SlideShare a Scribd company logo
1 of 76
Download to read offline
Ellen Boylen
June 2015
This thesis is presented as part of the requirements for
the Degree of Bachelor of Science in Marine Science
with Honours at Murdoch University
Factors affecting the growth of the Foxfish
Bodianus frenchii on the south and lower west
coasts of Western Australia
1
Declaration
I declare that the work presented here is my own research conducted from March to
October 2014, and has not been submitted for the award of any other degree at another
tertiary institution.
Elle Boylen
June 2015
Cover photograph: "Curious Foxfish at Rottnest Island, Western Australia"
Taken by Ellen Boylen, 2014
2
Abstract
The Foxfish Bodianus frenchii is a long-lived labrid that is endemic to Western
Australia, where it lives in shallow coastal reefs on the lower west and cooler south
coasts. A hybrid of sclerochronological techniques, Pearsonā€™s correlation coefficient
and additive mixed modelling was used to determine whether the pattern of growth of
the otoliths of B. frenchii was related to temperature over the year, seasonally and or to
the warmer months of the year. Emphasis was placed on testing the hypothesis that the
strength of the relationships between otolith growth and temperature variables is greater
on the lower west than south coast as actual temperatures and temperature range in any
given month are greater on that coast and the temperatures varied more between years.
Focus was also placed on determining whether the pattern of otolith growth of B.
frenchii on the lower west coast was related to sea level height on that coast, an
indicator of the strength of the Leeuwin Current. No attempt was made to relate otolith
growth with sea level height on the south coast as the Leeuwin current flows well
offshore on that coast.
Successive growth increments in the transverse sections of the otoliths of 53
B. frenchii, comprising adjacent translucent and opaque zones, were measured and
aligned according to year, using the increment width for the year of capture as the
anchor. Visual examination of the time series of individual increment widths for all
otoliths of individuals for each coast separately was undertaken to determine whether
the widths of the increments, i.e. narrow or wide, of each otolith corresponded. The
widths of the sequential increments of the different otoliths did not match, a common
feature of otoliths in which all increment widths are very narrow. Further attempts to
cross date the increment widths in the time series for each otolith, i.e. aimed at matching
corresponding increment widths, was thus undertaken using a statistical approach. This
3
was also not entirely successful. Subsequent analyses thus continued to use the year of
capture as the anchor aligning the otolith width of the different years for individuals on
both coasts.
The next step was detrending, whereby age-related declines in otolith growth are
removed, while at the same time preserving any climate signals present in the increment
width time series. Various detrending techniques have been used in otolith-based
increment width studies. A visual assessment of the results of different types of
detrending procedures demonstrated that the double detrending method provided the
best fit for the B. frenchii increment width data.
The detrended time series for the increment widths for each otolith of B. frenchii
was used to calculate a mean index chronology (MIC) for otoliths from individuals from
both the south and lower west coast populations. These MICs were then employed to
determine whether the trends in the otolith increment widths were related to the
environmental variables given earlier. Note that year in this study corresponds to the
July (mid-winter) of one year to June (early-winter) of the following year so that each
year encompasses the main growth period, i.e. spring to autumn. As each year
encompasses the last two months of winter of one year and the first month of winter of
the next year, it is inappropriate to test for differences between winters.
The pattern of change in MICs with age for B. frenchii on the lower west and
south coasts were similar, demonstrating that although actual temperatures varied
appreciably between the two coasts, the pattern of growth was influenced by the similar
overall trends.
On the basis of Pearsonā€™s correlation coefficient, the MIC for B. frenchii on the
south coast was positively correlated with mean annual SST (r = 0.32, P = 0.02) and, to
a slightly greater extent, for this species on the lower west coast (r = 0.37, P = 0.007).
4
The MIC for B. frenchii on the south coast was not related, however, to SST when
considered in the context of seasons or months. In contrast, the MIC for B. frenchii on
the lower west coast was related individually to spring, summer and autumn (r = 0.39, P
= 0.004) and thus to the times of year when temperatures were greatest and most growth
would be expected to occur.
The relationships between MICs and temperature were next explored using
generalized additive mixed modeling (GAMM), which permitted the age of each fish at
each increment and a random individual effect to be incorporated. The GAMM results
demonstrated that the MICs for B. frenchii were positively correlated with mean annual
SST for individuals on both the lower west and south coasts, with the relationship being
stronger for the former coast. These results parallel those of the Pearsonā€™s correlation
coefficients.
Pearsonā€™s correlation coefficients demonstrated that the MIC for B. frenchii was
not correlated with mean annual FSL for the population on the lower west coast and the
same was true for the MIC on both a seasonal and monthly basis. On the basis of
GAMM, the MIC for B. frenchii on the lower west coast was strongly (P = 0.008)
related to sea level height, contrasting the result of the Pearsonā€™s correlation coefficient.
The results, derived from Pearsonā€™s correlation coefficient, and by the use of
GAMM validate the hypothesis that otolith growth in B. frenchii are correlated with
water temperature and, to a slightly greater extent, on the lower west than south coasts.
They also demonstrate a relationship between otolith growth and sea level on the lower
west coast.
5
Acknowledgements
I would like to thank my primary supervisors Peter Coulson and Ian Potter for their
guidance and patience.
I would also like to thank my "unofficial" supervisors Adrian Hordyk and Norm Hall
for their endless technical expertise and time.
Thank you to my parents Greg and Deb for their unwavering support and belief in my
abilities.
Special thanks to Justine Arnold and Nicholas Zebegew for the emotional support and
last minute advice.
Without all of you, I would not be where I am today.
