 x g. The pellets were air-dried before being resuspended in 1mL of 90% v/v trifluoroacetic acid (TFA) in the presence of 50 mg of cyanogen bromide (CNBr). These were then incubated, in the dark, at ambient conditions for 48 h. Finally, the suspensions were frozen at -80

°C and lyophilized (Christ Alpha 1-4 LD Plus). Protein content of each otolith was determined through the Bradford Assay indicating that 0.1-0.5% w/w protein was extracted from each otolith.

Endolymph samples from wild caught adults (n=4) and the lyophilized otolith material were suspended in 100 µL of 8 M Urea 100 mM ammonium bicarbonate (ABC). The protein content of each sample was determined through the Bradford Assay. Tris(2-carboxyethyl)phosphine (TCEP,

ThermoFisher, Australia) was added to a final concentration of 10 mM and incubated at 60°C for

10 min. After this, the solution was alkylated through the addition of Iodoacetamide (IAA) to 40 mM and incubated at 37°C in the dark for 1 h. Each sample then underwent an eightfold dilution with 50 mM ABC/1 mM calcium chloride Finally, trypsin (Sigma T1426 bovine trypsin) was added at a weight ratio of 1:50 and the entire mixture allowed to incubate at 37°C, in the dark, for 12 h.

Following trypsin digestion, the samples were acidified in 10% v/v TFA and underwent C18 clean- up on a Waters Positive Pressure-96 Processor, employing 99.9% ACN/0.1% TFA as the initial medium, 0.1% TFA for equilibration and washing, and eluting in 70% v/v ACN/0.1% Formic acid

166

(FA). The cleaned material was vacuum dried (Savant SC100 SpeedVac) and then resuspended in

100 µL of 5% ACN/0.2% ammonium hydroxide (NH4OH).

Samples were pooled to create both an otolith and an endolymph sample that each contained 200

µg of peptides. These pooled samples were fractionated, following the method of Gilar et al. (Gilar et al., 2005), through high pH reverse phase high performance liquid chromatography (RP-HPLC) on an Agilent Technologies 1200 Series Infinity II fitted with a Waters Acquity UPLC BEH 300 Å column. The buffers employed were 0.2% NH4OH and 90% ACN/0.2% NH4OH. Fractions were collected every 0.7 min, resulting in 32 fractions collected from each starting material. These were concatenated to 16 samples by adding fractions that were 16 fractions apart (e.g. Fraction 1 +

Fraction 17, Fraction 2 + 18 etc.). The starting material was retained. Finally, the 34 subsequent samples (16 concatenated otolith, 16 concatenated endolymph, and corresponding starting materials) were vacuum dried and resuspended in 50 µL of a 2% ACN/0.05% TFA buffer. Each of the 34 samples was individually separated and sequenced on a Thermo Scientific Orbitrap Fusion

Lumos Tribrid Mass Spectrometer (LC-MS/MS) at the Mass Spectrometry and Proteomics Facility at the Bio21 Institute, the University of Melbourne, Australia.

Custom peptide library creation

The resulting raw mass spectra from the fractionated endolymph and otolith material were run through MaxQuant (v1.5.8.3) to provide peptide feature intensities. Following this, the

Andromeda Peak Lists (APL) were converted to Mascot Generic Format (MGF) using the WEHI APL to MGF converter, before being searched with Mascot (v2.4.0) on the WEHI Mascot server

(https://sysbio-mascot.wehi.edu.au/mascot/home.html) against our annotated cDNA FASTA files.

This search was conducted in six reading frames, and generated peptide-cDNA matches, while retaining accession information from Trinity. We employed the following thresholds for all

167

searches: a Mascot score of 13 or higher and E-value of 0.01 or lower. These transcript-matched peptides were then combined, in a database, with all currently known Sparidae proteins (retrieved on January 3 2017, http://www.uniprot.org). In addition to this, to capture any missing peptide sequences, we also included all top scoring proteins from the results of the transcriptome BLASTX alignments with RefSeq. As a preliminary evaluation of our transcripts revealed the absence of the cDNA entry corresponding to OMP-1, we included the full sequence of this from giant yellow croaker, (Larimichthys crocea). We did this because, while OMP-1 mRNA expression is constant in rainbow trout (Murayama et al., 2004, Takagi et al., 2005), it is possible that it was either not present at the time of death or was otherwise not captured by our mRNA extraction. Finally, we also included common contaminant sequences (e.g. trypsin, keratin, alpha-2-casein, serum albumin and others). Our custom peptide library is available as Appendix 4.1.

Quantitative assessment

In order to compare the presence of proteins in endolymph and otoliths extracted from individual fish, we undertook a label-free quantitative proteomic approach. Individual otoliths from wild caught adult bream (n=3, 37-59 mg) were manually abraded on all sides with sandpaper (3M

Wetordry 1000 grit) to physically remove any adhering membranous tissue or endolymph, and then individually dissolved in excess 20% w/v TCA. Endolymph samples (n=3, 34-90 µL) were thawed to room temperature and similarly dissolved. Both tissue types were then put through the

CNBr/trypsin digest and C18 clean-up protocols as described above for otoliths. Following clean- up, each sample (representing either endolymph or an otolith from an individual fish) was resuspended in 2% ACN/0.05% TFA to give an estimated protein concentration of 0.35 mg/mL.

Each sample was then individually run on a Thermo Scientific NanoLC-Q Exactive Plus Hybrid

168

Quadrupole-Orbitrap Mass Spectrometer at the Mass Spectrometry and Proteomics Facility at the

Bio21 Institute, University of Melbourne, Australia.

Mass spectra from each otolith (n=3) and endolymph (n=3) sample were similarly analysed with

MaxQuant, converted to MGF and searched with Mascot on the WEHI Mascot server against our custom database (peptides from RP-HPLC, Sparidae proteins, top-scoring proteins from BLASTX, L. crocea OMP-1, contaminants). We employed the same E-value and Mascot score thresholds as described above. Additionally, we performed decoy searches (i.e. utilizing amino acid sequences that had been reversed, shuffled, or Markov chain-generated) to validate these thresholds as per

Käll et al. (Kall et al., 2008). The outputs from these searches were ranked by feature intensity, providing an estimate of the most to least abundant peptides (and therefore proteins) in each sample. The average feature intensity (FI) for each peptide was determined for the three otolith samples and the three endolymph samples and log 2 (x+1) transformed (Fig. 4.3, 4.4). In addition to this, to identify potential post-translational modifications such as phosphorylation and glycosylation, we also performed an error tolerance Mascot search on our data. A flowchart of our library-creation and quantitative analysis methods is presented in Fig. 4.1. The parameters employed are presented in Table 4.3.

Identification of a starmaker-like protein

Preliminary Mascot searches revealed peptides for a candidate starmaker-like protein, and its full primary sequence was inferred from its cDNA entry. This sequence was analysed using the protein parameter tool at ExPASy (https://web.expasy.org/protparam/) to provide molecular weight, theoretical isoelectric point, percentage amino acid composition, instability index, aliphatic index and grand average of hydropathy (GRAVY). In addition to the candidate starmaker-like protein,

169

parameters were also extracted from the primary sequences of Otolin-1 (or its homolog), OMP-1,

Transferrin, Neuroserpin, Starmaker/OMM-64/Dentin sialophosphoprotein (DSPP), and Sparc for black bream as well as the following species: Danio rerio, Lates calcarifer, Larimichthys crocea,

Oryzias latipes, Oncorhynchus mykiss, Oreochromis niloticus, Salmo salar, Seriola lalandi, Stegastes partitus, Takifugu rubipes and Homo sapiens. These data were then analysed using Ward’s minimum variance clustering method, and visualized as a constellation plot, labelling our candidate protein as an unknown (Fig. 4.5; JMP v 13, SAS Institute Inc).

Table 4.3. Parameters employed for Mascot searches of proteomic data. Reverse-phase high pH liquid chromatography fractionated samples (RP-HPLC) were run on a Thermo Scientific Orbitrap Fusion Lumos Tribrid Mass Spectrometer. Quantitative samples were run on a Thermo Scientific NanoLC-Q Exactive Plus Hybrid Quadrupole-Orbitrap Mass Spectrometer. Abbreviations are as follows: CNBr – cyanogen bromide; CAM – carbamidomethyl; MetOx – methionine oxidation; N-acetyl – N-terminus protein acetylation; N- pyroGlu – N-terminus pyroglutamic acid; Hse – Homoserine; Hsl – Homoserine lactone.

Parameter RP-HPLC, Endolymph RP-HPLC, Quantitative, Quantitative, otolith Endolymph Otolith

Missed cleavages 2 2 2 2

Fixed CAM CAM CAM CAM modifications

Variable MetOx, N-acetyl, N- MetOx, N- MetOx, N-acetyl, MetOx, N-acetyl, modifications pyroGlu acetyl, N- N-pyroGlu, Hse, N-pyroGlu, Hse, pyroGlu, Hse, Hsl Hsl Hsl

Precursor mass 20 ppm 20 ppm 20 ppm 20 ppm tolerance

Fragment ion 0.3 Da 0.3 Da 20 mmu 20 mmu tolerance

Database used cDNA cDNA Custom Custom

Mascot score 13 13 13 13 threshold

E-value 0.01 0.01 0.01 0.01 threshold

170

Data deposition

The mass spectrometry proteomics data have been deposited to the ProteomeXchange

Consortium via the PRIDE partner repository with the dataset identifier PXD011021 and

10.6019/PXD011021. Transcript data have been deposited in NCBI BioSample Database with the accession numbers SAMN10272979, SAMN10272980, SAMN10272981.

ACKNOWLEDGEMENTS

We thank G. Jenkins, M. Sievers, K. Hassell, J. French, D. Chamberlain, and L. Barrett for assistance with sample collection and S. Mukherjee for assistance with laboratory work. We thank the Plant

Genome & Molecular Evolution Lab, Department of Life Science, Tunghai University, Taiwan for use of their supercomputer, and Steven Wilcox at the Systems Biology Division at the Walter and

Eliza Hall Institute of Medical Research for advice and sequencing of cDNA. We similarly thank

Karin Limburg for her advice and comments. Funding was provided by the Australian Research

Council (project LP140100087) and the University of Melbourne (John and Allan Gilmour Research

Award). ORBT was supported by an Australian Postgraduate Award.

Author contributions: ORBT – Conceptualization, Formal Analysis, Investigation, Methodology,

Visualization, Writing – original draft, review and editing; SES – Conceptualization, Formal analysis,

Funding acquisition, Methodology, Supervision, Validation, Writing – review and editing; EAK –

Data curation, Formal analysis, Methodology, Software; PP – Data curation, Formal Analysis,

Methodology, Resources, Software; GQT – Data curation, Formal analysis, Software; AP – Data curation, Formal analysis, Methodology, Resources, Software; AR – Investigation, Resources; PB –

Investigation, Resources; BRR – Conceptualization, Formal analysis, Funding acquisition,

Methodology, Project administration, Supervision, Validation, Writing – review and editing

171

REFERENCES

Bajoghli, B., Ramialison, M., Aghaallaei, N., Czerny, T., Wittbrodt, J. (2009) Identification of

Starmaker-like in medaka as a putative target gene of Pax2 in the otic vesicle.

Developmental Dynamics 238, 2860-2866.

Blasiole, B., Canfield, V.A., Vollrath, M.A., et al. (2006) Separate Na,K-ATPase genes are required

for otolith formation and semicircular canal development in zebrafish. Developmental

Biology 294, 148-160.

Borelli, G., Guibbohni, M.E., Mayer-Gostan, N., et al. (2003) Daily variations of endolymph

composition: relationship with the otolith calcification process in trout. Journal of

Experimental Biology 206, 2685-2692.

Borelli, G., Mayer-Gostan, N., De Pontual, H., Boeuf, G., Payan, P. (2001) Biochemical relationships

between endolymph and otolith matrix in the trout (Oncorhynchus mykiss) and turbot

(Psetta maxima). Calcified Tissue International 69, 356-364.

Campana, S.E. (1999) Chemistry and composition of fish otoliths: pathways, mechanisms and

applications. Marine Ecology Progress Series 188, 263-297.

Castillo, L.M., Guerrero, C.A., Acosta, O. (2017) Expression of typical osteoclast markers by PBMCs

after PEG-induced fusion as a model for studying osteoclast differentiation. Journal of

Molecular Histology 48, 169-185.

Charone, S., De Lima Leite, A., Peres-Buzalaf, C., et al. (2016) Proteomics of secretory-stage and

maturation-stage enamel of genetically distinct mice. Caries Research 50, 24-31. [In

English].

172

Cruz, S., Shiao, J.-C., Liao, B.-K., Huang, C.-J., Hwang, P.-P. (2009) Plasma membrane calcium

ATPase required for semicircular canal formation and otolith growth in the zebrafish inner

ear. Journal of Experimental Biology 212, 639-647.

Davis, J.G., Oberholtzer, J.C., Burns, F.R., Greene, M.I. (1995) Molecular cloning and

characterization of an inner ear-specific structural protein. Science 267, 1031-1034. de Vries, M.C., Gillanders, B.M., Elsdon, T.S. (2005) Facilitation of barium uptake into fish otoliths:

Influence of strontium concentration and salinity. Geochimica et Cosmochimica Acta 69,

4061-4072.

Deans, M.R., Peterson, J.M., Wong, G.W. (2010) Mammalian Otolin: A multimeric glycoprotein

specific to the inner ear that interacts with otoconial matrix protein Otoconin-90 and

Cerebellin-1. PloS One 5.

Degens, E.T., Deuser, W.G., Haedrich, R.L. (1969) Molecular structure and composition of fish

otoliths. Marine Biology 2, 105-113.

Disspain, M., Wallis, L.A., Gillanders, B.M. (2011) Developing baseline data to understand

environmental change: a geochemical study of archaeological otoliths from the Coorong,

South Australia. Journal of Archaeological Science 38, 1842-1857.

Drake, F.H., Dodds, R.A., James, I.E., et al. (1996) Cathepsin K, but not cathepsins B, L, or S, is

abundantly expressed in human osteoclasts. Journal of Biological Chemistry 271, 12511-

12516. [In English].

Drenckhahn, D., Schluter, K., Allen, D., Bennett, V. (1985) Colocalization of band 3 with ankyrin and

spectrin at the basal membrane of intercalated cells in the rat kidney. Science 230, 1287-

1289. [In English].

173

Elsdon, T.S., Gillanders, B.M. (2003) Relationship between water and otolith elemental

concentrations in juvenile black bream Acanthopagrus butcheri. Marine Ecology Progress

Series 260, 263-272.

Elsdon, T.S., Gillanders, B.M. (2004) Fish otolith chemistry influenced by exposure to multiple

environmental variables. Journal of Experimental Marine Biology and Ecology 313, 269-

284.

Elsdon, T.S., Gillanders, B.M. (2005) Alternative life-history patterns of estuarine fish: barium in

otoliths elucidates freshwater residency. Canadian Journal of Fisheries and Aquatic

Sciences 62, 1143-1152.

Gauldie, R.W. (1986) Vaterite otoliths from chinook salmon (Oncorhynchus tshawytscha). New

Zealand Journal of Marine and Freshwater Research 20, 209-217. [In English].

Gilar, M., Olivova, P., Daly, A.E., Gebler, J.C. (2005) Two-dimensional separation of peptides using

RP-RP-HPLC system with different pH in first and second separation dimensions. Journal of

Separation Science 28, 1694-1703.

Hojo, M., Omi, A., Hamanaka, G., et al. (2015) Unexpected link between polyketide synthase and

calcium carbonate biomineralization. Zoological Letters 1, 16. [In English].

Holubowicz, R., Wojtas, M., Taube, M., Kozak, M., Ozyhar, A., Dobryszycki, P. (2017) Effect of

calcium ions on structure and stability of the C1q-like domain of Otolin-1 from human and

zebrafish. FEBS Journal 284, 4278-4297.

Iozzo, R.V. (1998) Matrix proteoglycans: From molecular design to cellular function. Annual Review

of Biochemistry 67, 609-652. [In English].

174

Izzo, C., Doubleday, Z.A., Grammer, G.L., et al. (2016) Fish as proxies of ecological and

environmental change. Reviews in Fish Biology and Fisheries 26, 265-286. [In English].

Jeong, J.A., Ko, K.M., Bae, S., Jeon, C.J., Koh, G.Y., Kim, H. (2007) Genome-wide differential gene

expression profiling of human bone marrow stromal cells. Stem Cells 25, 994-1002. [In

English].

Kall, L., Storey, J.D., MacCoss, M.J., Noble, W.S. (2008) Posterior error probabilities and false

discovery rates: two sides of the same coin. Journal of Proteome Research 7, 40-44. [In

English].

Kang, Y.J., Stevenson, A.K., Yau, P.M., Kollmar, R. (2008) Sparc protein is required for normal

growth of zebrafish otoliths. J Assoc Res Otolaryngol 9, 436-451.

Kaplon, T.M., Rymarczyk, G., Nocula-Lugowska, M., et al. (2008) Starmaker exhibits properties of

an intrinsically disordered protein. Biomacromolecules 9, 2118-2125.

Kohfeldt, E., Sasaki, T., Gohring, W., Timpl, R. (1998) Nidogen-2: a new basement membrane

protein with diverse binding properties. Journal of Molecular Biology 282, 99-109.

Lambert, L.A. (2012) Molecular evolution of the transferrin family and associated receptors.

Biochimica Et Biophysica Acta-General Subjects 1820, 244-255.

Le Norcy, E., Lesieur, J., Sadoine, J., Rochefort, G.Y., Chaussain, C., Poliard, A. (2018)

Phosphorylated and non-phosphorylated leucine rich amelogenin peptide differentially

affect ameloblast mineralization. Frontiers in Physiology 9, 10. [In English].

Lima Leite, A., Silva Fernandes, M., Charone, S., Whitford, G.M., Everett, E.T., Buzalaf, M.A.R.

(2018) Proteomic mapping of dental enamel matrix from inbred mouse strains: Unraveling

potential new players in enamel. Caries Research 52, 78-87. [In English].

175

Liu, J., Li, X., Zhang, H., et al. (2017) Ubiquitin E3 ligase Itch negatively regulates osteoblast

function by promoting proteasome degradation of osteogenic proteins. Bone & Joint

Research 6, 154-161.

Lowenstein, O. (1936) The equilibrium function of the vertebrate labyrinth. Biological Reviews of

the Cambridge Philosophical Society 11, 113-145.

Mazzieri, R., Masiero, L., Zanetta, L., et al. (1997) Control of type IV collagenase activity by

components of the urokinase-plasmin system: A regulatory mechanism with cell-bound

reactants. EMBO Journal 16, 2319-2332. [In English].

Morison, A.K., Coutin, P.C., Robertson, S.G. (1998) Age determination of black bream,

Acanthopagrus butcheri (Sparidae), from the Gippsland Lakes of south-eastern Australia

indicates slow growth and episodic recruitment. Marine and Freshwater Research 49, 491-

498. [In English].

Mugiya, Y. (1987) Phase difference between calcification and organic matrix formation in the

diurnal growth of otoliths in the rainbow-trout, Salmo gairdneri. Fishery Bulletin 85, 395-

401. [In English].

Mugiya, Y., Uchimura, T. (1989) Otolith resorption induced by anaerobic stress in the goldfish,

Carassius auratus. Journal of Fish Biology 35, 813-818. [In English].