6
Table of Contents
Declaration........................................................................................................................1
Abstract .............................................................................................................................2
Acknowledgements...........................................................................................................5
Table of Contents..............................................................................................................6
List of Figures ...................................................................................................................8
List of Tables & Equations .............................................................................................10
Introduction.....................................................................................................................12
Objectives........................................................................................................19
Materials & Methods.......................................................................................................21
Sample collection............................................................................................21
Otolith preparation, imaging and increment measurement.............................22
Cross dating, detrending and otolith chronology construction .......................25
Analysis...........................................................................................................27
Data Exploration .........................................................................................29
Environmental Data ........................................................................................29
Results.............................................................................................................................31
Biological data ................................................................................................31
Cross dating and synchrony among individuals within ..................................31
Detrending.......................................................................................................35
Synchrony between regions ............................................................................38
Correlations with environmental variables .....................................................38
Sea Surface Temperature ............................................................................38
Fremantle sea level......................................................................................42
7
Discussion .......................................................................................................................45
Selection of otoliths and sample sizes.............................................................45
Cross dating and synchrony among individuals .............................................46
Detrending.......................................................................................................48
Synchrony between regions ............................................................................49
Financial year vs. calendar year......................................................................50
Relationships between MIC and environmental variables..............................52
References.......................................................................................................................56
Appendix.........................................................................................................................66
List of figures ..................................................................................................66
Data Exploration .............................................................................................67
Outliers........................................................................................................67
Homogeneity...............................................................................................69
Normality ....................................................................................................70
Independence...............................................................................................71
Motivation for Modeling.............................................................................73
Power Analysis................................................................................................73
8
List of Figures
Figure 1. Mean annual values for a) sea surface temperatures (Ā°C) at Rottnest Island on
the lower west coast and at Esperance on the south coast of Western Australia and b)
sea level (cm) at Fremantle. .................................................................................. 18
Figure 2. Map of south-western Australia showing the location of where samples of
Bodianus frenchii were collected from waters near Rottnest, on the lower west coast,
and Esperance, on the south coast. Red dots denote the approximate location for which
sea surface temperature data was obtained. .......................................................... 22
Figure 3. a) The sectioned otolith of a 51 year old Bodianus frenchii and a b) higher
magnification image of part of the dorsal side of the otolith were increment
measurements were undertaken. In a) white dots indicate the first and every tenth
opaque zone and vertical black lines are indicative of the axis followed when measuring
increments. In b) horizontal white lines show the position of the outer edge of each
opaque zone........................................................................................................... 24
Figure 4. A visual representation of the successive increment widths of the 29 Bodianus
frenchii individuals from the south coast used to assist in visually identifying
synchronous patterns in those widths. Numbers on the left and right y-axis are
identifying codes for individual fish. .................................................................... 33
Figure 5. A visual representation of the successive increment widths of the 24 Bodianus
frenchii individuals from the lower west coast used to assist in visually identifying
synchronous patterns in those widths. Numbers on the left and right y-axis are
identifying codes for individual fish. .................................................................... 33
Figure 6. A visual representation of the statistical cross-dating of the standardized raw
increment time series for the 29 individual Bodianus frenchii from the south coast. Blue
correlates well (p-values less or equal to the user-set critical value) while potential
dating problems are indicated by the red segments (p-values greater than the user set
9
critical value). Green lines show segments that do not completely overlap the time
period and thus have no correlations calculated (Bunn 2010). ............................. 34
Figure 7. A visual representation of the statistical cross-dating of the standardized raw
increment time series for the 24 individual Bodianus frenchii from the lower west coast.
Blue correlates well (p-values less or equal to the user-set critical value) while potential
dating problems are indicated by the red segments (p-values greater than the user set
critical value). Green lines show segments that do not completely overlap the time
period and thus have no correlations calculated (Bunn 2010). ............................. 34
Figure 8. Comparisons of the a, b) raw chronologies, c, d) negative exponential and e, f)
cubic smoothing spline g, h) double detrending methods for individuals of Bodianus
frenchii (grey lines) from the south (blue lines) and lower west (red lines) coasts of
Western Australia.................................................................................................. 37
Figure 9. a) A comparison of the mean index chronologies (MIC) for Bodianus frenchii
from the south (blue line) and lower west coasts (red line) and a comparison of the MIC
for B. frenchii from the b) south and c) lower west coasts and mean annual sea surface
temperature (black line) in those regions and comparison of the MIC for Bodianus
frenchii from the lower west and d) mean annual Fremantle sea level (black line) for the