Mugiya, Y., Yoshida, M. (1995) Effects of calcium antagonists and other metabolic modulators on

in vitro calcium deposition on otoliths in the rainbow trout Oncorhynchus mykiss. Fisheries

Science 61, 1026-1030. [In English].

176

Murayama, E., Herbomel, P., Kawakami, A., Takeda, H., Nagasawa, H. (2005) Otolith matrix

proteins OMP-1 and Otolin-1 are necessary for normal otolith growth and their correct

anchoring onto the sensory maculae. Mechanisms of Development 122, 791-803.

Murayama, E., Okuno, A., Ohira, T., Takagi, Y., Nagasawa, H. (2000) Molecular cloning and

expression of an otolith matrix protein cDNA from the rainbow trout, Oncorhynchus

mykiss. Comparative Biochemistry and Physiology B-Biochemistry & Molecular Biology 126,

511-520.

Murayama, E., Takagi, Y., Nagasawa, H. (2004) Immunohistochemical localization of two otolith

matrix proteins in the otolith and inner ear of the rainbow trout, Oncorhynchus mykiss:

comparative aspects between the adult inner ear and embryonic otocysts. Histochemistry

and Cell Biology 121, 155-166.

Murayama, E., Takagi, Y., Ohira, T., Davis, J.G., Greene, M.I., Nagasawa, H. (2002) Fish otolith

contains a unique structural protein, Otolin-1. European Journal of Biochemistry 269, 688-

696.

Panella, G. (1971) Fish otoliths: daily growth layers and periodical patterns. Science (New York,

N.Y.) 173, 1124-1127.

Partridge, G.J., Jenkins, G.I., Doupe, R.G., De Lestang, S., Ginbey, B.M., French, D. (2009) Factors

affecting mark quality of alizarin complexone-stained otoliths in juvenile black bream

Acanthopagrus butcheri and a prescription for dosage. Journal of Fish Biology 75, 1518-

1523.

Payan, P., De Pontual, H., Boeuf, G., Mayer-Gostan, N. (2004) Endolymph chemistry and otolith

growth in fish. Comptes Rendus Palevol 3, 535-547.

177

Petko, J.A., Millimaki, B.B., Canfield, V.A., Riley, B.B., Levenson, R. (2008) Otoc1: A novel Otoconin-

90 ortholog required for otolith mineralization in zebrafish. Developmental Neurobiology

68, 209-222.

Poznar, M., Hołubowicz, R., Wojtas, M., et al. (2017) Structural properties of the intrinsically

disordered, multiple calcium ion-binding Otolith Matrix Macromolecule-64 (OMM-64).

Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics 1865, 1358-1371.

Ranaldi, M.M., Gagnon, M.M. (2008) Trace metal incorporation in otoliths of black bream

(Acanthopagrus butcheri munro), an indicator of exposure to metal contamination. Water

Air and Soil Pollution 194, 31-43.

Reimer, T., Dempster, T., Wargelius, A., et al. (2017) Rapid growth causes abnormal vaterite

formation in farmed fish otoliths. Journal of Experimental Biology 220, 2965-2969. [In

English].

Renn, J., Winkler, C. (2014) Osterix/Sp7 regulates biomineralization of otoliths and bone in medaka

(Oryzias latipes). Matrix Biology 34, 193-204.

Rozycka, M., Wojtas, M., Jakob, M., et al. (2014) Intrinsically disordered and pliable Starmaker-like

protein from medaka (Oryzias latipes) controls the formation of calcium carbonate

crystals. PloS One 9, e114308.

Sakabe, R., Lyle, J.M., Crawford, C.M. (2011) The influence of freshwater inflows on spawning

success and early growth of an estuarine resident fish species, Acanthopagrus butcheri.

Journal of Fish Biology 78, 1529-1544.

Salmon, C.R., Giorgetti, A.P.O., Leme, A.F.P., et al. (2017) Microproteome of dentoalveolar tissues.

Bone 101, 219-229. [In English].

178

Sarre, G.A., Potter, I.C. (2000) Variation in age compositions and growth rates of Acanthopagrus

butcheri (Sparidae) among estuaries: some possible contributing factors. Fishery Bulletin

98, 785-799.

Shima, J.S., Swearer, S.E. (2009) Larval quality is shaped by matrix effects: implications for

connectivity in a marine metapopulation. Ecology 90, 1255-1267.

Siebers, M.C., ter Brugge, P.J., Walboomers, X.F., Jansen, J.A. (2005) Integrins as linker proteins

between osteoblasts and bone replacing materials. A critical review. Biomaterials 26, 137-

146. [In English].

Söllner, C., Burghammer, M., Busch-Nentwich, E., et al. (2003) Control of crystal size and lattice

formation by Starmaker in otolith biomineralization. Science 302, 282-286.

Su, X., Chen, D., Yang, K., et al. (2017) The circadian clock gene PER2 plays an important role in

tumor suppression through regulating tumor-associated genes in human oral squamous

cell carcinoma. Oncology Reports 38, 472-480.

Sumanas, S., Larson, J.D., Bever, M.M. (2003) Zebrafish chaperone protein GP96 is required for

otolith formation during ear development. Developmental Biology 261, 443-455.

Tagliabracci, V.S., Engel, J.L., Wen, J.Z., et al. (2012) Secreted kinase phosphorylates extracellular

proteins that regulate biomineralization. Science 336, 1150-1153. [In English].

Takagi, Y., Ishida, K., Mugiya, Y. (2000) Carbohydrates of the otolith organ in the rainbow trout

Oncorhynchus mykiss detected by lectins. Fisheries Science 66, 933-939.

Takagi, Y., Tohse, H., Murayama, E., Ohira, T., Nagasawa, H. (2005) Diel changes in endolymph

aragonite saturation rate and mRNA expression of otolith matrix proteins in the trout

otolith organ. Marine Ecology Progress Series 294, 249-256.

179

Thomas, O.R.B., Ganio, K., Roberts, B.R., Swearer, S.E. (2017) Trace element-protein interactions in

endolymph from the inner ear of fish: implications for environmental reconstructions

using fish otolith chemistry. Metallomics 9, 239-249.

Tohse, H., Ando, H., Mugiya, Y. (2004) Biochemical properties and immunohistochemical

localization of carbonic anhydrase in the sacculus of the inner ear in the salmon

Oncorhynchus masou. Comparative Biochemistry and Physiology a-Molecular & Integrative

Physiology 137, 87-94.

Tohse, H., Mugiya, Y. (2001) Effects of enzyme and anion transport inhibitors on in vitro

incorporation of inorganic carbon and calcium into endolymph and otoliths in salmon

Oncorhynchus masou. Comparative Biochemistry and Physiology a-Molecular and

Integrative Physiology 128, 177-184.

Tohse, H., Murayama, E., Ohira, T., Takagi, Y., Nagasawa, H. (2006) Localization and diurnal

variations of carbonic anhydrase rnRNA expression in the inner ear of the rainbow trout

Oncorhynchus mykiss. Comparative Biochemistry and Physiology B-Biochemistry &

Molecular Biology 145, 257-264.

Tohse, H., Saruwatari, K., Kogure, T., Nagasawa, H., Takagi, Y. (2009) Control of polymorphism and

morphology of calcium carbonate crystals by a matrix protein aggregate in fish otoliths.

Crystal Growth & Design 9, 4897-4901.

Tohse, H., Takagi, Y., Nagasawa, H. (2008) Identification of a novel matrix protein contained in a

protein aggregate associated with collagen in fish otoliths. FEBS Journal 275, 2512-2523.

Uversky, V.N., Dunker, A.K. (2010) Understanding protein non-folding. Biochimica Et Biophysica

Acta-Proteins and Proteomics 1804, 1231-1264.

180

Webb, S.D., Woodcock, S.H., Gillanders, B.M. (2012) Sources of otolith barium and strontium in

estuarine fish and the influence of salinity and temperature. Marine Ecology Progress

Series 453, 189-199.

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2016) Not all inner ears are the same: Otolith matrix

proteins in the inner ear of sub-adult cichlid fish, Oreochromis mossambicus, reveal

insights into the biomineralization process. Anatomical Record-Advances in Integrative

Anatomy and Evolutionary Biology 299, 234-245.

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2017) Formation of the inner ear during embryonic

and larval development of the cichlid fish (Oreochromis mossambicus). Connective Tissue

Research 58, 172-195.

Wiedemann-Bidlack, F.B., Kwak, S.Y., Beniash, E., Yamakoshi, Y., Simmer, J.P., Margolis, H.C. (2011)

Effects of phosphorylation on the self-assembly of native full-length porcine amelogenin

and its regulation of calcium phosphate formation in vitro. Journal of Structural Biology

173, 250-260. [In English].

Williams, J., Jenkins, G.P., Hindell, J.S., Swearer, S.E. (2018) Fine-scale variability in elemental

composition of estuarine water and otoliths: Developing environmental markers for

determining larval fish dispersal histories within estuaries. Limnology and Oceanography

63, 262-277. [In English].

Wojtas, M., Wolcyrz, M., Ozyhar, A., Dobryszycki, P. (2012) Phosphorylation of intrinsically

disordered Starmaker protein increases its ability to control the formation of calcium

carbonate crystals. Crystal Growth & Design 12, 158-168.

181

Xiao, Z., Camalier, C.E., Nagashima, K., et al. (2007) Analysis of the extracellular matrix vesicle

proteome in mineralizing osteoblasts. Journal of Cellular Physiology 210, 325-335. [In

English].

Ye, Y.P., Hu, W.H., Guo, F.J., Zhang, W.K., Wang, J., Chen, A.M. (2012) Glycosaminoglycan chains of

Biglycan promote Bone Morphogenetic Protein-4-induced osteoblast differentiation.

International Journal of Molecular Medicine 30, 1075-1080. [In English].

Yin, K., Paine, M.L. (2017) Bicarbonate transport during enamel maturation. Calcified Tissue

International 101, 457-464.

Zhang, Z., Runham, N.W. (1992) Temporal deposition of incremental and discontinuous zones in

the otoliths of Oreochromis niloticus (L.). Journal of Fish Biology 40, 333-339.

Zhen, E.Y., Brittain, I.J., Laska, D.A., et al. (2008) Characterization of metalloprotease cleavage

products of human articular cartilage. Arthritis and Rheumatism 58, 2420-2431. [In

English].

182

CHAPTER FIVE: SPATIO-TEMPORAL RESOLUTION OF SPAWNING AND LARVAL NURSERY

HABITATS USING OTOLITH MICROCHEMISTRY IS ELEMENT DEPENDENT

ABSTRACT

Otolith chemistry is frequently employed in the reconstruction of fish environmental histories.

While some elements have been strongly correlated with environmental physico-chemical factors

(e.g. salinity, temperature, water chemistry), others may not indicate exogeny and simply add endogenous variability to a data set. In a previous study I identified several commonly-assessed elements as being only present in the proteinaceous fraction of endolymph from black bream

(Acanthopagrus butcheri), suggesting that the choice of elements in otolith multi-elemental fingerprinting could influence their utility as a natural environmental marker. To test this hypothesis, I performed several cluster analyses based on different sets of trace element data extracted from both the cores and larval region of otoliths of juvenile black bream caught in the

Gippsland Lakes, Victoria, Australia, in 2015 and 2016. I used Ward’s method for hierarchical clustering, and clustered in three different ways: (a) all elements analysed; (b) elements identified as being primarily in the salt-fraction of the endolymph; and (c) elements identified as being associated with the protein-fraction of the endolymph. I then assessed the power of the resulting clusters to resolve temporal (among cohort) and/or spatial (collection location) differences in otolith chemistry using multinomial logistic regression. My results suggest that clustering based solely on endolymph salt-fraction elements is best for detecting the spawning sources and larval nursery habitats of black bream in the Gippsland Lakes. The results were consistent with known spawning locations but I also found evidence for a distinct and temporally consistent multi- elemental signature, suggesting an unknown source of black bream larvae. Identifying the novel

183

spawning location and assessing its importance to fisheries productivity in the Lakes requires further investigation.

INTRODUCTION

Trace element fingerprints from fish otoliths are routinely used in the reconstruction of environmental histories via laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-

MS) (FitzGerald et al., 2004, Thorrold and Swearer, 2009, Izzo et al., 2016b). Otoliths – biomineralized “earstones” in the inner ear labyrinth of all bony fish that are formed in embryo – are immersed in inner ear fluid (endolymph), and grow by depositing a protein-rich band and a calcium carbonate (aragonite) rich band each day (Pannella, 1971, Dunkelberger et al., 1980). In addition to this, they have been found to be metabolically inert, only undergoing resorption under extreme anaerobic conditions (Mugiya and Uchimura, 1989). Consequently, the otolith is considered to be a reliable physical record of endolymph change, as trace elements are typically incorporated during the process of increment formation. Therefore, trace-element:calcium ratios in otoliths can be seen as an indication of water chemistry over time, allowing for spatio-temporal reconstructions of an individual’s life history (Radtke, 1982, Secor et al., 1995, Dorval et al., 2007,

Elsdon et al., 2008). Otolith chemistry has been shown to aid in identifying movement patterns

(Kalish, 1990, Swearer et al., 1999, Crook et al., 2006), habitat use (Thorrold et al., 1998b, Beer et al., 2011) and natal sources (Thorrold et al., 2001, Barbee and Swearer, 2007). The methodology is particularly useful in estuarine systems, where variability in freshwater flow results in dynamic estuarine water chemistry (Gillanders and Kingsford, 2002).

Campana (1999) suggested that there were four main routes for trace element incorporation into the otolith: (a) direct substitution for calcium, as has been shown for Sr (Doubleday et al., 2014);

184

(b) formation of crystal impurities in the aragonite lattice; (c) trapping in the interstitial spaces of the crystal, as is likely with elements of small atomic radii (Thomas et al., 2017); and (d) interaction with otolith proteins (Campana, 1999, Izzo et al., 2016a, Thomas et al., 2017). Most otolith chemistry studies have focused on elements where incorporation is likely one of the first three

(e.g., Li, Mg, Mn, Sr, and Ba) (Limburg, 1995, Bath et al., 2000, Secor et al., 2001, Gillanders, 2002).

However, transition metals (e.g. Zn, Cu) and non-metals (e.g. P, S), which are strongly associated with proteins, are often also used in trace element fingerprinting (Swearer et al., 2003, Patterson et al., 2004, Vasconcelos et al., 2007, Tanner et al., 2012, Reis-Santos et al., 2013). This raises the question of how useful they are as indicators of environmental history, given the former are prime candidates to be protein co-factors and the latter are typically found in protein post-translational modifications (e.g. phosphorylation – a post-translational modification by which phosphate ions are used to “decorate” a protein). Furthermore, transition metals have been found to be bound primarily to proteins in the inner ear (Miller et al., 2006, Izzo et al., 2016a, Thomas et al., 2017).

Although not all transition metals are typical protein co-factors, they can be bound to a protein either through misincorporation (Kang et al., 2011), substituting for the normal co-factor, or by adhering to the protein surface by electrostatic attraction (Lothian et al., 2018). Consequently, it may be that these elements are more indicative of endogenous rather than exogenous processes, as has been suggested by Sturrock et al. (2015).

To assess how the suite of elements chosen influences the utility of fish otolith fingerprinting for reconstructing environmental histories, I conducted a study on black bream (Acanthopagrus butcheri) in the Gippsland Lakes, Australia, which supports the largest commercial fishery for this species in the State of Victoria (Nicholson et al., 2008).

Black bream is a member of the family Sparidae and is both commercially and recreationally important. It has been well characterised in terms of its otolith chemistry (Elsdon and Gillanders,

185

2002, Elsdon and Gillanders, 2003, Elsdon and Gillanders, 2004, de Vries et al., 2005, Elsdon and

Gillanders, 2005a, Thomas et al., 2017, Williams et al., 2018). It is a relatively long-lived estuarine- dependent species, with individuals tending to spend the entirety of their lives within their natal estuary (Butcher et al., 1962). Typical of many long-lived species, black bream populations are dominated by only a few year classes, as a result of large variations in annual recruitment (Jenkins et al., 2010, Jenkins et al., 2015).

Adult bream spawn serially, with spawning typically occurring daily throughout spring and early summer (Haddy and Pankhurst, 1998, Sakabe and Lyle, 2010). They have, however, been shown to vary in the timing of spawning between locations (Sarre and Potter, 1999). Jenkins et al. (2018) found evidence of spawning occurring in late summer/early autumn for juveniles sampled in the

Gippsland Lakes over February to April 2016. In periods of low freshwater flow, black bream have been found to move into the main Gippsland Lakes tributaries (the Mitchell, Nicholson and Tambo

Rivers) prior to spawning, where they remain well into the post-spawning season (Hindell et al.,

2008). Additionally, Williams et al. (2012) found that although adult bream disperse into the Lakes during flood events, they return back to their original tributary. Black bream produce pelagic eggs, which are generally found in the upper reaches of the estuary (Newton, 1996, Nicholson et al.,

2008) at salinities greater than 10 PSU (Jenkins et al., 2010). In a study conducted by Williams et al.

(2012), it was shown that nearly all (99%) of black bream larvae sampled were found in the riverine tributaries, with the majority being found in the Mitchell River.

As larvae grow older, it is thought that they are dispersed from the tributary in which they were spawned down to the Lakes proper, where they then settle to nursery habitats, particularly seagrass (Zostera and Ruppia spp.) (Rigby, 1984). Although larval dispersal has been traditionally thought to be passive, and as the result of water currents, several papers over the past two decades have suggested behaviour is a dominant driver (Leis et al., 1996, Kingsford et al., 2002,

186

Leis et al., 2003, Levin, 2006, Irisson et al., 2015). Nevertheless, in a group of studies on bream in the Mitchell River, Gippsland Lakes, it was found that despite significant swimming ability being present amongst larvae, this ability had little effect on population dispersal, and that the hydrodynamics of the estuarine system were a more important driver (Gee, 2014). In that study, it was postulated that interactions between flow and longitudinal spawning locations, as well as degree of stratification, were important factors to be considered. In addition to this, freshwater flow has been shown to influence survival of black bream at early life stages (Haddy and

Pankhurst, 2000, Hassell et al., 2008), as well as account for intra-estuarine variability in larval dispersal (Gillanders and Kingsford, 2002).

Australian estuarine systems have a unique geomorphology, where the interannual variation in discharge is far greater than that experienced by more studied Northern hemisphere systems, as well as having considerable differences in summer and winter discharge volumes (Kench, 1999). In recent years, in studies by both Jenkins et al. (2018) and Williams et al. (2018), the three main tributaries of the Gippsland Lakes were found to be comparable in terms of stratification as well as showing consistently low levels of dissolved oxygen at depths below the halocline region. The relative contribution of each of these three tributaries to black bream year class strength is currently unknown. Otolith chemistry has been shown to be indicative of spatio-temporal changes in ambient water conditions (Elsdon et al., 2008). Therefore, the specific trace element:calcium ratios in the region of the otolith associated with the larval life stage (i.e. at or around the core), may allow for the identification of natal sources (Thorrold et al., 2001) and reconstruction of dispersal patterns (Swearer et al., 1999) for black bream in the Gippsland Lakes.