years between 1954 and 2005. .............................................................................. 39
Figure 10. Three dimensional visualizations of the generalised additive mixed models
displaying the relationship between a) the otolith growth of Bodianus frenchii on the
south coast and SST and otolith growth of B. frenchii on the lower west coast and b)
SST and c) FSL. x, environmental variable; y, mean increment chronology; z, smoother
of age..................................................................................................................... 44
10
List of Tables & Equations
Table 1. Examples of the different detrending methods employed in otolith based
increment width chronology studies on fish species with varying times series length and
sample size. ........................................................................................................... 15
Table 2. Length and age range, method of capture and sex of the individuals of
Bodianus frenchii from the south and lower west coasts of Western Australia used in the
this study. .............................................................................................................. 32
Table 3. Time series range, mean increment size, interseries correlation, mean
sensitivity and standard deviation in the individual time series for Bodianus frenchii
chronologies for the south and lower west coasts and both coasts combined. n = sample
size......................................................................................................................... 32
Table 4. The expressed population signal (EPS) and š‘Ÿ values for the raw chronologies,
and the increment widths for Bodianus frenchii on the south and lower west coasts of
Western Australia after the negative exponential, spline and double detrending methods
were applied. ......................................................................................................... 36
Table 5. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses) for the
relationships between the MICs of Bodianus frenchii from the south and lower west
coasts and mean annual (financial year) and mean seasonal sea surface temperatures
(Ā°C) in waters off those coasts and mean annual and mean seasonal Fremantle sea level
(cm). ...................................................................................................................... 40
Table 6. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses) for the
relationships between the MICs of Bodianus frenchii from the south and lower west
coasts and mean monthly sea surface temperatures (Ā°C) in those waters, and between
the MIC of Bodianus frenchii from the lower west coast and mean monthly Fremantle
sea level (cm). * denotes significant correlations at 0.05, ** denotes significant
correlation at 0.01.................................................................................................. 40
11
Table 7. Results of generalised additive mixed models of the effect of increasing sea
surface temperature on the growth of otoliths of Bodianus frenchii on the south and
lower west coasts................................................................................................... 42
Table 8. Results of generalised additive mixed models of the effect of increasing
Fremantle sea level (a proxy for Leeuwin Current strength) on the growth of otoliths of
Bodianus frenchii on the lower west coast............................................................ 43
Equation 1. Generalised additive mixed model formula for B. frenchii biochronologies
............................................................................................................................... 28
12
Introduction
The otoliths, ear bones of fish, consist of calcium carbonate and an organic matrix,
which are used for sensory purposes such as balance and hearing (Campana & Neilson
1985, Popper et al. 2005). Of the three pairs of otoliths, including the lapilli and
asterisci, the largest, the sagittae, have been routinely used by researchers to age
individual fish (Campana 2001). The density at which calcium carbonate material is
deposited, which varies during the year (Pannella 1971), results in the production of
alternating opaque (slow winter growth, densely deposited material) and translucent
(faster summer growth, sparsely deposited material) growth zones in the otolith. One of
the unique properties of otoliths is the consistency of which calcium carbonate and an
organic matrix is deposited (Campana & Thorrold 2001). As this material is deposited
regularly, even during times of stress or starvation and after somatic growth has largely
ceased, the growth zones in otoliths are very valuable for determining the age of fish
(Morrongiello et al. 2012).
Otoliths contain more information than simply the age of a fish. Early
researchers showed that the distances between successive daily growth bands in otoliths
could vary by manipulating the environmental conditions under which fish were held in
aquaria (Pannella 1971, Campana & Neilson 1985), and that historical annual time
series of growth could be obtained by measuring the widths of annuli in otoliths, as a
proxy for annual fish growth (Boehlert 1985, Pereira et al. 1995, Lehodey &
Grandperrin 1996). More recently, trends in the widths of successive growth increments
have been used to investigate how inter-annual fluctuations in environmental variables,
such as water temperature and the strength of major ocean current systems, influence
the growth of fish species (Black et al. 2011a, Black et al. 2013a, Rountrey et al. 2014,
Morrongiello & Thresher 2015). It is hypothesized that variations in increment widths
13
are produced by a number of intrinsic and extrinsic factors (Miller et al. 2010,
Morrongiello et al. 2012).
Intrinsic factors include genetic predispositions (Morrissey 2011), biological or
physiological factors, while extrinsic factors include inter- and intra-species competition
and environmental influences, such as large scale oceanographic systems such as the El
Nino Southern Oscillation (Lehodey & Grandperrin 1996). The aim of previous
sclerochronological (sclero = hard part, chronos = time series) work employing otoliths
has been to detect relationships between the trends displayed by successive increment
widths in otoliths with extrinsic factors such as air temperature (Black et al. 2013b), sea
surface temperature (Coulson et al. 2014), sea bottom temperature (Black 2009), wind
direction (Black et al. 2011a), oceanic upwelling (Black et al. 2011b), the El Nino
Southern Oscillation (Meekan et al. 1999) and hydrological regimes of freshwater
environments (Morrongiello et al. 2011, Black et al. 2013b). The relationships
established between growth and environmental variables can then be used to make
ecological inferences about populations, and population interactions and to forecast
ecological reactions to those environmental drivers (e.g. Rountrey et al. 2014).
The methods for constructing biochronologies from increment width
measurements obtained from otoliths in many recent studies (i.e. Black et al. 2005,
2008a; Matta et al. 2010 Gillanders et al. 2012) have been heavily borrowed from those
used to construct chronologies from annual banding in the trunks of trees (Douglass
1920, 1941), which have been developed over the past 95 years. Special emphasis has
been placed, by dendrochronology (dendro = tree, chronos = time series), on the
importance of the process of cross dating, i.e. ensuring that each increment is assigned
correctly to the year of formation (Fritts 1976, Cook & Kairiukstis 1990, Maxwell et al.
2011). Cross dating, which is traditionally carried out by visually inspecting the
14
individual increment series and matching conspicuously wide and or narrow increments,
is possible on these otoliths for those species that display such trends (e.g. Matta et al.
2010; Coulson et al 2014; Tao et al. 2015). However, when increments are only 40-Āµm
wide, visually detecting patterns of wide and narrow increments is very difficult
(Rountrey et al. 2014, Nguyen et al. 2015). In these cases, statistical cross dating, a
process which mimics visual cross dating using correlation analyses, has become
increasingly important.
In order to tease out any climate signals in otolith increment time series,
detrending, i.e. the process of standardizing growth ring widths in chronologies to a
mean of one, is used to remove age-related growth declines while preserving as much
environmentally-induced variability as possible to better illustrate climate-driven
anomalies in ring widths (Cook 1985). The most common method of detrending in
otolith-based biochronology studies is a modified exponential or curvilinear method
(Table 1). Other detrending methods used include smoothing splines, double detrending
and regional curve standardization (Table 1). Historically, curvilinear standardization
has been used in dendrochronological studies carried out in the north-west of America,
where the trees exhibit exponential growth (Cook 1985). Curvilinear standardization is
inappropriate in those cases where the trend in the increment time series is not perfectly
exponential or linear (Cook 1985). The smoothing spline represents a highly flexible
technique, which represents a major improvement in standardizing ring-width series
compared to linear or curvilinear fits (Cook & Peters 1981). Double detrending is the
combination of negative exponential and smoothing spline detrending techniques,
where the data are detrended with a spline, then a negative exponential function, hence
it is detrended twice (Rountrey et al. 2014). Regional curve standardization is a
detrending technique whereby a single increment is standardized by taking the average
15
increment width of a calendar year over the average increment width for an age class
(Briffa et al. 1996, Melvin & Briffa 2008) and preserves low frequency variance but
removes the overall trends in long time series (Melvin & Briffa 2008). This present
study determined the best method for detrending chronologies for the Bodianus frenchii.
hom
Table 1. Examples of the different detrending methods employed in otolith based
increment width chronology studies on fish species with varying times series length and
sample size.
Detrending method Species Time series
length
Sample
size
Curvilinear
standardization
red snapper Lutjanus campechanus1
gray snapper Lutjanus griseus 1
lake trout Salvelinus namaycush2
freshwater drum Aplodinotus grunniens3
28
31
21
22
30
24
17
1351
Smoothing splines splitnose rockļ¬sh Sebastes diploproa4
yelloweye rockļ¬sh Sebastes ruberrimus5
49
50
50
66
Double detrending western blue groper Achoerodus gouldii6
51 56
Regional curve
standardization
rock flathead Platycephalus laevigatus7
longhead flathead Leviprora inops 7
yellowfin sole Limanda aspera8
14
12
21
96
120
17
Black et al. 4
2005, 5
2008a,1
2011a, 2
2013b, 8
2013a, 3
Davis-Foust, 2012; 6
Rountrey et al. 7
2014; Coulson et
al. 2014
The Foxfish Bodianus frenchii is a medium-sized, shallow-water, temperate reef
fish of considerable longevity, attaining a maximum age of 78 years (Cossington et al.
2010), making it the longest-lived species within the speciose Labridae family. A
comprehensive study of the biology of B. frenchii collected over reefs on the lower west
and south coasts of Western Australia demonstrated that this protogynous
hermaphroditic labrid attains a greater length and body mass at age throughout life in
waters on the cooler south coast than on the warmer lower west coast and that females
mature at an earlier age on the former coast and that those females exhibit substantially
greater fecundity (Cossington et al. 2010). Bodianus frenchii, endemic to temperate
Australian waters, is most common off the lower west and south coasts of Western
16
Australia, where it co-occurs with another long-lived, but much larger labrid, the
Western Blue Groper Achoerodus gouldii (Cossington et al. 2010).
The otolith growth of A. gouldii, a species which attains a maximum age of 70
years, was shown to be positively and significantly correlated with sea surface
temperatures (SST) off the south coast of Western Australia in the months from
November to May i.e. the time of year of increasing and elevated water temperatures
and thus the main growing period (Rountrey et al. 2014). This is similar to the period
for which there is a positive relationship between SST and otolith growth for two
species of flathead (Platycephalidae) in the same region (Coulson et al. 2014, Rountrey
et al. 2014). In deep water off south-western Australia, the growth of Hapuku Polyprion
oxygeneios was found to also have a significant, positive relationship with SST in
winter and spring months of the previous year (Nguyen et al. 2015). In neither the study
involving A. gouldii nor the two platycepahlids was there any correlation between
otolith growth and the strength of the Leeuwin Current, a dominant oceanographic
feature along the west Australian coast. However, Nguyen et al. (2015) found a lagged
positive relationship between the growth of Hapuku Polyprion oxygeneios and Leeuwin
Current strength in the previous calendar year. The lag in the effect of the Leeuwin
Current on the growth of P. oxygeneios was thought to be a result of the time required
for their main (> 50%) prey, squid, most likely arrow squid Nototodarus gouldi, to
attain a size where they form a substantial part of the hapuku diet (Nguyen et al. 2015).
The near shore waters of the lower west and south coasts of Western Australia
differ markedly in their habitat structures, reef types and water depths. It is generally
accepted that the habitat structure within a region affects the biological community
structure and that habitat structure is driven by physical oceanography within a region
(Carter & Woodroffe 1994, List & Terwindt 1995). The near-shore waters of the west
17
coast are protected by a semi-continuous, limestone-based reef and barrier island
system, which moderates wave energy (Howard 1989, Wells et al. 1993, Sanderson et
al. 2000). In contrast, granite boulder reefs and headlands dominate near-shore waters of
the south coast which is fully exposed to the southern ocean and is a high wave-energy
coast (Wells & Keesing 1990, Wells et al. 1993, Kendrick 1999, Sanderson et al. 2000).
In addition, the waters of the south coast increase in depth far more rapidly with
distance from shore than is the case with waters off the lower west coast (Kendrick
1999, Sanderson et al. 2000).
Another important difference between the two coasts is their temperature
regimes, with average temperature, and the range in average temperature, on the lower
west coast being greater than on the south coast (Kendrick 1999, Sanderson et al. 2000)
(Figure 1). An important contributor to this difference is not only lower latitude, but
also the influence of the pole-ward flowing, warm waters of the Leeuwin Current on
this coast. The Leeuwin Current has a strong influence on the distribution and
abundance of marine flora and fauna in south western Australia (Pearce & Phillips
1988, Caputi et al. 1996), and in recent years has led to marine heat wave-like
conditions in waters off the lower west coast (Pearce & Feng 2011, Rose et al. 2012,
Feng et al. 2013, Pearce & Feng 2013). Fremantle sea level (FSL) has been used as a
proxy for the strength of the Leeuwin Current, i.e. strong current strength results in
higher FSL while weak current strength results in lower FSL, which, until recently, has
been closely correlated with the recruitment of rock lobster to reefs along the lower west
coast (Pearce & Phillips 1988, Caputi et al. 2001).