A range of elements are typically used in LA-ICP-MS-based reconstructions of dispersal histories in estuaries, notably Sr and Ba. Sr and Ba are both Group 2 elements and are thought to substitute directly for calcium in the aragonite lattice of otolith increments (Doubleday et al., 2014, Thomas

187

et al., 2017). Levels of Sr and Ba in the otolith have both been shown to positively correlate with water chemistry in both the laboratory (Bath et al., 2000, Kraus and Secor, 2004, Elsdon and

Gillanders, 2005b, Walther and Thorrold, 2006, Hicks et al., 2010, Miller, 2011) and the field

(Thorrold et al., 1997, Dorval et al., 2007, Sturrock et al., 2012). However, it is thought that by including additional trace elements in analyses a finer separation of dispersal histories can be achieved (Hicks et al., 2010, Tanner et al., 2013).

To assess the utility of different multi-elemental data sets for environmental reconstructions, I collected juvenile bream in vegetated habitats (seagrass, macroalgae), for each month from

February-April in 2015 and 2016. I used LA-ICP-MS to collect larval period data in the form of 12 trace elements and separately clustered the data in three different ways, using my previous studies as a guideline (Thomas et al., 2017, Thomas et al., 2019). I then assessed each of the approaches in detecting variation in elemental signatures among cohorts and collection locations using multinomial logistic regression and used the results to infer the spawning locations and larval nursery environments for black bream in the Gippsland Lakes.

MATERIALS AND METHODS

Study Area

The Gippsland Lakes are located in Victoria, southeastern Australia. They form the largest estuarine lagoon system in the country, with an approximate area of 600 km2. Having a reasonably low tidal range (<30 cm), they are connected by an artificial channel to the open ocean at Lakes

Entrance. It is understood that the Lakes formed through the deposition of successive sandy barrier formations across a wide marine coastal embayment during the Quaternary period (Bird,

1965). The largest lake (and focus of this study), Lake King, is fed by three principle waterways –

188

the Mitchell, Nicholson and Tambo Rivers, all of which have been identified as potential spawning areas for Black bream (Williams et al., 2012). These three rivers are all reasonably close together, being within 5 km of each other. The benthos of Lake King is rich in seagrass and macroalgae.

Collection of fish

Juvenile black bream were sampled at several sites in the Gippsland Lakes estuary (but chiefly throughout Lake King) for each month of February-April in 2015 and 2016 (Fig. 5.1). At each site, sampling was conducted in both vegetated (containing macroalgae and/or Zostera spp., Ruppia spp.) and unvegetated areas. A small seine was employed (15 m long, 2 m deep, 2 mm knotless mesh) by two operators employing a pursing technique (Jenkins and Sutherland, 1997). During sampling, the seine was walked out by one operator to encircle the vegetated or bare area. The diameter of the area encapsulated ranged from 3.5 – 8 m with a water depth of 0.3 – 1.5 m. All fish were sorted, identified, counted and measured (total length, mm). Juvenile black bream of

≤160mm were placed on ice and then transported to the University of Melbourne where they were frozen at -20 °C. All juveniles were assigned a unique identification number. Fish were collected under Monash University Ethics Committee approval BSCI/2014/19 and a Department of

Environment and Primary Industries Fisheries research permit #1204.

189

Figure 5.1. Fishing sites for sampling black bream (Acanthopagrus butcheri), Gippsland Lakes estuary; Feb-Apr, 2015 and 2016.

190

Otolith preparation

Black bream juveniles were thawed, decapitated, and the cranium was cut open by making a transverse incision from the snout above the eyes to above the gill arches. The brain was removed and set aside, exposing the saccula. Using fine forceps, the left and right sagittal otoliths were gently removed, rinsed in ultra-pure Milli-Q water (18.2 MΩ; Merck Millipore, Australia), air-dried, and placed into a labelled 1.5 mL microcentrifuge tube (TechnoPlas, Australia).

Otoliths were then immersed in 500 µL of a buffered cleaning solution. The solution was made from 30% v/v ultra-pure H2O2 (VLSI Selectipur, BASF) and a solution of 0.1 M NaOH (NaOH pellets,

99.99%, semiconductor grade, Sigma Aldrich, Castle Hill, Australia in ultra-pure Milli-Q water, 18.2

MΩ; Merck Millipore, Australia) to give a final concentration of 15% v/v H2O2/0.05 M NaOH.

Following this, tubes underwent ultra-sonication for 5 min (Sonic Clean 250HT) before being left to sit for 12 h in the cleaning solution. The cleaning solution was replenished, and the samples resonicated for 5 min and allowed to rest for a further 12 h. After this, the cleaning solution was removed by aspiration and the otoliths rinsed three times in Milli-Q water, with each rinse accompanied by 5 min of ultrasonication and 30 min of resting. Finally, otoliths were air-dried in a

Class 100 laminar flow bench for 24 h.

Dried otoliths were individually transferred with plastic forceps into a drop of epoxy resin (EpoThin

2, Buehler, USA) on an 18 mm acrylic disc and allowed to set. Mounted otoliths were ground to 5 –

15 µm from their cores with 9 µm diamond lapping film (668X, 3M, USA). Following grinding, all mounted samples were transferred to 12-well polystyrene microplates (ThermoFisher, Australia), and immersed in buffered cleaning solution for 24 h, with one change of cleaning solution as described above. Similarly, all samples underwent the same rinsing/ultrasonication cycles before

191

being dried in a Class 100 laminar flow bench for 24 h. All containers, bottles and well trays were previously acid leached for 48 h in a solution of 1.0 M HNO3 (Sigma-Aldrich, Australia) before use.

LA-ICP-MS

Multi-elemental signatures of otolith cores (both the primordium and the early larval period) were determined by vertical profiling using LA-ICP-MS. I previously demonstrated the partitioning of elements in the endolymph in the inner ear of black bream, identifying those that are present in the salt fraction, and therefore likely to be indicative of exogeny, and those present in the proteinaceous fraction, and therefore likely indicative of endogeny (Thomas et al., 2017).

Consequently, I chose a range of elements for my analyses: salt fraction-dominant (7Li, 11B, 24Mg,

39K, 88Sr, 138Ba) and protein fraction-dominant (31P, 34S, 63Cu, 66Zn, 208Pb). I also previously found the presence of a key biomineralization protein, Extracellular serine/threonine protein kinase FAM20C

(FAM20C) in black bream otoliths (Thomas et al., 2019). FAM20C is a kinase, meaning that it modifies proteins by adding phosphate groups to them. In addition to this, it has Mn as a co-factor in its tertiary structure (Tagliabracci et al., 2012). The primordia of otoliths have been shown to be enriched in both P and Mn (Brophy et al., 2004, Ruttenberg et al., 2005, Shima and Swearer,

2009). Due to this, I also included 31P and 55Mn in my analyses as indicators of the primordium (the centre of the otolith, protein-rich and surrounded by the first calcified daily increment). Finally,

43Ca was used as an internal standard. All elements analysed were expressed as a ratio to 43Ca.

I employed a 7700x ICP-MS (Agilent, USA) fitted with a RESOlution LA system (Resonetics, Nashua

NH, USA). This was attached to a Compex 110 (Lambda Physik) excimer laser operating at 193 nm.

The system was calibrated for each element using one of two glass standards (NIST 610 and 612) doped with trace elements at known concentrations (National Institute of Standards and

192

Technology). Otoliths were analysed in blocks of 12, bracketed by analyses of the NIST standards and a calcium carbonate consistency standard (MACS3, United States Geological Survey).

Estimates of external precision (%RSD) based on 24 analyses of the MACS3 were as follows: 7Li,

7.11; 11B, 12.79; 24Mg, 9.11; 31P, 28.43; 34S, 20.97; 39K, 50.92; 55Mn, 5.11; 63Cu, 8.99; 66Zn, 14.09;

88Sr, 6.98; 138Ba, 6.42; 208Pb, 6.65. Because the MACS3 standard is not homogeneous (and contains low concentrations of some elements such as 39K), I also provide the following estimates of external precision (%RSD) based on 24 analyses of NIST612 (calibrated against 610): 7Li, 2.10; 11B,

3.24; 24Mg, 2.30; 31P, 36.71; 34S, 13.80; 39K, 7.92; 55Mn, 1.70; 63Cu, 2.26; 66Zn, 2.27; 88Sr, 1.37; 138Ba,

1.62; 208Pb, 1.80. Please note that %RSD listed above are for trace element ratios to the internal standard, 43Ca. Samples and standards were analysed in time-resolved mode, with a spot size of 18

µm, a laser energy density of ∼3 J/cm2 and a laser repetition rate of 5 Hz. Spot ablation was performed under pure He (400 ml/min) to minimise re-deposition of ablated material and the sample was then entrained into the Ar (1 L/min) carrier gas flow to the ICP-MS.

All 12 elements analysed were converted to either µmol (7Li, 55Mn, 63Cu, 66Zn, 138Ba, 208Pb) or mmol

(11B, 24Mg, 31P, 34S, 39K, 88Sr) per mol 43Ca. These values were averaged for each sample (2015, n=90; 2016, n=186), log-transformed (if required) and then standardized (mean of 0 and a standard deviation of 1) to give equal weighting of all elements in subsequent analyses. 7Li, 34S, 39K and 88Sr did not require log-transformation prior to standardization. Thirty second blanks were acquired before ablation of each standard and sample. I undertook a vertical profiling method for sampling, rather than horizontal transects. I previously identified and photographed the location of the core and used this as a guide to drilling (Olympus SZX7). I created an integrated signature of larval chemistry by averaging all scans. The mean number of scans per sample was 175.8 with a standard deviation of 50.9.

193

I was only able to definitively characterise the chemistry of the primordium for a subset of otoliths. Successfully “hitting” the primordium was indicated by at least an order of magnitude spike in the 55Mn:43Ca ratio accompanied by a peak in 31P:43Ca, as suggested by previous studies

(Brophy et al., 2004, Shima and Swearer, 2009). For this subset of otoliths (2015, n=38; 2016, n=96), I determined the primordium values for each element as the mean of a block of 11 scans that encompassed the peak 55Mn:43Ca value and the 5 scans preceding and following it. These average values were then standardized, as described above for the “larval period”.

Data analysis

In order to compare fish distributions across Lake King, I calculated catch per unit effort (CPUE):

푛표. 표푓 푏푙푎푐푘 푏푟푒푎푚 퐶푃푈퐸 = 휋 × (ℎ푎푢푙 푑𝑖푎푚푒푡푒푟 ÷ 2)2

I then standardized CPU to give comparative numbers of black bream for each of the fishing areas:

퐶푃푈퐸 × 푚푒푎푛(휋 × (ℎ푎푢푙 푑𝑖푎푚푒푡푒푟 ÷ 2)2)

I determined cohorts for each year by visual inspection of size-frequency distribution plots.

Estimates of juvenile black bream growth rates (Jenkins, unpublished) were used to eliminate fish that were not young-of-the-year.

Data from LA-ICP-MS was split by year, and then analysed using Ward’s Hierarchical clustering method (JMP v 13, SAS Institute Inc). I used the distance graph (which plots the distance between clusters as a function of the number of clusters) to determine the optimal number of clusters. The point where the slope of the distance graph levels off indicates a natural break in the data and was used to determine the optimal number of clusters for each method. I performed three different

194

cluster analyses per year, with different elements chosen for clustering: all elements (7Li, 11B, 24Mg,

31P, 34S, 39K, 55Mn, 88Sr, 63Cu, 66Zn, 138Ba, 208Pb); “salt-fraction dominant” (7Li, 11B, 24Mg, 39K, 88Sr,

138Ba); and, “protein fraction-dominant” (31P, 34S, 63Cu, 66Zn, 208Pb). This was done for both the

“larval period” and primordium data sets. I previously found 55Mn in the salt-fraction of adult endolymph of black bream (Thomas et al., 2017). However, as stated above, 55Mn is also a co- factor of the otolith protein FAM20C, which is likely to be in the primordium. It is likely then, that

Mn is only an indicator of exogeny when found outside the primordium. Consequently, when clustering the “larval period” data it was added to the suite of “salt-fraction dominant” elements, whereas when clustering the “primordium” data it was part of the suite of “protein-dominant” elements. I undertook a total of 12 cluster analyses (3 analyses for each of “larval period” or

“primordium” for both 2015 and 2016) (Fig. 5.2).

Figure 5.2. Schematic illustrating how LA-ICP-MS data were used as observations in the cluster analyses of juvenile black bream (Acanthopagrus butcheri) otoliths. Otoliths were extracted from juveniles caught in Lake King during the autumn months of 2015 and 2016.

195

I assessed the power of my cluster analysis methods to predict larval source and dispersal histories by performing a multinomial logistic regression (R v.3.4.3, R Core Team), treating the assigned cluster ID as a response variable with cohort and juvenile recruitment site as predictor variables (R

Core Team, 2017). I also performed MANOVA on each of the three cluster methods, followed by a

Tukey’s HSD to identify elements significantly contributing to clustering (R v.3.4.3, R Core Team).

There are some difficulties associated with using primordium chemistry as an indicator of spawning location. Firstly, it has been suggested that primordium chemistry may be influenced by maternal effects on the oocyte and thus not reflect well the site of spawning (Tzeng and Tsai,

1994, Zimmerman and Reeves, 2000). Furthermore, as the primordium is difficult to hit during laser ablation, there is inevitably a decrease in the subset’s sample size. Consequently, I performed chi-square tests to determine if there were any relationships between clusters assigned by primordium chemistry and those assigned by larval period chemistry (R v.3.4.3, R Core Team).

RESULTS

Cohort allocation

Sampled juvenile bream ranged in size from 10 to 130 mm total length in 2015 and 12 to 140 mm total length in 2016. When allocating cohorts, fish were excluded if their size indicated that they were unlikely to have been spawned in the previous Spring or Autumn. In February, for both years, all fish greater than 80 mm were excluded. In March and April, for both years, all fish greater than

100 mm were excluded. Additionally, a cohort cut-off of 40-50 mm was evident in both March and

April for 2015 and 2016. Cohort allocation is shown in Figure 5.3.

196

Figure 5.3. Size frequency distribution of juvenile black bream (Acanthopagrus butcheri) caught in Lake King, Victoria, Australia over February, March and April of 2015 (n=114) and 2016 (n=326). Two cohorts (C1 and C2) are evident each year.

Juvenile distribution, 2015

In 2015, 126 fish were caught with 87.5% of these found in vegetated patches at five juvenile recruitment sites: Broadlands, Metung West, Mitchell River Mouth, Nicholson River Mouth, and

Raymond Island North (Fig. 5.4a). Additionally, a single fish was caught at Rigby’s Island West.

More than half the juveniles were caught in February (54%), whereas March (23.2%) and April

(22.8%) had comparable catches (Fig. 5.4a).

197

In February, juveniles were mostly caught in the Mitchell River Mouth (55.6%), with Nicholson

River Mouth (15.9%), Raymond Island North (16.4%) and Broadlands (11.5%) contributing the remainder. In March, juveniles were mostly caught in Metung West (65.2%). Smaller numbers were caught in the Mitchell River Mouth (7.2%), Nicholson River Mouth (17.2%) and Raymond

Island North (10.4%). Juveniles were captured in April at Metung West (57.6%), Raymond Island

North (28.2%) and Nicholson River Mouth (14.1%) (Fig. 5.4a).

Juvenile distribution, 2016

In 2016, 535 fish were caught with 97.5% of these collected from vegetated patches at 10 sites:

Broadlands, Eagle Point, Jones Bay North, Metung West, Mitchell River Mouth, Newlands Arm,

Nicholson River Mouth, Raymond Island East, Raymond Island North and Wollaston Bay (Fig. 5.4b).

Like 2015, more than half of the juveniles were caught in February (53%), but in contrast to 2015, twice as many fish were caught in April (31.5%) than March (15.5%) (Fig. 5.4b).

In February, most juveniles were caught in either the Mitchell River Mouth (69.1%) or Nicholson

River Mouth (28.3%). A small number were caught in Eagle Point (0.3%), Metung West (0.6%) and

Raymond Island North (1.7%). In March, juvenile abundance was highest in the Nicholson River

Mouth (46.8%) and Mitchell River Mouth (21.2%). Metung West (7.7%), Raymond Island East

(7.1%), Broadlands (5.8%) and Wollaston Bay (5.8%) had lower abundances, and a small number were caught at Eagle Point (2.8%) and Raymond Island North (2.8%). Finally, in April, fish were distributed amongst the Nicholson River Mouth (41.8%), Metung West (20.9%), Raymond Island

East (15.3%) and Jones Bay North (13.9%). Smaller numbers of fish were found at Broadlands

(3.3%), Eagle Point (1.6%), Mitchell River Mouth (1.5%), Newlands Arm (0.9%) and Wollaston Bay

(0.8%) (Fig. 5.4b).

198

a)

199

b)

Figure 5.4. Pie charts showing the number (chart size) and monthly proportion (segment colour) of juvenile black bream (Acanthopagrus butcheri) caught per standard haul in a) 2015 (n=126) and b) 2016 (n=535). For comparison, the standarized haul for the Mitchell River in 2015 was 47, whereas in 2016 it was 151.

200

Cluster Analysis, 2015

Clustering of “larval period” data based on all 12 analysed elements resulted in 11 clusters in 2015.

The “salt-fraction dominant” approach yielded 6 clusters whereas clustering by “protein-fraction dominant elements” produced 5 clusters. While all clustering approaches resulted in significant differences between cohorts, only the “salt-fraction dominant” method yielded significant differences among juvenile recruitment sites.

Clustering of the primordia data by all 12 elements resulted in 4 clusters whereas clustering by either “salt-fraction dominant” or “protein-fraction” dominant yielded 3 clusters each. No method of clustering of primordia data resulted in significant differences between juvenile recruitment sites. However, both “salt-fraction dominant” and “all elements” resulted in significant differences between cohorts.

For 2015, the “salt-fraction dominant” approach produced the model with the greatest predictive power for both “larval period” and primordia data, having the highest R-squared value and lowest

Akaike Information Criterion (corrected) statistic (AICc). The elements contributing most to differences among clusters were identified through MANOVA (Appendix 5.1). Summaries of the six cluster analyses are presented in Table 5.1.

201

Table 5.1. Summaries of the 2015 cluster analyses of LA-ICP-MS data from sagittal otoliths extracted from juvenile black bream (Acanthopagrus butcheri). Juveniles were caught in vegetated patches throughout Lake King, Victoria, Australia during the autumn months (Feb-Apr). Clustering methods were assessed by multinomial logistic regression. Provided are the R-squared values (R2), an Akaike Information Criterion (corrected) test statistics (AICc) and p-values for the predictor variables, Cohort and Recruitment Site.

Approach R2 AICc p-value, p-value, Site Cohort

All elements 0.39 604.17 <0.0001 0.38

Larval period Salt-fraction 0.44 316.93 <0.0001 0.01 dominant n = 89 Protein-fraction 0.17 316.14 0.04 0.50 dominant

All elements 0.32 132.32 0.002 0.60

Primordium Salt-fraction 0.46 81.71 <0.0001 0.35 dominant n = 38 Protein-fraction 0.18 93.25 0.83 0.17 dominant

Cluster Analysis, 2016

Clustering based on all 12 analysed elements produced 13 separate clusters for 2016. Clustering by

“salt-fraction dominant” elements or “protein-fraction dominant” elements both resulted in 5 clusters each. Only the “salt-fraction dominant” approach detected significant effects of both cohort and juvenile recruitment site. Clustering by all 12 elements could only detect differences among juvenile recruitment sites, whereas the “protein-fraction dominant” yielded no significant effects.