18
Figure 1. Mean annual values for a) sea surface temperatures (Ā°C) at Rottnest Island on
the lower west coast and at Esperance on the south coast of Western Australia and b)
sea level (cm) at Fremantle.
Objectives
The species that is the focus of this thesis is a long-lived, temperate fish species that
lives in a region experiencing increases in temperature (Pearce & Feng 2007, Pearce &
Feng 2011). Previous otolith-based biochronology studies have demonstrated that water
temperature during the summer months, and thus those when the majority of growth is
occurring, has an important influence on the growth of fish in Western Australia
19
(Coulson et al. 2014; Rountrey et al. 2014). Those studies were restricted, however, to
fish populations along the south coast of Western Australia. The current study thus
investigates otolith growth of B. frenchii from the lower west as well as south coast
waters, it also employs the otoliths of this species caught in waters off the lower west
coast, where temperatures are greater and have increased since 1970. Furthermore, the
influence of the Leeuwin Current is greater on the lower west coast and occurs more
inshore than on the south coast.
The first aim of this study was to develop a biochronology, using growth
increment widths in otoliths, for the long-lived B. frenchii on both the south and lower
west coasts in south-western Australia, which could be used for the following purposes.
1) Determine whether the patterns of otolith growth of the two populations are
synchronized, thereby implying that the growth of individuals of this species in these
two distantly-located populations respond similarly to environmental influences.
2) Test the hypotheses that, the waters off the lower west coast of Australia exhibit
greater inter-annual fluctuations in temperature than off the south coast of Western
Australia, the relationship between the otolith increment width chronology for B.
frenchii will be stronger in the waters of the former coast.
3) Test the hypotheses that, as the Leeuwin Current exhibits a substantial influence on
waters along the west coast of Australia as far south as the lower west coast, it will have
an effect on the pattern of otolith growth of B. frenchii in these waters.
20
Materials & Methods
Sample collection
Bodianus frenchii, whose otoliths were used in this current study, were collected
between April 2004 and 2006 as part of an earlier study on the biological characteristics
of this species in south-western Australia (Cossington et al. 2010). In that study,
individuals were collected from numerous locations in marine waters between Jurien
Bay at ~ 30Ā°18ā€™S, 115Ā°02ā€™E on the lower west coast and from Esperance at ~ 33Ā°51ā€™S,
121Ā°53ā€™E on the south coast (Figure 2). To enable the current study to investigate the
influence of environmental variables on the growth of B. frenchii at the northern and
southern extent of this species distribution for which sufficient samples were available,
B. frenchii sampled from Rottnest Island at ~ 32Ā°00ā€™S, 115Ā°52ā€™E (hereafter referred to
ā€˜lower west coastā€™) and Esperance (hereafter referred to ā€˜south coastā€™) were selected for
otolith increment analysis. Bodianus frenchii from Rottnest Island were sampled by rod
and line angling and spear fishing while SCUBA diving, whereas those from Esperance
where sampled by spear fishing while snorkeling and collected from recreational line
and commercial gillnet fishers (Table 2).
One of the aims of this study is to investigate the influence of fluctuating
environmental conditions on the growth of B. frenchii (see earlier). In order to
maximize temporal resolution of the data and the greater certainty of resultant
correlation tests that is afforded by using long time series (Wigley et al. 1984), i.e. a
series of successive increment measurements from an individual, the sectioned otoliths
from the oldest individuals from each of the two coasts were preferentially selected. In
addition, only those otoliths from the oldest fish, whose increment boundaries could be
clearly defined, allowing for the most precise increment measurements, were employed
for otolith chronology construction.
21
Figure 2. Map of south-western Australia showing the location of where samples of
Bodianus frenchii were collected from waters near Rottnest, on the lower west coast,
and Esperance, on the south coast. Red dots denote the approximate location for which
sea surface temperature data was obtained.
Otolith preparation, imaging and increment measurement
The otoliths employed during the current study were previously sectioned for
age determination (Cossington et al. 2010). This involved embedding one otolith from
each individual in clear epoxy resin before cutting a thin, ~ 300 Ī¼m thick transverse
section, using a low-speed diamond saw (Buehler), through the otolith primordia. The
surface of the sections were then lightly polished using fine wet and dry carborundum
paper (grade 1200) and mounted on slides using DePX mounting medium with a cover
slip. Multiple images of the dorsal side of the sectioned otoliths were taken at a
22
magnification of 10X using an Olympus DP70 12.0 megapixel digital camera mounted
on an Olympus BX51 stereo microscope. These images were then stitched together
using the stitch function in Leica Application Suite V. 4.3 (Leica Microsystems).
The measurement of the widths of successive growth increments on the digital
images of sectioned otoliths was carried out using ImageJ V. 1.47 (AbrĆ moff et al.