The cluster analysis of primordia data using all 12 elements resulted in 9 clusters. In contrast to this, clustering on the basis of salt-fraction dominant elements produced 3 clusters and by protein-

202

fraction dominant elements yielded 5 clusters. No approach detected a significant effect of cohort.

Only the “salt-fraction dominant” yielded a significant effect of juvenile recruitment site.

As with the 2015 analyses, the “salt-fraction dominant” approach produced the highest R-squared values for both the “larval period” and primordium data. However, unlike 2015, it did not produce the lowest AICc statistic. It was, however, the best predictor of the effect of cohort and recruitment site. The elements contributing most to differences among clusters were identified through MANOVA (Appendix 5.2). Summaries of the six cluster analyses are presented in Table

5.2. As the “salt-fraction dominant” clustering approach explained the most variability, resulted in stronger differences among cohorts and recruit sites, and was, on average, the most supported model (based on AICc), it was chosen as the best approach for describing spatio-temporal differences in otolith chemistry.

Table 5.2. Summaries of the 2016 cluster analyses of LA-ICP-MS data from sagittal otoliths extracted from juvenile black bream (Acanthopagrus butcheri). Juveniles were caught in vegetated patches throughout Lake King, Victoria, Australia during the autumn months (Feb-Apr). Clustering methods were assessed by multinomial logistic regression. Provided are the R-squared values (R2), an Akaike Information Criterion (corrected) test statistics (AICc) and p-values for the predictor variables, Cohort and Recruitment Site.

Approach R2 AICc p-value, p-value, Site Cohort

All elements 0.26 2651.387 0.770 0.0009

Larval period Salt-fraction 0.27 536.6572 0.026 <0.0008 dominant n = 186 Protein-fraction 0.12 469.2616 0.82 0.50 dominant

All elements 0.29 -1327.05 0.48 0.30

Primordium Salt-fraction 0.35 198.6246 0.71 0.0006 dominant n = 96 Protein-fraction 0.15 562.9985 0.79 0.75 dominant

203

Cluster spatial distributions, 2015

Larval period clusters were strongly predictive of primordium clusters (χ2 likelihood-ratio 47.6, p<0.001, n=38). Primordium cluster A was mostly representative of larval period clusters A and D, primordium cluster B was composed almost exclusively of larval period cluster B juveniles, and primordium cluster C was constituted by larval period clusters E and F. Only 2 larval period cluster

C fish had primordium data. These two fish were assigned to primordium cluster A and B. The distribution of primordium and larval period chemistry clusters in Lake King for 2015 are presented in Figures 5.5 and 5.6, respectively.

Although 2015 had six larval period clusters, two clusters dominated each cohort. The first cohort was dominated by larval period clusters A and D, which were responsible for most of the cohort

(79%). Juveniles from cluster A were mostly found in the Mitchell River Mouth, whereas cluster D fish were found at the recruitment sites at Raymond Island North, Broadlands and the Nicholson

River Mouth. Cohort 2 was composed mainly of larval period cluster E and F juveniles (74%). These juveniles were mostly found at Metung West and Raymond Island North.

These four clusters (A, D, E, F) differed significantly (Tukey’s HSD pairwise comparisons) in 7Li, 11B,

55Mn, 88Sr and 138Ba (Fig. 5.7). Cluster A and D, dominant in the first cohort, both showed relative depletion in 7Li and 11B. Cluster A was relatively enriched in 55Mn and 88Sr and depleted in 138Ba.

Cluster D fish showed relative depletion of 55Mn and enrichment of both 88Sr and 138Ba. Clusters E and F, dominant in the second cohort fish, were distinguished by relatively low levels of both 88Sr and 138Ba. Cluster E showed relative enrichment in 7Li, 11B and 55Mn, whereas cluster F only showed enrichment of 7Li. A comparison of clusters A-F, based on all elements analysed, is presented in Appendix 5.2.

204

a)

205

b)

Figure 5.5. Pie charts showing proportional distribution of clusters based on otolith primordium chemistry for black bream (Acanthopagrus butcheri) for the (a) first and (b) second cohort for 2015, n = 38. For comparison, the number of fish from both the Mitchell River Mouth and Metung West was 8.

206

a)

207

b)

Figure 5.6. Pie charts showing proportional distribution of clusters based on otolith larval period chemistry for black bream (Acanthopagrus butcheri) for the (a) first and (b) second cohort for 2015, n = 89. For comparison, the number of fish from the Mitchell River Mouth was 24, whereas the number from Metung West was 14.

208

a)

b)

Figure 5.7. Comparison of dominant clusters based on otolith chemistry in the a) larval period and b) primordium for black bream (Acanthopagrus butcheri) caught in Lake King, Australia, 2015. Error bars represent +/- SE. For each element, letters indicate groupings assigned by a Tukey’s Honestly Significant Difference test.

209

Cluster spatial distributions, 2016

As with 2015, larval period clusters were strongly predictive of primordium clusters (χ2 likelihood- ratio 46.4, p<0.001, n=96). Primordium cluster A was mostly representative of equal amounts of larval period clusters A and B, primordium cluster B was mostly composed of larval period cluster

A, and primordium cluster C was constituted by larval period clusters C and D. Primordium data was available for only 3 larval period cluster E fish. These were assigned to primordium cluster B.

The distribution of larval period and primordium chemistry clusters in Lake King for 2016 are presented in Figures 5.8 and 5.9 respectively.

There were five larval period chemistry cluster groups for 2016. Cohort 1 was dominated by cluster B (52.8%), which corresponded to juveniles mainly from the Mitchell and Nicholson River

Mouth sites. Smaller amounts were contributed by clusters A (19.2%), C (12%) and E (16%). Nearly half (47.5%) of cohort 2 was cluster C. The remainder was made up by clusters A (18%), B (11.5%),

D (14.8%) and E (8.2%). Cluster C and D were generally found in either Metung West or Raymond

Island East.

These five clusters (A, B, C, D, E) differed significantly (Tukey’s HSD pairwise comparisons) in 7Li,

11B, 55Mn, 88Sr and 138Ba (Fig. 5.10). Clusters A, B and C showed either depletion or relatively low levels of 7Li, whereas clusters D and E were relatively enriched. In the case of 11B, clusters A, B and

E showed relatively low levels and clusters C and D showed enrichment. Clusters B, C and E showed relative depletion of 55Mn and clusters A and D showed higher levels. Like in 2015, clusters

C and D (corresponding to Metung West and Raymond Island East recruitment sites), showed very low levels of both 88Sr and 138Ba, whereas the other three clusters showed enrichment of these elements. A comparison of clusters A-E, based on all elements analysed, is presented in Appendix

5.2.

210

a)

211

b)

Figure 5.8. Pie charts showing proportional distribution of clusters based on otolith primordium chemistry for black bream (Acanthopagrus butcheri) for the (a) first and (b) second cohort for 2016, n=96. For comparison, the number of fish from Raymond Island North was 5, whereas the number from Raymond Island East was 13.

212

a)

213

b)

Figure 5.9. Pie charts showing proportional distribution of clusters based on otolith larval period chemistry for black bream (Acanthopagrus butcheri) for the (a) first and (b) second cohort for 2016, n=186. For comparison, the number of fish from Raymond Island North was 11, whereas the number from Raymond Island East was 18.

214

a)

b)

Figure 5.10. Comparison of dominant clusters based on otolith chemistry in the a) larval period and b) primordium for black bream (Acanthopagrus butcheri) caught in Lake King, Australia, 2016. Error bars represent +/- SE. For each element, letters indicate groupings assigned by a Tukey’s Honestly Significant Difference test.

215

DISCUSSION

Selection of elements for cluster analysis

The findings of this study show how the suite of elements chosen can influence the utility of otolith microchemistry data in reconstructing environmental histories in fish. I found, through comparison of clustering approaches, that only using elements that exist primarily (or completely) in the free ion phase in the endolymph resulted in elemental fingerprints that were the most distinct among juvenile cohorts and collection locations. These “salt-fraction dominant” elements are more likely to reflect ambient water chemistry due to their mechanisms of incorporation into the growing otolith (Thomas et al., 2017). Campana (1999) suggested that the incorporation of trace elements into the otolith is a highly regulated process, with barriers at each stage of the pathway. A given trace element enters the blood plasma from the water column via uptake at either the gills or intestine. From here, blood-borne ions must be transported into the epithelial cells of the inner ear (sacculum). Finally, they must be transported (either passively or actively) across the saccular cell membrane into the endolymph in the extracellular space (Payan et al.,

2004). Trace elements that enter the endolymph can be broadly classified into three categories: free ions (e.g. certain Group 1 and 2 elements), essential co-factors of proteins (e.g. transition metals – Mn, Cu, Zn), and non-metal protein constituents (e.g. S in cysteine, methionine; P in phosphate-groups). The utility of each these categories in physico-chemical reconstructions is explored below.

Among-individual variation in the concentrations of protein-associating trace elements is more likely to indicate differences in physiological histories than past environments. The inclusion of these elements (Cu, Zn, P, S, Pb) may at best be only useful in heavily polluted areas and may at worst add noise that obscures meaningful spatial variation. For example, Shima and Swearer

216

(2016) analysed larval otoliths from the common triplefin ( lapillum) to reconstruct dispersal histories. In that study, they assayed 12 trace elements, but excluded Cu and Zn (but not

Pb) from their cluster analyses as the concentrations of these elements failed to converge at the otolith edge of individuals collected at the same location and time (i.e., when individuals were experiencing the same environmental conditions). Similarly, in a multi-elemental fingerprinting study on two sole species (Solea solea and S. senegalensis), Tanner et al. (2012) identified both Pb and Cu as being unimportant in separating estuaries in a canonical discriminant analysis.

Interestingly, this research group conducted another study on S. solea, where they attempted to correctly classify juveniles to their estuary of collection through linear discriminant function analysis (LDFA) of trace element ratios (including Pb and Cu) from otolith edges (representing the period of collection). In that study, the average cross-validated classification success was only 57%, which in part was likely due to neither element showing significant differences in otolith edge concentrations among estuaries (Tanner et al., 2011). In another study, a multi-elemental LDFA

(employing 9 elements, including Cu, Zn and Pb) was used to classify French grunt (Haemulon flavolineatum) to site of capture (Chittaro et al., 2005). In that study, the authors were only able to poorly predict site of capture (47%), noting that all elements only contributed weakly to discrimination between sites. It may have been that poor classification was due to a lack of differences in elemental signatures among locations. Alternatively, Cu, Zn and Pb, being protein- associated and physiologically regulated, may have varied considerably among individuals with similar environmental histories, reducing the accuracy of classification.

Although Cu, Zn and Pb may be physiologically regulated, they have been shown to be incorporated in otoliths of fish from highly polluted water sources. In ecotoxicological studies, there has been keen interest in the use of trace elements in otoliths as potential markers of pollution exposure. For example, in a study by Ranaldi and Gagnon (2008a) on black bream, the

217

authors assessed 14 trace elements to determine whether there were site-specific differences within the Swan River, Australia. The Swan River is an estuary in an industrial urban area, and consequently has high levels of transition metals. In that study, Pb (but not Cu and Zn) showed significant differences between sites. This suggests that Pb can be used as a potential marker of industrial pollution (as it positively correlated with Pb-sediment concentrations). Barbee et al.

(2014) conducted a laboratory study investigating the effects of elevated Pb, Cu and Zn during fertilization and development on otoliths from larvae of an amphidromous fish, Galaxias maculatus. In that study, they found that while Pb and Cu were incorporated at moderate treatment levels, Zn was only incorporated at extremely high levels (4000 mg/kg). This is in agreement with Ranaldi and Gagnon (2008b) who found that Zn was only incorporated into otoliths of snapper (Pagrus auratus) via a dietary, rather than waterborne, route.

While otoliths can have levels of Cu, Zn and Pb that reflect water chemistry, this is only under concentrations that are high enough to overwhelm endogenous signals. Under normal environmental concentrations, it is likely that Cu, Zn and Pb are present in otoliths either wholly, or partially, due to physiological factors. Thomas et al. (2019) found considerable levels of Matrix metalloproteinase 2 (MMP 2) and Carbonic anhydrase (CA) in endolymph and otoliths from black bream. These two proteins are likely to play roles in otolith growth and contain Zn as a cofactor.

CA, in particular, is essential for biomineralization as it catalyses the conversion of carbon dioxide into biocarbonate ions necessary for aragonite precipitation (Tohse et al., 2004). In addition to this, Otolith Matrix Protein-1, an abundant otolith structural protein, also likely binds Zn as part of its tertiary structure (Murayama et al., 2000). Interestingly, Limburg and Elfman (2010) noted marked taxa-specific differences amongst Salmoniformes with regards to otolith Zn:Ca ratios and suggested that these could be explained by variation in expression of otolith proteins. I tested for any correlations between all elements analysed in the current study and found that Zn moderately

218

correlated with P (P=0.378), Cu (P=0.333) and Pb (P=0.403). This further supports the hypothesis that these four elements are under physiological control but does not rule out the possibility that their bioavailability and inclusion may be affected by environmental processes. Interestingly, P showed the highest correlation with Mg (P=0.8213) amongst all samples analysed. It may be that these two elements are together indicative of endogenous events.

Mg is tightly regulated and is an essential element in a number of cellular processes, including the binding of adenosine triphosphate (ATP) (Silva and Williams, 2001). Limburg et al. (2018) showed a positive correlation between Mg incorporation and metabolic rate. In that review, the authors hypothesized an elegant two-step process for Mg incorporation. Firstly, that Mg is only able to be transported through ionocytes (in sufficient amounts) during periods of high metabolic activity.

They speculate that only during these periods is there surplus energy for Mg transport. Secondly, the authors suggest that an increased level of water-soluble otolith proteins (i.e. those occurring in the otolith during high metabolic periods) allows for endolymph Mg to be incorporated into the otolith. In a study conducted by Heimbrand, Limburg and Hüssy, Mg and P were found to be the first and second most useful predictors of age based on otolith chemistry (K. Limburg, personal communication). I suggest that the fluctuations in Mg and P amounts in otoliths can be used to indicate seasonal differences in annuli that are reflective of temperature-driven changes in metabolic rate. As well as this, a decrease in overall Mg and P incorporation over time may be indicative of a change in metabolic rate with age, independent of seasonal fluctuations.

While the protein-associating elements were uninformative in clustering juvenile bream, the key elements in distinguishing between larval chemistry clusters were the free ion elements; Li, B, Mn,

Sr and Ba. Otolith levels of the latter two elements have been shown in numerous laboratory and field studies to be effective indicators of water chemistry (Elsdon and Gillanders, 2003, Martin et al., 2004, Elsdon and Gillanders, 2005c, Elsdon and Gillanders, 2005b, Walther and Thorrold, 2006,

219

Hicks et al., 2010). Sr is usually positively correlated with salinity (Panfili et al., 2015) whereas Ba has been shown to be negatively correlated with salinity, although with differences between laboratory and field studies (Elsdon and Gillanders, 2005a, Elsdon and Gillanders, 2005b).

Interestingly, I found a strong, positive correlation between Sr and Ba (P=0.775). Indeed, Sr:Ca and

Ba:Ca levels clearly distinguished between the dominant clusters for each cohort, across both years. I speculate below about the possible mechanisms leading to this unexpected positive correlation between Sr and Ba.

Li and B both proved to be useful indicators of spatio-temporal distributions in this study. In terms of the relationship of otolith Li:Ca levels with those of water, there is disagreement over whether its correlation is positive (Hicks et al., 2010) or negative (Milton and Chenery, 2001). However, it has been shown to be positively correlated with salinity (Milton and Chenery, 2001) and used successfully in environmental reconstructions (Hicks et al., 2010, Longmore et al., 2010, Miyan et al., 2016). There has also been some use of B in physico-chemical reconstructions (Tournois et al.,

2013). It has been suggested that it may have utility as a proxy for pH due to its incorporation into both inorganic aragonite (Mavromatis et al., 2015) as well as biogenic aragonite in other taxa

(Anagnostou et al., 2012, McCulloch et al., 2012). It is assumed that B incorporates into the aragonite lattice by substituting carbonate as either BO3 or BO4, and that B incorporation is positively correlated with pH (Mavromatis et al., 2015).

When including Mn in clustering, it was considered a “salt-fraction dominant” element in larval period analyses but a “protein-fraction dominant” element in primordia analyses. It has been frequently found in the primordia of otoliths, along with other elements, where it has long been supposed to act as a cofactor for proteins present during nucleation (Brophy et al., 2004,

Ruttenberg et al., 2005, Limburg et al., 2013). As stated above, I previously identified a Mn-binding biomineralization protein, FAM20C, in high amounts in the otoliths of adult black bream (Thomas

220

et al., 2019). I speculated in that study that FAM20C causes the phosphorylation of a key biomineralization protein, Starmaker-like (Stm-l). Stm-l is crucial to crystal nucleation and polymorph selection, with its effectiveness being directly related to its degree of phosphorylation

(Wojtas and Dobryszycki, 2011, Wojtas et al., 2012, Rozycka et al., 2014). While FAM20C is undoubtedly expressed at otolith nucleation (along with Stm-l), and likely is the reason for elevated Mn in the primordia, it cannot be ruled out that it continues to be expressed throughout otolith growth, and therefore contributes to daily increment Mn:Ca levels. It is worth noting, here, however, that Thomas et al. (2019) did not detect FAM20C in endolymph of adult black bream, suggesting that it is expressed either at an earlier life stage, or at a different time than that when endolymph was harvested for the study. Regardless, I also previously found Mn present as a free ion in the endolymph of adult black bream, indicating that it is capable of traversing the cellular barriers between water, blood and endolymph. Mn has a comparable ionic radius to Ca as well as the same oxidation state (+2). It has also been shown to substitute for Ca in bivalve shells (Soldati et al., 2016). It is therefore a reasonable candidate for substitution for Ca in the otolith aragonite lattice (Shannon, 1976, Thomas et al., 2017).

There is disagreement, however, over whether otolith Mn:Ca levels are indicative of Mn:Ca ratios in water. In a laboratory study by Elsdon and Gillanders (2003) on juvenile black bream, the authors found that, unlike Ba and Sr, otolith Mn:Ca levels were mildly negatively correlated with water Mn:Ca levels. In a separate laboratory study on black rockfish (Sebastes melanops), Miller

(2009) found no relationship between otolith and water Mn:Ca levels. However, in contrast to this, in field studies, both Dorval et al. (2007) and Hamer et al. (2006) found positive correlations between water chemistry and otolith Mn incorporation.

Limburg et al. (2015) hypothesised that hypoxic conditions lead to a change in the oxidation state of Mn, allowing it to be in freely dissolved in water and therefore available for incorporation into

221

calcified structures such as the otolith. In a study on Baltic cod (Gadus morhua), Limburg et al.

(2011) found clear correlations between hypoxic conditions and Mn:Ca otolith levels. Similar relationships between oxygenation of the water column and otolith Mn incorporation were found in studies on Atlantic croaker (Micropogonias undulates) (Mohan and Walther, 2016, Altenritter et al., 2018). In contrast to this, in a laboratory study by Mohan et al. (2014), the authors found no relationship between anoxic stress conditions and otolith chemistry, concluding that physiology did not play a major role in its incorporation into increments.

Considering that several studies have successfully used Mn:Ca ratios to reclassify fish to their natal estuaries (Thorrold et al., 1998a, Swearer et al., 2003, Reis-Santos et al., 2012) and that kinetically,

Mn2+ is a likely substituent for Ca2+ (Thomas et al., 2017), it is likely that Mn:Ca ratios are primarily indicative of exogenous processes. Notably, Williams et al. (2018) found strong spatio-temporal patterns for Mn:Ca (and Li:Ca, K:Ca) in otoliths from black bream larvae caught in the Gippsland

Lakes.

In summary, protein-associating elements are more likely to be indicative of endogenous events and are only incorporated into the otolith as a result of water chemistry when ambient concentrations are extremely high. In contrast to this, elements that are present in the endolymph as free ions (i.e. Li, B, Mn, Sr) are more likely to be incorporated into the otolith as a result of water chemistry and so be indicative of dispersal and natal sources.

Spawning sources for black bream in the Gippsland Lakes

In 2015, there were two dominant clusters in each cohort: A and D for cohort 1, and E and F for cohort 2. Li and Sr proved to be the most useful in distinguishing between cohorts. Those clusters most strongly associated with the Mitchell River (Cluster A) and Nicholson River (Cluster D) were

222

both low in Li but high in Sr. For clusters E and F (that were largely found in Raymond Island North and Metung West) the opposite was true. Li and Sr have been previously shown to be effective in multi-element characterization of black bream movement in the Gippsland Lakes (Williams et al.,

2018). In 2016, a similar patterning of Li and Sr was seen across clusters dominating either the

Mitchell/Nicholson rivers (A, B) or Raymond Island East/Metung West (C, D). Under low flow conditions, larvae are thought to be retained in salt-wedge estuaries, recruiting close to the mouths of the tributaries (Williams et al., 2012, Gee, 2014). It may be possible that at high flow conditions, larvae are swept away from their natal tributary and recruit to vegetation patches further away from their origin. The differences in otolith chemistry may indicate either that these clusters originated at different spawning locations that are distinguished by differences in salinity, or that they originated in the same spawning tributaries but experienced large freshwater flow events that transported them further into the Lakes. It should be noted here, however, that Sarre and Potter (2000), suggested that high freshwater flow may negatively impact recruitment success for black bream.

Clusters most strongly associated with the Mitchell River and Nicholson River Mouth recruitment sites (A, D – 2015; A, B – 2016) were depleted in Li and B but enriched in Sr. Intra-annual differences were mostly driven by Mn for both years, and Ba for 2015. It is likely, considering previous studies on black bream larvae in the Gippsland Lakes (Williams et al., 2012) that the majority of fish were spawned in the closest river and were subsequently swept to its mouth where they settled in vegetated patches. The distributions of these clusters across the northern recruitment sites of Lake King, however, suggest that a reasonable amount of “mixing” of larvae was occurring, with juveniles recruiting to a variety of sites such as Raymond Island North.

Williams et al. (2018) found spatio-temporal patterns in Mn:Ca for black bream larval otoliths from the Gippsland Lakes. In that study, they suggested a negative correlation between Mn:Ca and

223

dissolved oxygen and were thus useful in discriminating between water layers. Additionally,

Jenkins et al. (2018) reported that dissolved oxygen was consistently low beneath the halocline of all three main tributaries in the Gippsland Lakes and that variation in hypoxia was likely to be driven by seasonal freshwater flow (Newton, 1996, Kurup and Hamilton, 2002). Therefore, the differences in Mn:Ca levels between clusters may be the result of the relative amounts of dissolved oxygen present in the halocline at the time of spawning.

Interestingly, a comparison of the crucial Metung West/Raymond Island clusters for each year

(2015 – E, F; 2016 – C, D) revealed an interannual consistency of otolith chemistry (Fig. 5.11).

Figure 5.11. Interannual comparison of otolith larval chemistry for juvenile black bream (Acanthopagrus butcheri) caught in Lake King, Australia, in 2015 and 2016. Clusters are associated with recruitment sites at Raymond Island and Metung West. Error bars represent 95% confidence intervals.

Swearer et al. (2003) previously suggested the possibility that otolith chemistry might be spatially consistent across years. The concentration of Sr in water varies non-linearly with salinity, with the greatest drop in its concentration occurring at salinities below 10 PSU (Ingram and Sloan, 1992,

224

Secor et al., 1995). Consequently, depletion in Sr is usually indicative of a freshwater, rather than brackish or marine, environment. However, these clusters also showed markedly lower Ba – inconsistent with the prevalent view that Sr and Ba are inversely proportional (Kimura et al., 2000,

Secor and Rooker, 2000, Gillanders, 2005). Additionally, Williams et al. (2018) reported negative correlations between Sr and Ba in the Gippsland Lakes tributaries. There are two scenarios that might have led to these unusual signatures. Firstly, all black bream larvae may have been spawned in the Mitchell and Nicholson Rivers, as consistent with previous studies (Williams et al., 2012,

Williams et al., 2013, Gee, 2014, Williams et al., 2018). These larvae may have experienced an increase of river flow, dispersing them to vegetated patches on Raymond Island and Metung West.

This scenario would explain the depletion in Sr, but not Ba. Barium has been shown, however, to be utilised by phytoplankton species, which can deplete it from the water column (Sternberg et al.,

2005), suggesting a second, and more compelling scenario. It is possible that fish caught in Metung

West and Raymond Island during the later months originated in a relatively low saline environment with high phytoplankton activity. In 2015, some cluster F fish were found in the

Tambo River Mouth, a site that lacked vegetation. Considering the distances that larvae can be dispersed (Jenkins et al., 1999, Schludermann et al., 2012, Lechner et al., 2016), these juveniles may have been spawned in the Tambo River and then recruited to vegetated patches at Metung

West and Raymond Island. However, as the Sr and Ba signatures are not consistent with the

Tambo River (Williams et al., 2018), it is more likely that the larvae originated in a different, unknown freshwater source. In support of this, in an associated study to mine, modelling of black bream spawning for the years 2015 and 2016 indicated dispersal over a large distance and suggested the presence of a previously unidentified natal source (Woodland et al., Unpublished).

Finally, the differences in B seen across these four clusters is likely explained by variations in pH over time.

225

CONCLUSION

From the microchemical clusters found in juvenile black bream distributed across 2015 and 2016 in Lake King, it is evident that multiple tributaries/areas are being used as spawning sites.

Consequently, major shifts to the discharge and stratification of any of these tributaries may have strong impacts on year class strength. Furthermore, although the Mitchell, Nicholson and Tambo

Rivers have been considered to be the main tributaries of interest, it is clear that there is a yet-to- be-identified natal source. The fish from this unknown water source are spawned later in the year than originally thought, and, come from a distinctly different chemical environment.

With regards to multi-element fingerprinting as a means to construct physico-chemical histories, it is recommended that elements that are primarily in the salt-phase of the endolymph are employed in discriminant analyses or clustering approaches. However, although protein- associated elements may not be indicative of ambient water conditions over time, they may be indicative of endogenous events. This potentially opens up an extremely exciting and promising field – the reconstruction of physiological events and factors such as maternal investment, growth, condition and sexual maturation.

ACKNOWLEDGEMENTS

We thank D. Brehm, A. Gebler, R. Woodland, F. Warry, S. Chia, V. Komyakova, and D. Chamberlain for field and laboratory assistance. Funding was provided by the Australian Research Council

(project LP140100087) and a Holsworth Wildlife Research Endowment – Equity Trustee Charitable

Foundation Grant (to ORBT). ORBT was supported by an Australian Postgraduate Award. The present study was conducted under Monash University Animal Ethics Committee approval

BSCI/2014/19 and the State of Victoria Fisheries research permit #1204.

226

REFERENCES

Altenritter, M.E., Cohuo, A., Walther, B.D. (2018) Proportions of demersal fish exposed to

sublethal hypoxia revealed by otolith chemistry. Marine Ecology Progress Series 589, 193-

208. [In English].

Anagnostou, E., Huang, K.F., You, C.F., Sikes, E.L., Sherrell, R.M. (2012) Evaluation of boron isotope

ratio as a pH proxy in the deep sea coral Desmophyllum dianthus: Evidence of

physiological pH adjustment. Earth and Planetary Science Letters 349, 251-260. [In

English].

Barbee, N.C., Ganio, K., Swearer, S.E. (2014) Integrating multiple bioassays to detect and assess

impacts of sublethal exposure to metal mixtures in an estuarine fish. Aquatic Toxicology

152, 244-255.

Barbee, N.C., Swearer, S.E. (2007) Characterizing natal source population signatures in the

diadromous fish Galaxias maculatus, using embryonic otolith chemistry. Marine Ecology

Progress Series 343, 273-282.

Bath, G.E., Thorrold, S.R., Jones, C.M., Campana, S.E., McLaren, J.W., Lam, J.W.H. (2000) Strontium

and barium uptake in aragonitic otoliths of marine fish. Geochimica et Cosmochimica Acta

64, 1705-1714.

Beer, N.A., Wing, S.R., Swearer, S.E. (2011) Otolith elemental evidence for spatial structuring in a

temperate reef fish population. Marine Ecology Progress Series 442, 217-227.

Bird, E.C.F. (1965) The geomorphology of the Gippsland lakes, Vol., Australian National University,

Canberra.

227

Brophy, D., Jeffries, T.E., Danilowicz, B.S. (2004) Elevated manganese concentrations at the cores

of clupeid otoliths: possible environmental, physiological, or structural origins. Marine

Biology 144, 779-786.

Butcher, A.D., Ling, J.K., Victoria, Fisheries, Wildlife, D. (1962) Bream tagging experiments in East

Gippsland during April and May 1944, Vol., Fisheries and Wildlife Dept., Victoria,

[Melbourne].

Campana, S.E. (1999) Chemistry and composition of fish otoliths: pathways, mechanisms and

applications. Marine Ecology Progress Series 188, 263-297.

Chittaro, P.M., Usseglio, P., Fryer, B.J., Sale, P.F. (2005) Using otolith microchemistry of Haemulon

flavolineatum (French Grunt) to characterize mangroves and coral reefs throughout

Turneffe Atoll, Belize: Difficulties at small spatial scales. Estuaries 28, 373-381.

Crook, D.A., Macdonald, J.I., O'Connor, J.P., Barry, B. (2006) Use of otolith chemistry to examine

patterns of diadromy in the threatened Australian grayling Prototroctes maraena. Journal

of Fish Biology 69, 1330-1344. de Vries, M.C., Gillanders, B.M., Elsdon, T.S. (2005) Facilitation of barium uptake into fish otoliths:

Influence of strontium concentration and salinity. Geochimica et Cosmochimica Acta 69,

4061-4072.

Dorval, E., Jones, C.M., Hannigan, R., van Montfrans, J. (2007) Relating otolith chemistry to surface

water chemistry in a coastal plain estuary. Canadian Journal of Fisheries and Aquatic

Sciences 64, 411-424.

Doubleday, Z.A., Harris, H.H., Izzo, C., Gillanders, B.M. (2014) Strontium randomly substituting for

calcium in fish otolith aragonite. Analytical Chemistry 86, 865-869.

228

Dunkelberger, D.G., Dean, J.M., Watabe, N. (1980) The ultrastructure of the otolithic membrane

and otolith in the juvenile mummichog, Fundulus heteroclitus. Journal of Morphology 163,

367-377. [In English].

Elsdon, T.S., Gillanders, B.M. (2002) Interactive effects of temperature and salinity on otolith

chemistry: challenges for determining environmental histories of fish. Canadian Journal of

Fisheries and Aquatic Sciences 59, 1796-1808.

Elsdon, T.S., Gillanders, B.M. (2003) Relationship between water and otolith elemental

concentrations in juvenile black bream Acanthopagrus butcheri. Marine Ecology Progress

Series 260, 263-272.

Elsdon, T.S., Gillanders, B.M. (2004) Fish otolith chemistry influenced by exposure to multiple

environmental variables. Journal of Experimental Marine Biology and Ecology 313, 269-

284.

Elsdon, T.S., Gillanders, B.M. (2005a) Alternative life-history patterns of estuarine fish: barium in

otoliths elucidates freshwater residency. Canadian Journal of Fisheries and Aquatic

Sciences 62, 1143-1152. [In English].

Elsdon, T.S., Gillanders, B.M. (2005b) Consistency of patterns between laboratory experiments and

field collected fish in otolith chemistry: an example and applications for salinity

reconstructions. Marine and Freshwater Research 56, 609-617.

Elsdon, T.S., Gillanders, B.M. (2005c) Strontium incorporation into calcified structures: separating

the effects of ambient water concentration and exposure time. Marine Ecology Progress

Series 285, 233-243.

Elsdon, T.S., Wells, B.K., Campana, S.E., et al. (2008) Otolith chemistry to describe movements and

life-history parameters of fishes: Hypotheses, assumptions, limitations and inferences. In:

229

Oceanography and Marine Biology: An Annual Review, Vol 46. Vol. 46. Oceanography and

Marine Biology (Eds. R.N. Gibson, R.J.A. Atkinson, J.D.M. Gordon), pp. 297-+.

FitzGerald, J.L., Thorrold, S.R., Bailey, K.M., Brown, A.L., Severin, K.P. (2004) Elemental signatures

in otoliths of larval walleye pollock (Theragra chalcogramma) from the northeast Pacific

Ocean. Fishery Bulletin 102, 604-616.

Gee, E.M. (2014) Linking freshwater flows, salt wedge dynamics and black bream dispersal in the

Mitchell River sub-estuary, Vol.

Gillanders, B.M. (2002) Temporal and spatial variability in elemental composition of otoliths:

implications for determining stock identity and connectivity of populations. Canadian

Journal of Fisheries and Aquatic Sciences 59, 669-679.

Gillanders, B.M. (2005) Using elemental chemistry of fish otoliths to determine connectivity

between estuarine and coastal habitats. Estuarine Coastal and Shelf Science 64, 47-57.

Gillanders, B.M., Kingsford, M.J. (2002) Impact of changes in flow of freshwater on estuarine and

open coastal habitats and the associated organisms. In: Oceanography and Marine

Biology, Vol 40. Vol. 40. Oceanography and Marine Biology (Eds. R.N. Gibson, M. Barnes,

R.J.A. Atkinson), Taylor & Francis Ltd, London, pp. 233-309.

Haddy, J.A., Pankhurst, N.W. (1998) Annual change in reproductive condition and plasma

concentrations of sex steroids in black bream, Acanthopagrus butcheri (Munro) (Sparidae).

Marine and Freshwater Research 49, 389-397. [In English].

Haddy, J.A., Pankhurst, N.W. (2000) The effects of salinity on reproductive development, plasma

steroid levels, fertilisation and egg survival in black bream Acanthopagrus butcheri.

Aquaculture 188, 115-131. [In English].

230

Hamer, P.A., Jenkins, G.P., Coutin, P. (2006) Barium variation in Pagrus auratus (Sparidae) otoliths:

A potential indicator of migration between an embayment and ocean waters in south-

eastern Australia. Estuarine Coastal and Shelf Science 68, 686-702.

Hassell, K.L., Coutin, P.C., Nugegoda, D. (2008) Hypoxia, low salinity and lowered temperature

reduce embryo survival and hatch rates in black bream Acanthopagrus butcheri (Munro,

1949). Journal of Fish Biology 72, 1623-1636. [In English].

Hicks, A.S., Closs, G.P., Swearer, S.E. (2010) Otolith microchemistry of two amphidromous galaxiids

across an experimental salinity gradient: A multi-element approach for tracking

diadromous migrations. Journal of Experimental Marine Biology and Ecology 394, 86-97.

Hindell, J.S., Jenkins, G.P., Womersley, B. (2008) Habitat utilisation and movement of black bream

Acanthopagrus butcheri (Sparidae) in an Australian estuary. Marine Ecology Progress

Series 366, 219-229.

Ingram, B.L., Sloan, D. (1992) Strontium isotopic composition of estuarine sediments as

paleosalinity-paleoclimate indicator. Science 255, 68-72. [In English].

Irisson, J.O., Paris, C.B., Leis, J.M., Yerman, M.N. (2015) With a little help from my friends: Group

orientation by larvae of a coral reef fish. PloS One 10, e0144060. [In English].

Izzo, C., Doubleday, Z.A., Gillanders, B.M. (2016a) Where do elements bind within the otoliths of

fish? Marine and Freshwater Research 67, 1072-1076.

Izzo, C., Doubleday, Z.A., Grammer, G.L., et al. (2016b) Fish as proxies of ecological and

environmental change. Reviews in Fish Biology and Fisheries 26, 265-286. [In English].

Jenkins, G.P., Black, K.P., Keough, M.J. (1999) The role of passive transport and the influence of

vertical migration on the pre-settlement distribution of a temperate, demersal fish:

231

numerical model predictions compared with field sampling. Marine Ecology Progress

Series 184, 259-271. [In English].

Jenkins, G.P., Conron, S.D., Morison, A.K. (2010) Highly variable recruitment in an estuarine fish is

determined by salinity stratification and freshwater flow: implications of a changing

climate. Marine Ecology Progress Series 417, 249-261. [In English].

Jenkins, G.P., Kent, J.A., Woodland, R.J., Warry, F., Swearer, S.E., Cook, P.L.M. (2018) Delayed

timing of successful spawning of an estuarine dependent fish, black bream Acanthopagrus

butcheri. Journal of Fish Biology 0.

Jenkins, G.P., Spooner, D., Conron, S., Morrongiello, J.R. (2015) Differing importance of salinity

stratification and freshwater flow for the recruitment of apex species of estuarine fish.

Marine Ecology Progress Series 523, 125-144. [In English].

Jenkins, G.P., Sutherland, C.R. (1997) The influence of habitat structure on nearshore fish

assemblages in a southern Australian embayment: Colonisation and turn-over rate of

fishes associated with artificial macrophyte beds of varying physical structure. Journal of

Experimental Marine Biology and Ecology 218, 103-125.

Kalish, J.M. (1990) Use of otolith microchemistry to distinguish the progeny of sympatric

anadromous and non-anadromous salmonids. Fishery Bulletin 88, 657-666. [In English].

Kang, Y., He, Y.X., Zhao, M.X., Li, W.F. (2011) Structures of native and Fe-substituted SOD2 from

Saccharomyces cerevisiae. Acta Crystallographica Section F-Structural Biology

Communications 67, 1173-1178. [In English].

Kench, P.S. (1999) Geomorphology of Australian estuaries: Review and prospect. Australian

Journal of Ecology 24, 367-380.

232

Kimura, R., Secor, D.H., Houde, E.D., Piccoli, P.M. (2000) Up-estuary dispersal of young-of-the-year

bay anchovy Anchoa mitchilli in the Chesapeake Bay: inferences from microprobe analysis

of strontium in otoliths. Marine Ecology Progress Series 208, 217-227.

Kingsford, M.J., Leis, J.M., Shanks, A., Lindeman, K.C., Morgan, S.G., Pineda, J. (2002) Sensory

environments, larval abilities and local self-recruitment. Bulletin of Marine Science 70,

309-340. [In English].

Kraus, R.T., Secor, D.H. (2004) Incorporation of strontium into otoliths of an estuarine fish. Journal

of Experimental Marine Biology and Ecology 302, 85-106.

Kurup, R.G., Hamilton, D.P. (2002) Flushing of dense, hypoxic water from a cavity of the Swan River

estuary, Western Australia. Estuaries 25, 908-915. [In English].