2004) and employing the plugin ā€˜IncMeasā€™ (Rountrey 2009). As the width of an
increment at one particular location on an otolith may vary from the width of that same
increment measured elsewhere, increments were measured along three to five transects
in order to obtain a more precise, otolith-wide measurement of that increment. The
average of these transects was used as the time series for an individual. Increments were
measured to the nearest 1 Āµm along transects that were drawn perpendicularly to the
otolith increments on the digital images of the sectioned otoliths, i.e. parallel to growth
direction (Figure 3a). Starting at the outermost opaque zone, the outer edge of each
individual opaque zone was marked until the edges of those opaque zones where no
longer clearly defined, after which increment measurement ceased (Figure 3b). This
often meant that the increments representing the growth undertaken in the first 5 years
of life were not measured. Partial increments on the peripheral side of the outermost
opaque zone were not measured.
23
Figure 3. a) The sectioned otolith of a 51 year old Bodianus frenchii and a b) higher
magnification image of part of the dorsal side of the otolith were increment
measurements were undertaken. In a) white dots indicate the first and every tenth
opaque zone and vertical black lines are indicative of the axis followed when measuring
increments. In b) horizontal white lines show the position of the outer edge of each
opaque zone.
24
Cross dating, detrending and otolith chronology construction
Cross dating is an important process that relies on the assumption that there are
synchronous growth patterns in multiple individuals due to the influence of common
environmental drivers of growth (Black et al. 2005). By matching growth patterns
among individuals, missing or false increments can be identified and taken into account
when assigning calendar years, ensuring that each increment is assigned to the correct
year of formation (Kastelle et al. 2011). While visual cross dating has been possible for
some fish species (Black et al. 2005, Gillanders et al. 2012, Coulson et al. 2014), it was
not possible for B. frenchii, as was the case for Achoerodus gouldii (Rountrey et al.
2014), due to the very small increment widths, ~ 10-20 Āµm, thus making it difficult to
visually detect any changes in those widths. As the date of capture of the individuals
used in this study was known, this served as an anchor for each increment measurement
series and accordingly back-dated. Statistical cross-dating was performed using the
dendrochronology program library dplR in R (Bunn 2010) following the methods
outlined in Bunn et al. (2015). This cross dating program was used to fit each series of
increment measurements with a highly flexible cubic smoothing spline using a 15-year
moving window. Each time series was then divided by the values predicted by the
spline function, removing low-frequency variability and standardizing all measurement
time series to a mean of one. The detrended time series for each individual was then
correlated, at segment lengths of 20 years, with a lag of 10 years, with the average for
all other standardised time series of measurements for all individuals. The mean,
calculated as a Tukey's biweight robust mean, was reported as the inter-series
correlation. This approach of isolating only the high frequency, serially-independent
growth pattern mathematically mimicked the process of visual cross-dating (Holmes
1983, Grissino-Mayer 2001). dplR also calculated the mean sensitivity, a measure of
25
the relative change in increment width between successive years that ranges from a
minimum of zero, i.e. two increments with the same width, to a theoretical maximum of
two, i.e. a pair of increments in which one value is zero (Fritts 1976). The
corresponding otoliths whose time series were highlighted in the output by the dplR
program to have potential errors, were visually expected, and if no error was found the
increment measurement was not adjusted.
To develop the mean index chronology (MIC) for climate analysis, each set of
the original growth-increment measurement time series was detrended employing
double detrending method outlined in Rountrey et al. (2014). The detrending effectively
removed the rapid ontogenetic decline in growth rate that occurs in early life, and
removed other low frequency variation including departures from the negative
exponential fit. Additional detrending techniques were also explored such as negative
exponential detrending (Black 2009, Black et al. 2010, Matta et al. 2010, Black et al.
2011a, Gillanders et al. 2012) and cubic smoothing splines (Black et al. 2005). All
detrending analyses were carried out employing the R package dplR (Bunn 2008) and
only those years with a sample depth of > five individuals were retained to ensure a true
mean is calculated (Matta et al. 2010, Black et al. 2013a).
26
Analysis
For the analysis the chronologies of B. frenchii a hybridization of sclerochronology
techniques, Pearson's correlation analysis and additive mixed-effects modeling was
used, similar to Rountrey et al (2014), who investigated the relationship between otolith
growth of the co-occurring A. gouldii and environmental variables on the south coast.
The analysis can be broken down into three parts: 1) data exploration, 2) Pearson's
correlation analyses, and 3) modeling otolith growth, via mean index chronology
(MIC), using a generalized additive mixed model, with age as a smooth function and an
individual random effect. Reasoning for using the GAMM approach is outlined in the
Appendix.
To determine a parsimonious predictive relationship between the MICs for B.
frenchii on the south and lower west coasts and the mean annual sea surface temperature
in waters of those regions and Fremantle sea level, generalised additive mixed models
(GAMMs) were used. This resulted in three models, each model was created, using one
predictor, either mean annual sea surface temperature (for both the south coast and
lower-west coast) or mean annual Fremantle sea level (lower-west coast). Models were
created using gamm4 in R (Wood et al. 2014).
27
To investigate the effects of variations in sea surface temperature on B.
frenchii growth, we created a generalised additive mixed model in the form of:
Equation 1. Generalised additive mixed model formula for B. frenchii
biochronologies
š‘”(š‘€š¼š¶š‘–,š‘—
š‘
) = š›½0 + š‘“1(š“š‘”š‘’š‘–,š‘—) + š‘‹š‘— + š‘§š‘–
š‘‡
š‘
where:
š‘”(š‘€š¼š¶š‘–,š‘—
š‘
) is a monotonic differential link function of fish otolith
width, where MIC(i, j) where i is the MIC increment and j
is year relating to increment;
š›½0 is the intercept parameter;
š‘“1(š“š‘”š‘’š‘–š‘—) is a centered twice-differentiated smooth function for fish
age;
š‘‹š‘— is the average annual (financial calendar) environmental
variable for year j;
š‘§š‘–
š‘‡
š‘ is an error term; where š‘ are random effects, assumed to be
distributed
~š‘{0, š·( šœƒ)} where šœƒ is a š‘ Ɨ 1 vector of variance components
28
Data Exploration
Data exploration is the first step in data analysis (Tukey 1977, Zuur et al. 2010) with the
objective of summarizing the dataset's main characteristics, often with visual methods.