Lechner, A., Keckeis, H., Humphries, P. (2016) Patterns and processes in the drift of early

developmental stages of fish in rivers: a review. Reviews in Fish Biology and Fisheries 26,

471-489. [In English].

Leis, J.M., Carson-Ewart, B.M., Hay, A.C., Cato, D.H. (2003) Coral-reef sounds enable nocturnal

navigation by some reef-fish larvae in some places and at some times. Journal of Fish

Biology 63, 724-737. [In English].

Leis, J.M., Sweatman, H.P.A., Reader, S.E. (1996) What the pelagic stages of coral reef fishes are

doing out in blue water: Daytime field observations of larval behavioural capabilities.

Marine and Freshwater Research 47, 401-411. [In English].

Levin, L.A. (2006) Recent progress in understanding larval dispersal: New directions and

digressions. Oxford University Press, p. 282.

233

Limburg, K.E. (1995) Otolith strontium traces environmental history of subyearling american shad

Alosa sapidissima. Marine Ecology Progress Series 119, 25-35. [In English].

Limburg, K.E., Elfman, M. (2010) Patterns and magnitude of Zn:Ca in otoliths support the recent

phylogenetic typology of Salmoniformes and their sister groups. Canadian Journal of

Fisheries and Aquatic Sciences 67, 597-604.

Limburg, K.E., Hayden, T.A., Pine, W.E., III, Yard, M.D., Kozdon, R., Valley, J.W. (2013) Of travertine

and time: Otolith chemistry and microstructure detect provenance and demography of

endangered humpback chub in grand canyon, USA. PloS One 8.

Limburg, K.E., Olson, C., Walther, Y., Dale, D., Slomp, C.P., Hoie, H. (2011) Tracking Baltic hypoxia

and cod migration over millennia with natural tags. Proceedings of the National Academy

of Sciences of the United States of America 108, E177-E182.

Limburg, K.E., Walther, B.D., Lu, Z., et al. (2015) In search of the dead zone: Use of otoliths for

tracking fish exposure to hypoxia. Journal of Marine Systems 141, 167-178.

Limburg, K.E., Wuenschel, M.J., Hüssy, K., Heimbrand, Y., Samson, M. (2018) Making the otolith

magnesium chemical calendar-clock tick: Plausible mechanism and empirical evidence.

Reviews in Fisheries Science & Aquaculture 26, 479-493. [In English].

Longmore, C., Fogarty, K., Neat, F., et al. (2010) A comparison of otolith microchemistry and

otolith shape analysis for the study of spatial variation in a deep-sea teleost,

Coryphaenoides rupestris. Environmental Biology of Fishes 89, 591-605.

Lothian, A., Lago, L., Mukherjee, S., et al. (2018) Characterization of the metal status of natively

purified Alpha Synuclein from human blood, brain tissue, or recombinant sources using

size exclusion ICP-MS reveals no significant binding of Cu, Fe or Zn. Metallomics.

234

Martin, G.B., Thorrold, S.R., Jones, C.M. (2004) Temperature and salinity effects on strontium

incorporation in otoliths of larval spot (Leiostomus xanthurus). Canadian Journal of

Fisheries and Aquatic Sciences 61, 34-42. [In English].

Mavromatis, V., Montouillout, V., Noireaux, J., Gaillardet, J., Schott, J. (2015) Characterization of

boron incorporation and speciation in calcite and aragonite from co-precipitation

experiments under controlled pH, temperature and precipitation rate. Geochimica et

Cosmochimica Acta 150, 299-313. [In English].

McCulloch, M., Trotter, J., Montagna, P., et al. (2012) Resilience of cold-water scleractinian corals

to ocean acidification: Boron isotopic systematics of pH and saturation state up-regulation.

Geochimica et Cosmochimica Acta 87, 21-34. [In English].

Miller, J.A. (2009) The effects of temperature and water concentration on the otolith

incorporation of barium and manganese in black rockfish Sebastes melanops. Journal of

Fish Biology 75, 39-60.

Miller, J.A. (2011) Effects of water temperature and barium concentration on otolith composition

along a salinity gradient: Implications for migratory reconstructions. Journal of

Experimental Marine Biology and Ecology 405, 42-52.

Miller, M.B., Clough, A.M., Batson, J.N., Vachet, R.W. (2006) Transition metal binding to cod otolith

proteins. Journal of Experimental Marine Biology and Ecology 329, 135-143.

Milton, D.A., Chenery, S.R. (2001) Sources and uptake of trace metals in otoliths of juvenile

barramundi (Lates calcarifer). Journal of Experimental Marine Biology and Ecology 264, 47-

65.

235

Miyan, K., Khan, M.A., Patel, D.K., Khan, S., Prasad, S. (2016) Otolith fingerprints reveal stock

discrimination of Sperata seenghala inhabiting the Gangetic river system. Ichthyological

Research 63, 294-301. [In English].

Mohan, J., Rahman, M.S., Thomas, P., Walther, B. (2014) Influence of constant and periodic

experimental hypoxic stress on Atlantic croaker otolith chemistry. Aquatic Biology 20, 1-

11.

Mohan, J., Walther, B. (2016) Out of breath and hungry: Natural tags reveal trophic resilience of

Atlantic croaker to hypoxia exposure. Marine Ecology Progress Series 560, 207-221.

Mugiya, Y., Uchimura, T. (1989) Otolith resorption induced by anaerobic stress in the goldfish,

Carassius auratus. Journal of Fish Biology 35, 813-818. [In English].

Murayama, E., Okuno, A., Ohira, T., Takagi, Y., Nagasawa, H. (2000) Molecular cloning and

expression of an otolith matrix protein cDNA from the rainbow trout, Oncorhynchus

mykiss. Comparative Biochemistry and Physiology B-Biochemistry & Molecular Biology 126,

511-520.

Newton, G.M. (1996) Estuarine ichthyoplankton ecology in relation to hydrology and zooplankton

dynamics in a salt-wedge estuary. Marine and Freshwater Research 47, 99-111. [In

English].

Nicholson, G., Jenkins, G.P., Sherwood, J., Longmore, A. (2008) Physical environmental conditions,

spawning and early-life stages of an estuarine fish: climate change implications for

recruitment in intermittently open estuaries. Marine and Freshwater Research 59, 735-

749. [In English].

236

Panfili, J., Darnaude, A.M., Vigliola, L., Jacquart, A., Labonne, M., Gilles, S. (2015) Experimental

evidence of complex relationships between the ambient salinity and the strontium

signature of fish otoliths. Journal of Experimental Marine Biology and Ecology 467, 65-70.

Pannella, G. (1971) Fish otoliths: daily growth layers and periodical patterns. Science (New York,

N.Y.) 173, 1124-1127.

Patterson, H.M., McBride, R.S., Julien, N. (2004) Population structure of red drum (Sciaenops

ocellatus) as determined by otolith chemistry. Marine Biology 144, 855-862.

Payan, P., De Pontual, H., Boeuf, G., Mayer-Gostan, N. (2004) Endolymph chemistry and otolith

growth in fish. Comptes Rendus Palevol 3, 535-547.

R Core Team (2017) R: A language and environment for statistical computing., R Foundation for

Statistical Computing, Vienna, Austria.

Radtke, R.L. (1982) Strontium-calcium concentration ratios in fish otoliths as temperature

chronometers. American Zoologist 22, 963-963. [In English].

Ranaldi, M.M., Gagnon, M.M. (2008a) Trace metal incorporation in otoliths of black bream

(Acanthopagrus butcheri munro), an indicator of exposure to metal contamination. Water

Air and Soil Pollution 194, 31-43. [In English].

Ranaldi, M.M., Gagnon, M.M. (2008b) Zinc incorporation in the otoliths of juvenile pink snapper

(Pagrus auratus Forster): The influence of dietary versus waterborne sources. Journal of

Experimental Marine Biology and Ecology 360, 56-62.

Reis-Santos, P., Gillanders, B.M., Tanner, S.E., Vasconcelos, R.P., Elsdon, T.S., Cabral, H.N. (2012)

Temporal variability in estuarine fish otolith elemental fingerprints: Implications for

connectivity assessments. Estuarine Coastal and Shelf Science 112, 216-224.

237

Reis-Santos, P., Tanner, S.E., Vasconcelos, R.P., Elsdon, T.S., Cabral, H.N., Gillanders, B.M. (2013)

Connectivity between estuarine and populations: contributions of estuaries

are not consistent over time. Marine Ecology Progress Series 491, 177-186.

Rigby, B.A. (1984) The ecology of fish inhabiting estuarine seagrass habitats in the Gippsland Lakes.

Masters of Science Thesis (M.Sc.), University of Melbourne, 93 pages.

Rozycka, M., Wojtas, M., Jakob, M., et al. (2014) Intrinsically disordered and pliable Starmaker-like

protein from medaka (Oryzias latipes) controls the formation of calcium carbonate

crystals. PloS One 9, e114308.

Ruttenberg, B.I., Hamilton, S.L., Hickford, M.J.H., et al. (2005) Elevated levels of trace elements in

cores of otoliths and their potential for use as natural tags. Marine Ecology Progress Series

297, 273-281.

Sakabe, R., Lyle, J.M. (2010) The influence of tidal cycles and freshwater inflow on the distribution

and movement of an estuarine resident fish Acanthopagrus butcheri. Journal of Fish

Biology 77, 643-660. [In English].

Sarre, G.A., Potter, I.C. (1999) Comparisons between the reproductive biology of black bream

Acanthopagrus butcheri (Teleostei: Sparidae) in four estuaries with widely differing

characteristics. International Journal of Salt Lake Research 8, 179-210.

Sarre, G.A., Potter, I.C. (2000) Variation in age compositions and growth rates of Acanthopagrus

butcheri (Sparidae) among estuaries: some possible contributing factors. Fishery Bulletin

98, 785-799.

Schludermann, E., Tritthart, M., Humphries, P., Keckeis, H. (2012) Dispersal and retention of larval

fish in a potential nursery habitat of a large temperate river: an experimental study.

Canadian Journal of Fisheries and Aquatic Sciences 69, 1302-1315.

238

Secor, D.H., Hendersonarzapalo, A., Piccoli, P.M. (1995) Can otolith microchemistry chart patterns

of migration and habitat utilization in anadromous fishes. Journal of Experimental Marine

Biology and Ecology 192, 15-33. [In English].

Secor, D.H., Rooker, J.R. (2000) Is otolith strontium a useful scalar of life cycles in estuarine fishes?

Fisheries Research 46, 359-371.

Secor, D.H., Rooker, J.R., Zlokovitz, E., Zdanowicz, V.S. (2001) Identification of riverine, estuarine,

and coastal contingents of Hudson River striped bass based upon otolith elemental

fingerprints. Marine Ecology Progress Series, 245-253.

Shannon, R.D. (1976) Revised effective ionic-radii and systematic studies of interatomic distances

in halides and chalcogenides. Acta Crystallographica Section A 32, 751-767. [In English].

Shima, J.S., Swearer, S.E. (2009) Larval quality is shaped by matrix effects: implications for

connectivity in a marine metapopulation. Ecology 90, 1255-1267. [In English].

Shima, J.S., Swearer, S.E. (2016) Evidence and population consequences of shared larval dispersal

histories in a marine fish. Ecology 97, 25-31. [In English].

Silva, J.J.R.F.d., Williams, R.J.P. (2001) The biological chemistry of the elements: the inorganic

chemistry of life, 2nd edn Vol., Oxford University Press, Oxford, UK.

Soldati, A.L., Jacob, D.E., Glatzel, P., Swarbrick, J.C., Geck, J. (2016) Element substitution by living

organisms: the case of manganese in mollusc shell aragonite. Scientific Reports 6, 9. [In

English].

Sternberg, E., Tang, D.G., Ho, T.Y., Jeandel, C., Morel, F.M.M. (2005) Barium uptake and adsorption

in diatoms. Geochimica et Cosmochimica Acta 69, 2745-2752. [In English].

239

Sturrock, A.M., Hunter, E., Milton, J.A., et al. (2015) Quantifying physiological influences on otolith

microchemistry. Methods in Ecology and Evolution 6, 806-816. [In English].

Sturrock, A.M., Trueman, C.N., Darnaude, A.M., Hunter, E. (2012) Can otolith elemental chemistry

retrospectively track migrations in fully marine fishes? Journal of Fish Biology 81, 766-795.

Swearer, S.E., Caselle, J.E., Lea, D.W., Warner, R.R. (1999) Larval retention and recruitment in an

island population of a coral-reef fish. Nature 402, 799-802. [In English].

Swearer, S.E., Forrester, G.E., Steele, M.A., Brooks, A.J., Lea, D.W. (2003) Spatio-temporal and

interspecific variation in otolith trace-elemental fingerprints in a temperate estuarine fish

assemblage. Estuarine Coastal and Shelf Science 56, 1111-1123.

Tagliabracci, V.S., Engel, J.L., Wen, J.Z., et al. (2012) Secreted kinase phosphorylates extracellular

proteins that regulate biomineralization. Science 336, 1150-1153. [In English].

Tanner, S.E., Reis-Santos, P., Vasconcelos, R.P., et al. (2013) Does otolith geochemistry record

ambient environmental conditions in a temperate tidal estuary? Journal of Experimental

Marine Biology and Ecology 441, 7-15.

Tanner, S.E., Reis-Santos, P., Vasconcelos, R.P., Franca, S., Thorrold, S.R., Cabral, H.N. (2012)

Otolith geochemistry discriminates among estuarine nursery areas of Solea solea and S.

senegalensis over time. Marine Ecology Progress Series 452, 193-203.

Tanner, S.E., Vasconcelos, R.P., Reis-Santos, P., Cabral, H.N., Thorrold, S.R. (2011) Spatial and

ontogenetic variability in the chemical composition of juvenile common sole (Solea solea)

otoliths. Estuarine Coastal and Shelf Science 91, 150-157.

240

Thomas, O.R.B., Ganio, K., Roberts, B.R., Swearer, S.E. (2017) Trace element-protein interactions in

endolymph from the inner ear of fish: implications for environmental reconstructions

using fish otolith chemistry. Metallomics 9, 239-249.

Thomas, O.R.B., Swearer, S.E., Kapp, E.A., et al. (2019) The inner ear proteome of fish. FEBS Journal

286, 66-81.

Thorrold, S.R., Jones, C.M., Campana, S.E. (1997) Response to otolith microchemistry to

environmental variations experienced by larval and juvenile Atlantic croaker

(Micropogonias undulatus). Limnology and Oceanography 42, 102-111.

Thorrold, S.R., Jones, C.M., Campana, S.E., McLaren, J.W., Lam, J.W.H. (1998a) Trace element

signatures in otoliths record natal river of juvenile American shad (Alosa sapidissima).

Limnology and Oceanography 43, 1826-1835.

Thorrold, S.R., Jones, C.M., Swart, P.K., Targett, T.E. (1998b) Accurate classification of juvenile

weakfish Cynoscion regalis to estuarine nursery areas based on chemical signatures in

otoliths. Marine Ecology Progress Series 173, 253-265.

Thorrold, S.R., Latkoczy, C., Swart, P.K., Jones, C.M. (2001) Natal homing in a marine fish

metapopulation. Science 291, 297-299.

Thorrold, S.R., Swearer, S.E. (2009) Otolith Chemistry. In: Tropical Fish Otoliths: Information for

Assessment, Management and Ecology. Vol. 11. Reviews-Methods and Technologies in Fish

Biology and Fisheries (Eds. B.S. Green, B.D. Mapstone, G. Carlos, G.A. Begg), pp. 249-295.

Tohse, H., Ando, H., Mugiya, Y. (2004) Biochemical properties and immunohistochemical

localization of carbonic anhydrase in the sacculus of the inner ear in the salmon

Oncorhynchus masou. Comparative Biochemistry and Physiology a-Molecular & Integrative

Physiology 137, 87-94.

241

Tournois, J., Ferraton, F., Velez, L., et al. (2013) Temporal stability of otolith elemental fingerprints

discriminates among lagoon nursery habitats. Estuarine Coastal and Shelf Science 131,

182-193.

Tzeng, W.N., Tsai, Y.C. (1994) Changes in otolith microchemistry of the Japanese eel, Anguilla

japonica, during its migration from the ocean to the rivers of Taiwan. Journal of Fish

Biology 45, 671-683. [In English].

Vasconcelos, R.P., Reis-Santos, P., Tanner, S., et al. (2007) Discriminating estuarine nurseries for

five fish species through otolith elemental fingerprints. Marine Ecology Progress Series

350, 117-126.

Walther, B.D., Thorrold, S.R. (2006) Water, not food, contributes the majority of strontium and

barium deposited in the otoliths of a marine fish. Marine Ecology Progress Series 311, 125-

130.

Williams, J., Hindell, J.S., Swearer, S.E., Jenkins, G.P. (2012) Influence of freshwater flows on the

distribution of eggs and larvae of black bream Acanthopagrus butcheri within a drought-

affected estuary. Journal of Fish Biology 80, 2281-2301. [In English].

Williams, J., Jenkins, G.P., Hindell, J.S., Swearer, S.E. (2013) Linking environmental flows with the

distribution of black bream Acanthopagrus butcheri eggs, larvae and prey in a drought

affected estuary. Marine Ecology Progress Series 483, 273-287. [In English].

Williams, J., Jenkins, G.P., Hindell, J.S., Swearer, S.E. (2018) Fine-scale variability in elemental

composition of estuarine water and otoliths: Developing environmental markers for

determining larval fish dispersal histories within estuaries. Limnology and Oceanography

63, 262-277. [In English].

242

Wojtas, M., Dobryszycki, P. (2011) Hyperphosphorylation of intrinsically disordered Starmaker

protein increases its biomineralization activity. FEBS Journal 278, 482-482.

Wojtas, M., Wolcyrz, M., Ozyhar, A., Dobryszycki, P. (2012) Phosphorylation of intrinsically

disordered Starmaker protein increases its ability to control the formation of calcium

carbonate crystals. Crystal Growth & Design 12, 158-168.

Woodland, R.J., Warry, F., Zhu, Y., et al. (Unpublished) The factors affecting recruitment of

estuarine fish and management under increased water harvesting and climate change.

Zimmerman, C.E., Reeves, G.H. (2000) Population structure of sympatric anadromous and

nonanadromous Oncorhynchus mykiss: evidence from spawning surveys and otolith

microchemistry. Canadian Journal of Fisheries and Aquatic Sciences 57, 2152-2162. [In

English].

243

244

CHAPTER SIX: GENERAL DISCUSSION

An understanding of the biochemical mechanisms of otolith biomineralization is crucial to proper interpretation of otolith increment data in fish and fisheries ecology. For example, increment measurements are essential to ageing and stock assessment studies (Campana, 2001), but rely on a mechanism of daily accretion that is poorly understood. We similarly lack understanding as to the reasons behind the apparent metabolic inertness of the otolith (Campana and Neilson, 1985,

Campana, 1999), the supply of its inorganic components as well as the observation that sagittal otoliths are often aragonite (rather than the more stable calcite) (Carlstrom, 1963, Degens et al.,

1969, Mann et al., 1983, Oliveira et al., 1996). In addition to this, the specifics of trace element- otolith protein interactions are unclear, raising questions as to whether a given element is indicative of an environmental or a physiological process. In this thesis, I undertook several studies to clarify and elucidate the mechanisms that underpin the assumptions made in otolith ecological research.

I began by an assessment of current state of knowledge with regards to otolith biochemistry. In

Chapter Two, I assessed over 9000 studies, and found that relatively little was known about otolith proteins. This allowed me to identify present knowledge gaps, which were considerable in comparison to other biominerals. The small amount of information about otolith proteins was largely due to the technical limitations faced by early otolith protein researchers. Quite simply, the small number of otolith proteins identified in previous studies was a direct consequence of the authors not having access to more modern proteomic techniques such as shotgun mass spectrometry. I noted that more recent proteomic studies were either restricted to sequenced fish

(Danio rerio, Oncorhynchus spp., Oryzias latipes) or made use of large non-specific peptide databases. For example, in the most recent proteomic analysis of otoliths conducted by Weigele et al. (2016), a homolog of a key protein to otolith biomineralization, Starmaker (Söllner et al., 2003),

245

was not found, leading the authors to conclude that the proteins present in the otolith organic matrix varied considerably between taxa. However, Starmaker (and its functional homologs,

Starmaker-like (Bajoghli et al., 2009) and Otolith Matrix Macromolecule-64 (Tohse et al., 2008)) is an intrinsically disordered protein (Kaplon et al., 2008). This means that although their functions

(i.e. control of crystal of polymorph and regulation of crystallization) are the same, their primary sequences can be markedly different (Radivojac et al., 2007). A consequence of this is that, if one undertakes proteomic analyses by matching peptides against databases comprised of the entirety of Actinopterygii proteins or a distantly related species such as D. rerio, the homolog will not be found (contrast Rozycka et al. (2014) with Weigele et al. (2016)). In addition to this, I found that few studies, if any, had analysed both otoliths and endolymph from the same individual fish. It is important to analyse endolymph in conjunction with otoliths, as the latter are continually bathed in the former (Payan et al., 2004). This means that as increments form, small amounts of inner ear fluid will be trapped. This has three important implications for otolith studies. Firstly, in microchemistry studies, trace element-protein interactions are relevant not solely for otolith proteins, but also for endolymph proteins. This has implications for laser ablation-inductively coupled plasma-mass spectrometry (LA-ICP-MS) based environmental reconstructions, as detection of a trace element in the otolith may be indicative of endogenous, rather than exogenous, processes (Sturrock et al., 2014). Secondly, in proteomics, analysis of endolymph can aid in distinguishing between otolith proteins that are playing a functional role, and those that are the result of internal endolymph contamination (e.g. transferrin, Weigele et al. (2016)). Finally, and excitingly, the likelihood that endolymph is trapped during increment formation means that the otolith is potentially archiving physiological information as well as environmental data. This means that a spatio-temporal analysis of otolith proteins (as described below) may allow for the

246

reconstruction of additional life-history events (e.g. sexual maturation, spawning) that are not recorded by changes in otolith elemental chemistry.

In Chapter Three, I determined the trace-element protein interactions for a range of elements via size exclusion chromatography-inductively coupled plasma-mass spectrometry (SEC-ICP-MS) for adult black bream (Acanthopagrus butcheri) endolymph. The aim of this study was to comment on the utility of trace elements typically used in LA-ICP-MS based environmental reconstructions.

There have been previous studies into trace element: protein interactions in the inner ear, but these studies – Miller et al. (2006), Izzo et al. (2016) – were focused on the otolith, and so potentially risked the decoupling of trace elements from proteins through sample preparation techniques (i.e. metal chelation during EDTA dissolution). In my study, I found that there were three categories of element present in the endolymph: (1) elements solely present as free ions, (2) elements solely bound to large biomolecules (such as proteins) and (3) elements that were present in both bound and unbound forms. For those elements that are either exclusively or predominantly present as free ions, I suggest that they are appropriate indicators of exogenous processes. However, when undertaking LA-ICP-MS based discriminant analyses, it is imperative that a given element’s mechanism of incorporation is considered before its inclusion. For free ions, the most likely mechanisms are through either random trapping in crystal interstitial spaces or direct substitution for calcium. In order for random trapping to occur, the element must be small enough to get caught in the aragonite interstices. Consequently, smaller ions such as Li will be favoured over larger ones such as Rb. Direct substitution for calcium is likely to occur as part of

Ostwald ripening – the dissolution of small crystals and their subsequent re-precipitation during the course of normal crystal growth (Vetter et al., 2013). Free ions with the same charge (+2) as calcium are candidates for substitution. It has been found that elements with a larger ionic radius than calcium’s are favoured in otolith incorporation (Melancon et al., 2005). Notably, the most

247

utilized elements in otolith chemistry, Sr and Ba, were found to be interacting, to a small extent, with the same molecular weight fraction as calcium, indicating that they are potentially misincorporating into calcium-binding proteins. This requires further investigation, as the mechanism of Sr and Ba supply to the endolymph from the saccular epithelial cells is still largely unknown. It may be that this is evidence that Sr and Ba enter the endolymph via the same route as

Ca – through interaction with active transport proteins such as Plasma membrane Ca2+-ATPase

(PMCA, Cruz et al. (2009)).

I also found that several elements were solely interacting with biomolecules, either as co-factors

(e.g. Cu, Zn, Fe), through electrostatic surface adhesion (e.g. Pb, As), or as a post-translational modification (e.g. phosphorylation, P) (Silva and Williams, 2001). Of the proteinaceous elements,

P, Cu, Zn and Pb are of most interest as they are often used in multi-elemental fingerprinting of otoliths. As these elements are likely to be biologically regulated (i.e. as a consequence of protein interactions), I caution against their use as environmental markers as it is more likely that their presence in the otolith are indicative of endogeny, rather than exogeny. Although this study provided insight into which trace elements were interacting with proteins in the endolymph, the identities of those metalloproteins were still unclear.

In order to determine which metalloproteins were present in the otolith, I undertook a comprehensive proteomic analysis of adult black bream otoliths (Chapter Four). As black bream has not had its genome sequenced (like most fish), this necessitated the creation of the transcriptome of its brain and inner ear – a valuable resource not only for this project, but for future ones as well. Through my uncovering of the black bream inner ear proteome, I was not only able to increase the number of identified otolith proteins by a factor of 30, but also able to form several hypotheses relating to, until now, poorly understood aspects of otolith growth.

Although I only touch on a few dozen of the proteins identified within the otolith and endolymph, I

248

found, at the quantitative level, nearly a thousand inner ear proteins. It is likely that only a few of these proteins are actively involved in biomineralization pathways whereas others (e.g. vitellogenin) are incorporated inadvertently during increment formation. The presence of non- biomineralization proteins in the otolith is intriguing and suggests that otoliths are not only archives of physico-chemical data, but of physiological information as well. If the positions of those proteins within the otolith can be maintained during analysis, this will allow, excitingly, for the prospect of reconstruction of endogenous life-history events.

I identified several of the saccular transmembrane proteins involved in inorganic constituent supply to the endolymph as well as two separate isozymes of Carbonic anhydrase. One of these

(CA 6) was exclusively in otolith protein samples, whereas the other (CA 1) was restricted to the endolymph. I also found the presence of a pH pump (V-ATPase), which is thought to act in tandem with CA 6 in tooth enamel to regulate pH and bicarbonate (Yin and Paine, 2017). This suggests two things of interest to otolith researchers: Firstly, that there is a great deal of conservation across biomineral systems. Indeed, a comparison of otolith proteins of A. butcheri with pig (Sus domesticus) enamel matrix proteins (D. Green, personal communication) as well as shell matrix proteins of pearl oysters (Pinctada fucata, Liu et al. (2015)) revealed considerable overlap between these three systems. Secondly, the presence of CA 6 in otolith tissue (but not in the endolymph) suggests that bicarbonate supply is being directly regulated during otolith growth – an area that will benefit from further exploration.

I also determined the primary sequence of a Starmaker homolog. This would not have been possible without the creation of black bream’s inner ear and brain transcriptome. Starmaker (and its related proteins in biomineralization, the SIBLING family) is an intrinsically disordered protein

(Kaplon et al., 2008) and will thus evade detection through alignment with generic peptide databases. The creation of a transcriptome also allowed for the identification of Extracellular

249

serine/threonine protein kinase FAM20C (FAM20C), a protein known to phosphorylate SIBLING proteins (Tagliabracci et al., 2013). By discovery of FAM20C and its potential substrate (Starmaker- like), coupled with the observed enrichment of Mn in otolith cores (Brophy et al., 2004,

Ruttenberg et al., 2005), I was able to develop hypotheses describing the mechanisms of aragonite polymorph selection, nucleation and regulation of crystal growth. It is imperative that we understand the intricacies of polymorph selection, as choice in polymorph (i.e. calcite, vaterite or aragonite) likely affects rate of trace metal incorporation (Melancon et al., 2005, Macdonald et al.,

2012), impacting upon physico-chemical reconstructions. Additionally, as vaterite has been linked with hearing impairment in farmed fish (Reimer et al., 2016), an understanding of the protein- mediation in polymorph selection may aid in finding a solution to this potential welfare issue in farmed fish.

In addition to this, the discovery of Matrix metalloproteinase 2 in otoliths (along with Neuroserpin and its downstream targets) allowed me to create a working model for the mechanism behind diel-entrained increment deposition. Although I produced a theoretical model, the timing of expression of these proteins can be determined through genetic manipulation and immunohistochemistry studies as discussed below.

In Chapter Five, I used the findings from the two biochemical studies (Chapters Three and Four) to inform the design for a project focused on the utility of multi-elemental clustering in environmental reconstructions from otolith chemistry. In this study, I investigated the utility of different trace elements in environmental reconstructions by comparing three different cluster analysis approaches with regards to their ability to resolve spawning and larval nursery habitats for black bream in the Gippsland Lakes, Australia. The three approaches taken were largely based on my data in Chapter Three. I clustered based on: (1) all elements analysed, (2) salt fraction- dominant elements, and (3) protein-fraction dominant elements. My comparison suggested that

250

choosing to cluster based on “salt-fraction dominant” elements (i.e. Groups 1, 2, and Mn) provided the best ability to predict cohort and spatial differences amongst individuals.

Interestingly, I found a strong correlation between Mg and P in all samples (P=0.8213, n=276).

Limburg et al. (2018) have suggested the potential utility of Mg as an indicator of metabolic rate.

Mg is known to bind ATP (Silva and Williams, 2001) and ATP (along with a SIBLING protein) is a necessary substrate for the kinase, FAM20C (Tagliabracci et al., 2012, Tagliabracci et al., 2013). The investigation of the roles of Mg, ATP and FAM20C in otolith growth is a potential fruitful avenue of research.

My thesis has allowed for the creation of several hypotheses relating to otolith formation and growth. It is necessary, however, to test many of these hypotheses. Below, I list several current and future research projects that will continue to elucidate otolith biochemistry, and more broadly, biomineralization mechanisms. In addition to this, as I created a transcriptome of the brain and inner ear of black bream, I was also able to identify potential proteins of interest in fish eye lenses, which have an analogous pattern of growth to otoliths and may be of interest to fish ecologists in the reconstruction of dietary histories.

FUTURE DIRECTIONS

In order to test several of the hypotheses developed in this thesis, it is necessary to undertake both immunohistochemical and genetic manipulation studies on the fish inner ear. Through labelling, we can confirm the expression of key biomineralization proteins such as FAM20C,

Plasminogen, Sparc, Neuroserpin, and MMP 2 in the fish inner ear. In collaboration with developmental geneticists at the University of Melbourne, I successfully labelled MMP 2, showing that it is almost exclusively expressed in the macula (Fig. 6.1). I envisage a larger study, wherein

251

larvae are killed periodically (e.g. every 2h post fertilisation) and subsequently labelled for proteins of interest. This study will help confirm FAM20C’s location in the primordium, as well as validate my proposed mechanism for the circadian regulation of increment deposition. If indeed protein expression is being regulated by a clock gene (i.e. PER2, Su et al. (2017)), this will be evident through a difference in intensity of its expression over a 24 h period.

Brain Brain

200 µm 200 µm

200 µm 200 µm

Figure 6.1. Fluorescence micrograph of Matrix metalloproteinase 2 (MMP 2) labelled tissues in coronal section of the inner ear of larval zebrafish (Danio rerio). Red colour indicates intensity of protein expression, and that MMP 2 (red) is primarily expressed in the saccular macula of D. rerio (n=3). 4′,6-diamidino-2- phenylindole (DAPI) is employed here as a nuclear stain (blue) for contrast.

Furthermore, genetic manipulation of a model organism (e.g. D. rerio, Salmo salar) will shed further light on the mechanisms of polymorph selection. If it is indeed the result of post- translational modification (i.e. phosphorylation) of Starmaker by FAM20C, then a change in the expression of the latter should result in aberrant otoliths. To determine this, I intend to knock out

252

FAM20C in developing fish (zebrafish, Atlantic salmon) and characterize the crystal polymorph of their otoliths both visually and via X-ray diffraction.

Another potential study is one into the abundances and relative post-translational modifications of proteins extracted from the sagitta and asteriscus of the same individual. As the asteriscus tends to be vaterite (Lowenstam, 1989, Oliveira et al., 1996), any differences between its proteins and that of the aragonitic sagitta will provide further useful clues as to the specifics of protein- mediated polymorph selection. Proteomics, genetic manipulation and immunohistochemical studies have the potential to reveal a great deal more about the structures, abundances and timings of expression of otolith proteins. The specific locations of both biomineralization and non- biomineralization proteins, however, continue to prove to be elusive.

Recall that spatio-temporal analysis of otolith proteins is difficult as they are sandwiched between mineral layers. Due to this, it is not possible to analyse the proteins in situ as the mineral phase can interfere with typical spatio-temporal protein analysis approaches, such as matrix assisted laser desorption ionization-imaging mass spectrometry (MALDI-IMS) (Fujino et al., 2016). Similarly, extraction of proteins through acid or EDTA dissolution results in the irreversible destruction of any spatio-temporal data.

However, there is a protocol, “CLARITY”, that has been used to successfully 3D image the proteins in mice brains (Tomer et al., 2014). In a pilot study, I adapted this approach in a novel way to remove the otolith mineral phase, while maintaining the position and integrity of proteins. This now allows me to apply a variety of staining and imaging techniques (e.g. immunohistochemistry and confocal microscopy, MALDI-IMS) following “CLARITY” to identify the specific locations of proteins in the otolith (Fig. 6.2). With a CLARITY-treated otolith, we can determine the locations of biomineralization proteins identified in Chapter Four (e.g. Otolith matrix protein-1, Otolin-1,

253

FAM20C, Sparc, Neuroserpin, MMP 2 etc.), further aiding our understanding of biomineralization mechanisms. In addition to this, I will also be able to determine the locations of non- biomineralization proteins (e.g. vitellogenin) and other biomolecules (e.g. steroids such as cortisol, metabolites etc.), remarkably allowing for the reconstruction of key physiological events such as maturation or stress.

Figure 6.2. Z-stack of CLARITY-treated otolith from adult black bream, Acanthopagrus butcheri. Oriole staining indicates presence of vertically radiating collagen fibres (bright white) through the otolith and the presence of ~horizontal increments (alternating light and dark bands).

254

Future proteomic approaches need not be restricted to the inner ear. In a separate pilot study, I investigated the relative abundances of metabolites and proteins occurring in the eye lens of juvenile black bream. Using the transcriptome that I had created previously, I identified proteins of potential interest (Crystallin alpha, Grey and Schey (2008); Aquaporin 0, Schey et al. (2013)) and metabolites known to occur in vertebrate eye lenses (Grey, 2016). I then, with my collaborators, used MALDI-IMS to map the distributions of proteins and biomolecules of interest. Vertebrate eye lenses grow continually from in embryo (Bloemendal, 1977). As lens biomolecules have been shown to undergo age-related modification (Korlimbinis et al., 2009, Grey, 2016, Truscott and

Friedrich, 2016), they may potentially be the key to determining changes in trophic ecology through stable isotope ratios (Trueman et al., 2012, Gronkjaer et al., 2013). In our pilot study, it was possible to identify temporally distinct regions within the fish eye lens (Fig. 6.3). Specifically, we were able to detect differences in molecular species between the lens nucleus, and layers of the lens cortex. As these regions are known to form at different life stages, we can use the change in molecular species to construct a biochemical age-gradient through the fish lens. This provides the temporal resolution that, when coupled with determination of stable isotope ratios, allows for the reconstruction of an individual’s lifetime dietary history.

255

Figure 6.3. MALDI images of major tissues of the juvenile black bream eye. Representative MALDI images from the (A) choroid (m/z 834.5264), (B) retina (m/z 401.0553), (C, D) lens cortex (m/z 438.1429, m/z 503.1608) lens nucleus (m/z 902.4642), and multiple ocular tissues (m/z 426.0256). Scale bar = 500um.

256

A final major outcome from my thesis was the development of a workflow for effective assembly of species-specific protein landscapes. As noted above, successful protein inference relies upon alignment with sequenced genomes and proteins in public domain databases. This proves to be a significant challenge for the study of non-model organisms for several reasons. Firstly, as such alignments are typically done against large, non-specific datasets (e.g. Actinopterygii, Vertebrata), the power of these searches is inevitably negatively impacted upon. Secondly, the false discovery rate for peptides is increased. Finally, large numbers of mass spectra end up being unassigned, causing numerous novel proteins to be overlooked (Choudhary et al., 2001). To avoid this, proteomic approaches will typically focus on sequenced organisms (i.e. D. rerio, Mus musculus).

This means that although we are deepening our knowledge of the genomes and proteomes of select species, we are not broadening our knowledge of overall biochemical systems. Accordingly, we developed a method for creating a species-specific peptide database that can now be easily adapted to other biomineral systems (such as cephalopod statoliths and shark vertebrae). By mapping these structures’ proteomes, and subsequently comparing them with otoliths (and other biomineral proteomes) we can greatly enhance our general understanding of biomineralization mechanisms.

The otolith has provided a cornucopia of information to fish ecologists, allowing for the reconstruction of age and growth chronologies as well as environmental histories. By embracing the techniques and methodologies afforded by biochemistry, genetics and molecular biology, future scientists will continue to plumb its secrets. While the otolith has already revealed so much, it is abundantly clear that it has still far more to tell.

257

REFERENCES

Bajoghli, B., Ramialison, M., Aghaallaei, N., Czerny, T., Wittbrodt, J. (2009) Identification of

Starmaker-like in medaka as a putative target gene of Pax2 in the otic vesicle.

Developmental Dynamics 238, 2860-2866.

Bloemendal, H. (1977) The vertebrate eye lens. Science 197, 127-138. [In English].

Brophy, D., Jeffries, T.E., Danilowicz, B.S. (2004) Elevated manganese concentrations at the cores

of clupeid otoliths: possible environmental, physiological, or structural origins. Marine

Biology 144, 779-786.

Campana, S.E. (1999) Chemistry and composition of fish otoliths: pathways, mechanisms and

applications. Marine Ecology Progress Series 188, 263-297.

Campana, S.E. (2001) Accuracy, precision and quality control in age determination, including a

review of the use and abuse of age validation methods. Journal of Fish Biology 59, 197-

242. [In English].

Campana, S.E., Neilson, J.D. (1985) Microstructure of fish otoliths. Canadian Journal of Fisheries

and Aquatic Sciences 42, 1014-1032. [In English].