Data exploration is described as "graphical detective work" (Tukey 1972) of factors,
such as outliers, unequal variance, or dependence structures, which affect the type of
analysis which can be performed. The results of the data exploration can be found in the
Appendix.
Environmental Data
A range of environmental variables and climate indices (i.e. bottom temperature, wind
direction, oceanic upwelling, El Nino Southern Oscillation Index or Multivariate ENSO
Index, and Dipole Mode Index) were considered in order to determine their influence on
the growth of B. frenchii in south-western Australia. However, many of these variables
were not available throughout much of the early years of the MIC on both the west and
south coasts. The variables which were selected, were also the most likely to influence
the growth of fish in south-western Australia, were sea surface temperature (SST) and
Fremantle sea level (FSL), a measure of the strength of the Leeuwin Current. No
attempt was made to relate otolith growth with sea level height on the south coast as the
Leeuwin current flows well offshore on that coast.
Mean monthly SST data was collected from Hadley Satellites and obtained from
the Met Office Marine Data Bank (Rayner 2003). Mean monthly SST data was obtained
for every month between January 1938 and December 2005 from the grid cell off the
lower-west coast (-32Ā°, 116ā€™E) and south coast (-35Ā°, 119ā€™E). Fremantle sea level was
obtained from the Bureau of Meteorology. Pearsonā€™s correlation coefficients (r) were
tested to determine whether there was a significant relationship between the B. frenchii
29
MIC and mean annual, seasonal and monthly SST data for each coast and mean annual,
seasonal and monthly FSL.
In the present study, correlation tests between otolith increment widths and
environmental variables were conducted with those variables organized on a financial
year scale, rather than calendar year. The financial year is preferred to the calendar year,
because it follows more closely the biological growth year of teleosts in the southern
hemisphere (Coulson et al. 2014), i.e. the peak growing period is summer (December -
February) which would be divided if the traditional calendar year scale was employed.
30
Results
Biological data
The ages of individuals used in this chronology study ranged from 39 - 78 years for the
south coast and 41 - 62 years for the west coast, which corresponded to lengths ranging
from 372 - 435 and 315 - 446 mm, respectively (Table 2). As B. frenchii is a
protogynous hermaphrodite that changes sex at ~ 29 years (Cossington et al. 2010), all
but one of the individuals from each coast that were used in this chronology study were
male (Table 2).
Cross dating and synchrony among individuals within
The increments measured encompassed years between 1938 and 2005 for the south
coast and 1947 and 2005 for the west coast. A sample depth of at least six occurred in
those years between 1954 and 2005 and 1954 and 2004 on south and lower west coasts,
respectively. Visually, there is no pattern of variation in the widths of the increments in
the otoliths of B. frenchii within and between the otoliths of the individuals on both
coasts (Figure 4, Figure 5). As visual cross dating was not possible due to the very small
increment widths (Table 3) increasing the difficultly in identifying synchrony in
conspicuously wide and/or narrow increments, statistical cross dating was employed,
with the capture year used as the anchor for this process. Although there was some
synchrony among a few individuals on each coast (indicated by the blue segments in
Figure 6 and Figure 7) shown by the statistical cross dating, on the whole, this process
also demonstrated a lack of synchrony among those same individuals (indicated by the
red segments in Figure 6 and Figure 7).
Despite this apparent lack of synchrony identified by statistical cross dating, the
interseries correlation (ISC), a measure of how much correlation exists amongst the
31
increment widths (Grissino-Mayer 2001), demonstrates that there is synchrony among
the standardised increment time series for those individuals of B. frenchii on the south
coast, and to a lesser extent, the lower west coast (Table 3). The very low mean
sensitivity values, a measure of the change in increment width between successive
years, of 0.14 for the south coast and 0.01 for the lower west coast indicate that there is
very little variation between successive increment widths, which is probably why visual
and statistical cross dating failed to find any synchrony (Table 3).
Table 2. Length and age range, method of capture and sex of the individuals of
Bodianus frenchii from the south and lower west coasts of Western Australia used in the
this study.
Length range
(mm)
Age range
(years)
Method of capture Sex
Gillnet Line Spear F M
South 372-435 39-78 8 10 11 1 28
Lower west 315-446 41-62 20 4 1 23
Table 3. Time series range, mean increment size, interseries correlation, mean
sensitivity and standard deviation in the individual time series for Bodianus frenchii
chronologies for the south and lower west coasts and both coasts combined. n = sample
size.
Time series
(years)
Mean
increment
size (Āµm)
Interseries
correlation
Mean
sensitivity
Standard
deviation
n
South 33-68 46 0.32 0.14 4.87 29
Lower west 33-58 42 0.17 0.11 3.02 24
Combined coasts 31-59 44 0.19 0.12 4.15 53
32
Figure 4. A visual representation of the successive increment widths of the 29 Bodianus
frenchii individuals from the south coast used to assist in visually identifying
synchronous patterns in those widths. Numbers on the left and right y-axis are
identifying codes for individual fish.
Figure 5. A visual representation of the successive increment widths of the 24 Bodianus
frenchii individuals from the lower west coast used to assist in visually identifying
synchronous patterns in those widths. Numbers on the left and right y-axis are
identifying codes for individual fish.
33
Figure 6. A visual representation of the statistical cross-dating of the standardized raw
increment time series for the 29 individual Bodianus frenchii from the south coast. Blue
correlates well (P less or equal to the user-set critical value) while potential dating
problems are indicated by the red segments (P greater than the user set critical value).