Carlstrom, D.D. (1963) Crystallographic study of vertebrate otoliths. Biological Bulletin 125, 441-&.

[In English].

Choudhary, J.S., Blackstock, W.P., Creasy, D.M., Cottrell, J.S. (2001) Matching peptide mass spectra

to EST and genomic DNA databases. Trends in Biotechnology 19, S17-22.

258

Cruz, S., Shiao, J.-C., Liao, B.-K., Huang, C.-J., Hwang, P.-P. (2009) Plasma membrane calcium

ATPase required for semicircular canal formation and otolith growth in the zebrafish inner

ear. Journal of Experimental Biology 212, 639-647.

Degens, E.T., Deuser, W.G., Haedrich, R.L. (1969) Molecular structure and composition of fish

otoliths. Marine Biology 2, 105-113.

Fujino, Y., Minamizaki, T., Yoshioka, H., Okada, M., Yoshiko, Y. (2016) Imaging and mapping of

mouse bone using MALDI-imaging mass spectrometry. Bone Rep 5, 280-285.

Grey, A.C. (2016) MALDI imaging of the eye: Mapping lipid, protein and metabolite distributions in

aging and ocular disease. International Journal of Mass Spectrometry 401, 31-38.

Grey, A.C., Schey, K.L. (2008) Distribution of bovine and rabbit lens Alpha Crystallin products by

MALDI imaging mass spectrometry. Molecular Vision 14, 171-179.

Gronkjaer, P., Pedersen, J.B., Ankjaero, T.T., et al. (2013) Stable N and C isotopes in the organic

matrix of fish otoliths: validation of a new approach for studying spatial and temporal

changes in the trophic structure of aquatic ecosystems. Canadian Journal of Fisheries and

Aquatic Sciences 70, 143-146.

Izzo, C., Doubleday, Z.A., Gillanders, B.M. (2016) Where do elements bind within the otoliths of

fish? Marine and Freshwater Research 67, 1072-1076.

Kaplon, T.M., Rymarczyk, G., Nocula-Lugowska, M., et al. (2008) Starmaker exhibits properties of

an intrinsically disordered protein. Biomacromolecules 9, 2118-2125.

Korlimbinis, A., Berry, Y., Thibault, D., Schey, K.L., Truscott, R.J.W. (2009) Protein aging: truncation

of Aquaporin 0 in human lens regions is a continuous age-dependent process.

Experimental Eye Research 88, 966-973.

259

Limburg, K.E., Wuenschel, M.J., Hüssy, K., Heimbrand, Y., Samson, M. (2018) Making the otolith

magnesium chemical calendar-clock tick: Plausible mechanism and empirical evidence.

Reviews in Fisheries Science & Aquaculture 26, 479-493. [In English].

Liu, C., Li, S.G., Kong, J.J., et al. (2015) In-depth proteomic analysis of shell matrix proteins of

Pinctada fucata. Scientific Reports 5, 14. [In English].

Lowenstam, H.A. (1989) On biomineralization / Heinz A. Lowenstam, Stephen Weiner, Vol.

Accessed from https://nla.gov.au/nla.cat-vn1683340), Oxford University Press, New York.

Macdonald, J.I., McNeil, D.G., Crook, D.A. (2012) Asteriscus v. lapillus: comparing the chemistry of

two otolith types and their ability to delineate riverine populations of common carp

Cyprinus carpio. Journal of Fish Biology 81, 1715-1729.

Mann, S., Parker, S.B., Ross, M.D., Skarnulis, A.J., Williams, R.J.P. (1983) The ultrastructure of the

calcium carbonate balance organs of the inner-ear - an ultrahigh resolution electron-

microscopy study. Proceedings of the Royal Society Series B-Biological Sciences 218, 415-+.

[In English].

Melancon, S., Fryer, B.J., Ludsin, S.A., Gagnon, J.E., Yang, Z.P. (2005) Effects of crystal structure on

the uptake of metals by lake trout (Salvelinus namaycush) otoliths. Canadian Journal of

Fisheries and Aquatic Sciences 62, 2609-2619.

Miller, M.B., Clough, A.M., Batson, J.N., Vachet, R.W. (2006) Transition metal binding to cod otolith

proteins. Journal of Experimental Marine Biology and Ecology 329, 135-143.

Oliveira, A.M., Farina, M., Ludka, I.P., Kachar, B. (1996) Vaterite, calcite, and aragonite in the

otoliths of three species of piranha. Naturwissenschaften 83, 133-135.

260

Payan, P., De Pontual, H., Boeuf, G., Mayer-Gostan, N. (2004) Endolymph chemistry and otolith

growth in fish. Comptes Rendus Palevol 3, 535-547.

Radivojac, P., Iakoucheva, L.M., Oldfield, C.J., Obradovic, Z., Uversky, V.N., Dunker, A.K. (2007)

Intrinsic disorder and functional proteomics. Biophysical Journal 92, 1439-1456.

Reimer, T., Dempster, T., Warren-Myers, F., Jensen, A.J., Swearer, S.E. (2016) High prevalence of

vaterite in sagittal otoliths causes hearing impairment in farmed fish. Scientific Reports 6,

25249. [In English].

Rozycka, M., Wojtas, M., Jakob, M., et al. (2014) Intrinsically disordered and pliable Starmaker-like

protein from medaka (Oryzias latipes) controls the formation of calcium carbonate

crystals. PloS One 9, e114308.

Ruttenberg, B.I., Hamilton, S.L., Hickford, M.J.H., et al. (2005) Elevated levels of trace elements in

cores of otoliths and their potential for use as natural tags. Marine Ecology Progress Series

297, 273-281.

Schey, K.L., Grey, A.C., Nicklay, J.J. (2013) Mass spectrometry of membrane proteins: a focus on

aquaporins. Biochemistry 52, 3807-3817.

Silva, J.J.R.F.d., Williams, R.J.P. (2001) The biological chemistry of the elements: the inorganic

chemistry of life, 2nd edn Vol., Oxford University Press, Oxford, UK.

Söllner, C., Burghammer, M., Busch-Nentwich, E., et al. (2003) Control of crystal size and lattice

formation by Starmaker in otolith biomineralization. Science 302, 282-286.

Sturrock, A.M., Trueman, C.N., Milton, J.A., Waring, C.P., Cooper, M.J., Hunter, E. (2014)

Physiological influences can outweigh environmental signals in otolith microchemistry

research. Marine Ecology Progress Series 500, 245-264.

261

Su, X., Chen, D., Yang, K., et al. (2017) The circadian clock gene PER2 plays an important role in

tumor suppression through regulating tumor-associated genes in human oral squamous

cell carcinoma. Oncology Reports 38, 472-480.

Tagliabracci, V.S., Engel, J.L., Wen, J.Z., et al. (2012) Secreted kinase phosphorylates extracellular

proteins that regulate biomineralization. Science 336, 1150-1153. [In English].

Tagliabracci, V.S., Pinna, L.A., Dixon, J.E. (2013) Secreted protein kinases. Trends in Biochemical

Sciences 38, 121-130. [In English].

Tohse, H., Takagi, Y., Nagasawa, H. (2008) Identification of a novel matrix protein contained in a

protein aggregate associated with collagen in fish otoliths. FEBS Journal 275, 2512-2523.

Tomer, R., Ye, L., Hsueh, B., Deisseroth, K. (2014) Advanced CLARITY for rapid and high-resolution

imaging of intact tissues. Nature Protocols 9, 1682-1697. [In English].

Trueman, C.N., MacKenzie, K.M., Palmer, M.R. (2012) Stable isotopes reveal linkages between

ocean climate, plankton community dynamics, and survival of two populations of Atlantic

salmon (Salmo salar). ICES Journal of Marine Science 69, 784-794. [In English].

Truscott, R.J., Friedrich, M.G. (2016) The etiology of human age-related cataract. Proteins don't

last forever. Biochimica et Biophysica Acta 1860, 192-198. [In eng].

Vetter, T., Iggland, M., Ochsenbein, D., Haenseler, F., Mazzotti, M., Hänseler, F. (2013) Modeling

nucleation, growth, and Ostwald ripening in crystallization processes: A comparison

between population balance and kinetic rate equation. Crystal Growth & Design 13, 4890-

4905.

Weigele, J., Franz-Odendaal, T.A., Hilbig, R. (2016) Not all inner ears are the same: Otolith matrix

proteins in the inner ear of sub-adult cichlid fish, Oreochromis mossambicus, reveal

262

insights into the biomineralization process. Anatomical Record-Advances in Integrative

Anatomy and Evolutionary Biology 299, 234-245.

Yin, K., Paine, M.L. (2017) Bicarbonate transport during enamel maturation. Calcified Tissue

International 101, 457-464.

263

264

APPENDICES

Appendix 3.1. Thomas, O.R.B., Ganio, K., Roberts, B.R., Swearer, S.E. (2017) Trace element-protein interactions in endolymph from the inner ear of fish: implications for environmental reconstructions using fish otolith chemistry. Metallomics 9, 239-249

265

266

267

268

269

270

271

272

273

274

275

276

Appendix 3.2. Size exclusion chromatograms of 7Li, 11B, 24Mg, 31P, 39K, 44Ca, 51V, 52Cr, 55Mn, 56Fe, 59Co, 60Ni, 63Cu, 66Zn, 75As, 85Rb, 88Sr, 95Mo, 111Cd, 118Sn, 127I, 137Ba, 201Hg and 208Pb in endolymph of adult black bream (Acanthopagrus butcheri), n=3.

Li B

400 200

d

d n

300 n 150

o

o

c

c

e

e

s

s

200

r 100

r

e

e

p

p

s s t 100

t 50

n

n

u

u o 0 o C 0 2 4 6 8 C 2 4 6 8 Volume (mL) -100 -50 Volume (mL) Mg P

150000 25000 d

d 20000

n

n

o o

100000 c e

c 15000

s

e

r

s

e

r 10000

p e

50000

s

p

t s

n 5000

t

u

n

o u C 0 o 0 2 4 6 8 C 2 4 6 8 -5000 Volume (mL) Volume (mL) -50000 K Ca

2.510 7 40000

d d n

n 2.010 7 o

o 30000

c c

e e s

s 7

1.510

r r e

e 20000

p p

7

s s t

t 1.010

n n

u u o o 10000

5.010 6

C C

0 0 0 2 4 6 8 0 2 4 6 8 Volume (mL) Volume (mL)

277

V Cr

250 3000

d d n

n 200

o o

c c e

e 2000

s s

150

r r

e e

p p

s s t

t 100 n

n 1000

u u o

o 50

C C

0 0 0 2 4 6 8 0 2 4 6 8 Volume (mL) Volume (mL)

Mn Fe

6000 400000

d d

n n o

o 300000

c c e

e 4000

s s

r r e

e 200000

p p

s s

t t n

n 2000

u u o

o 100000

C C

0 0 0 2 4 6 8 0 2 4 6 8 Volume (mL) Volume (mL)

Co Ni

15000 60000

d d

n n

o o

c c e

e 10000 40000

s s

r r

e e

p p

s s

t t n

n 5000 20000

u u

o o

C C

0 0 0 2 4 6 8 0 2 4 6 8 Volume (mL) Volume (mL)

278

Cu Zn

200000 80000

d

n

d o 150000 n

c 60000

o

e

c

s

e

r

s e

100000 40000

p

r

e

s

t

p

n

s u

t 20000

o 50000

n

C u o 0 0 C 0 2 4 6 8 2 4 6 8 Volume (mL) -20000 Volume (mL)

As Rb

1500 20000

d

d

n n

o 15000 o

1000 c

e

c

s

e

r

s

e

r 10000

p

e

500 s

p

t

n

s

t

u n

o 5000

u C

o 0

C 2 4 6 8 0 Volume (mL) 0 2 4 6 8 -500 Volume (mL)

Sr Mo

100000 6000

d

d n

80000 n 5500

o

o

c

c

e

e

s

s

60000 5000

r

r

e

e

p

p

s

s t

40000 t 4500

n

n

u

u o

20000 o 4000

C C

0 3500 0 2 4 6 8 0 2 4 6 8 Volume (mL) Volume (mL)

279

Cd Sn

150 300

d d

n n o

o 100 200

c c

e e

s s

r r e

e 50 100

p p

s s

t t

n n

u u o

o 0 0 C C 2 4 6 8 2 4 6 8 Volume (mL) Volume (mL) -50 -100 I Ba

200000 10000

d d n

n 8000 o

o 150000

c c

e e

s s

6000

r r e

e 100000

p p

s s t

t 4000

n n

u u o o 50000

2000

C C

0 0 0 2 4 6 8 0 2 4 6 8 Volume (mL) Volume (mL)

Hg Pb

80 2500

d n

d 2000

o n

60 c

o

e

c

s

e 1500

r

s e

40

p

r

e s

t 1000

p

n

s u

t 20 o

n 500

C u o 0 C 0 2 4 6 8 0 2 4 6 8 -20 Volume (mL) Volume (mL)

280

Appendix 4.1 Supplementary material for Thomas et al., 2018. FEBS Journal (In press). Please note that due to the size of these databases, they have not been reprinted. They can be accessed at: https://doi.org/10.1111/febs.14715

Custom peptide library, Sparidae_trans_v7.fasta

Peptide library created for Mascot searches of peptides extracted from otoliths and endolymph from black bream (Acanthopagrus butcheri). The flowchart for library creation is summarized in Figure 4.1.

Full peptide list from quantitative proteomics.xlsx Identified peptides in endolymph and otoliths from adult black bream (Acanthopagrus butcheri, n=3). Endolymph sample 2 was split into two subsamples, 2a and 2b, due to its large volume.

Error_Tolerant Search Results.xlsx Results of error tolerant Mascot search of peptides from endolymph and otoliths from adult black bream (Acanthopagrus butcheri, n=3), indicating potential post translational modifications such as phosphorylation or glycosylation. Endolymph sample 2 was split into two subsamples, 2a and 2b, due to its large volume.

281

Appendix 5.1 Comparison of MANOVA and ANOVA results for different cluster analysis approaches using LA-ICP-MS data from sagittal otoliths extracted from juvenile black bream (Acanthopagrus butcheri).

Li B Mg P S K MANOVA ANOVA ANOVA ANOVA ANOVA ANOVA ANOVA Cluster Approach, 2015 F P F p F p F P F p F p F P Larval period All 6.390 < 2.2e-16 8.891 1.415e-09 13.387 2.479e-13 12.022 2.845e-12 11.338 1.024e-11 6.844 1.417e-07 3.596 0.0006 n = 89 Salt 20.284 < 2.2e-16 12.978 2.482e-09 34.125 < 2.2e-1 23.884 8.356e-15 4.686 0.0008

Protein 6.222 3.762e-14 6.065 0.0002 1.867 0.1238

Primordium All 7.083 6.047e-13 17.998 3.613e-07 6.459 0.0014 22.297 3.66e-08 19.745 1.373e-07 2.999 0.0441 4.000 0.0153 n = 38 Salt 11.752 7.015e-12 0.733 0.4878 5.468 0.0086 57.217 9.297e-12 6.4391 0.0042

Protein 11.350 1.404e-11 19.545 1.997e-06 14.755 2.253e-05

Mn Cu Zn Sr Ba Pb ANOVA ANOVA ANOVA ANOVA ANOVA ANOVA Cluster Approach, 2015 F p F p F p F P F p F P Larval period All 6.9611 1.075e-07 5.114 1.026e-05 10.668 3.728e-11 24.908 < 2.2e-16 40.679 < 2.2e-16 17.302 4.653e-16 n = 89 Salt 16.246 3.802e-11 41.386 < 2.2e-16 103.570 < 2.2e-16

Protein 4.229 0.0036 87.526 < 2.2e-16 6.131 0.0002

Primordium All 4.134 0.0133 8.675 0.0002 5.853 0.0025 11.670 2.064e-05 2.604 0.0678 1.986 0.1347 n = 38 Salt 23.530 3.341e-07 5.677 0.0073

Protein 4.194 0.0233 8.332 0.0011 25.429 1.514e-07 0.969 0.3893

282

Li B Mg P S K MANOVA ANOVA ANOVA ANOVA ANOVA ANOVA ANOVA Cluster F P F p F p F P F p F P F P Approach, 2016 Larval period All 8.808 < 2.2e-16 12.425 < 2.2e-16 16.895 < 2.2e-16 10.066 7.797e-15 10.942 4.709e-16 24.959 < 2.2e-16 5.373 1.044e-07 n = 186 Salt 24.273 < 2.2e-16 32.485 < 2.2e-16 47.509 < 2.2e-16 12.166 8.828e-09 3.165 0.0152

Protein 12.222 < 2.2e-16 12.328 6.907e-09 0.975 0.4225

Primordium All 7.001 < 2.2e-16 2.813 0.0081 20.103 < 2.2e-16 3.350 0.0022 4.570 0.0001 3.502 0.0015 21.435 < 2.2e-16 n = 96 Salt 24.141 < 2.2e-16 8.973 0.0003 24.057 3.797e-09 1.034 0.3596 80.273 < 2.2e-16

Protein 16.608 < 2.2e-16 8.807 4.782e-06 2.473 0.0500

Mn Cu Zn Sr Ba Pb ANOVA ANOVA ANOVA ANOVA ANOVA ANOVA Cluster F P F p F p F P F p F P Approach, 2016 Larval period All 18.826 < 2.2e-16 11.864 < 2.2e-16 13.858 < 2.2e-16 9.183 1.429e-13 16.002 < 2.2e-16 38.558 < 2.2e-16 n = 186 Salt 45.368 < 2.2e-16 29.438 < 2.2e-16 55.914 < 2.2e-16

Protein 8.160 4.516e-06 141.590 < 2.2e-16 12.247 7.809e-09

Primordium All 12.017 2.3e-11 16.429 1.673e-14 20.321 < 2.2e-16 4.903 5.229e-05 9.933 1.090e-09 23.057 < 2.2e-16 n = 96 Salt 17.735 2.988e-07 27.924 3.176e-10

Protein 42.567 < 2.2e-16 9.986 9.847e-07 48.380 < 2.2e-16 69.735 < 2.2e-16

283

Appendix 5.2 Comparison of clusters based on otolith chemistry for the larval period of juvenile black bream (Acanthopagrus butcheri) caught in Lake King, Australia (2015, n = 89; 2016, n = 186). Clusters were assigned using the “salt-dominant” approach. Error bars represent +/- SE.

2015 4

3

2

1

0

Standardized calcium to Standardized ratio -1

-2

-3 Li B Mg P S K Mn Cu Zn Sr Ba Pb

A B C D E F

284

2016 4

3

2

1

Standardized calcium to Standardized ratio 0

-1

-2 Li B Mg P S K Mn Cu Zn Sr Ba Pb

A B C D E

285

Minerva Access is the Institutional Repository of The University of Melbourne

Author/s: Thomas, OIiver Robert Bion

Title: Biochemical mechanisms of biomineralization and elemental incorporation in otoliths: implications for fish and fisheries research

Date: 2018

Persistent Link: http://hdl.handle.net/11343/221421

File Description: Complete Thesis

Terms and Conditions: Terms and Conditions: Copyright in works deposited in Minerva Access is retained by the copyright owner. The work may not be altered without permission from the copyright owner. Readers may only download, print and save electronic copies of whole works for their own personal non-commercial use. Any use that exceeds these limits requires permission from the copyright owner. Attribution is essential when quoting or paraphrasing from these works.