Green lines show segments that do not completely overlap the time period and thus
have no correlations calculated (Bunn 2010).
Figure 7. A visual representation of the statistical cross-dating of the standardized raw
increment time series for the 24 individual Bodianus frenchii from the lower west coast.
Blue correlates well (P less or equal to the user-set critical value) while potential dating
problems are indicated by the red segments (P greater than the user set critical value).
Green lines show segments that do not completely overlap the time period and thus
have no correlations calculated (Bunn 2010).
34
Detrending
A range of detrending methods that are most commonly used in otolith increment
width-based chronology studies were employed in order to determine which of these
was most suitable to apply to the B. frenchii increment width data.
The negative exponential detrending method is a rigid function that removes age
related growth declines in the increment series that are more prevalent in tree ring data,
for which it was developed. The negative exponential detrending method does not
appear to have has not completely removed the age-related decline in the otolith
increment-width data for B. frenchii, as there is still a declining trend in the time series
(Figure 8c, d). The cubic smoothing spline method (often termed 'spline'), which is by
far the most common in dendrochronology, but not so in otolith sclerochronology,
works by taking a mean of the surrounding four data points, in front, behind, above and
below. Double detrending also employs a cubic-smoothing spline after the increment
width data is initially detrended by applying a negative exponential function (Cook &
Kairiukstis 1990). When applied to the increment width data for B. frenchii from each
coast, both the spline and double detrending methods have largely removed any age
related decline in the increment width data (Figure 8e, f, g, h). In addition, the detrended
time series also display a homogeneous variance around a mean of 1. Furthermore,
these detrending methods, while eliminating the age related decline in otolith growth,
have preserved the high frequency climate effects, which is particularly evident in the
south coast B. frenchii data where there is conspicuous decrease in increment width in
the early 1990s (Figure 8e, g).
The higher expressed population signal (EPS), a measure of how well the
sample means represent the mean of the theoretical population (Wigley et al. 1984),
provided by negative exponential detrending suggests that this best detrending option
for the B. frenchii increment width data (Table 4). Chronologies with EPS values above
35
the theoretical threshold of 0.85 or high š‘ŸĢ… values are generally considered to be
acceptable; however, detrending with the aim of maximizing these indicator parameters
may result in the retention of aspects of an ontogenetic trend (Nguyen et al. 2015).
Thus, despite the negative exponential detrending option providing the highest EPS and
š‘ŸĢ… values, visually it was still shown to retain an ontogentic decline and was therefore
discarded. The EPS and š‘ŸĢ… values for the spline and double detrending methods for the
increment data for south coast population were identical (Table 4). As the EPS and
š‘ŸĢ… values for double detrending of the increment data for lower west coast population
were marginally higher than those obtained employing the spline detrending, and to
enable comparisons with the previous otolith chronology study of the Western Blue
Groper Achoerodus gouldii in the same region (Rountrey et al. 2014), double detrending
was chosen as the detrending method for B. frenchii on the south and lower west coasts.
Table 4. The expressed population signal (EPS) and š‘ŸĢ… values for the raw chronologies,
and the increment widths for Bodianus frenchii on the south and lower west coasts of
Western Australia after the negative exponential, spline and double detrending methods
were applied.
Raw Negative
Exponential
Spline Double
Detrending
South
EPS 0.854 0.878 0.697 0.697
š‘ŸĢ… 0.221 0.259 0.100 0.100
Lower West
EPS 0.741 0.821 0.421 0.433
š‘ŸĢ… 0.131 0.195 0.037 0.039
36
Figure 8. Comparisons of the a, b) raw chronologies, c, d) negative exponential and e, f) cubic smoothing spline g, h) double detrending methods for individuals
of Bodianus frenchii (grey lines) from the south (blue lines) and lower west (red lines) coasts of Western Australia.
37
Synchrony between regions
The MICā€™s for the two B. frenchii populations were significantly correlated (r =
0.15, P= 0.005), indicating that there is a high level of synchrony in the otolith
growth of the individuals of these two populations (Figure 9a). This finding
suggests that B. frenchii on the south and lower west coasts respond in a similar
way to variations in temperature, despite the fact that these two populations are
spatial separated (~1200km) and occur on coasts that are markedly different in
terms of their reef types and the environmental conditions experienced. Even
though synchrony in the trends of increment widths is present between the two B.
frenchii populations, in order to investigate the relationships between otolith
growth of individuals in those two populations and region specific environmental
variables, the MICs have not been combined to generate a single MIC.
Correlations with environmental variables
Sea Surface Temperature
On the basis of Pearsonā€™s correlation coefficient, the MIC for B. frenchii on the
south coast was weakly and positively correlated with mean annual (financial
year) SST (r = 0.32, P = 0.02) in those years between 1954 and 2005 (Table 5).
There was a stronger positive relationship (r = 0.37, P = 0.007) between the MIC
for B. frenchii on the lower west coast mean annual (financial year) SST (Table
5). This positive relationship on an annual level for the lower west coast was
largely driven by the significant positive correlation (r = 0.39, P = 0.004) with
mean seasonal SST for summer, the time of year when temperature is highest and
when the majority of growth is expected to occur (Table 6).
38
Figure 9. a) A comparison of the mean index chronologies (MIC) for Bodianus
frenchii from the south (blue line) and lower west coasts (red line) and a
comparison of the MIC for B. frenchii from the b) south and c) lower west coasts
and mean annual sea surface temperature (black line) in those regions and
comparison of the MIC for Bodianus frenchii from the lower west and d) mean
annual Fremantle sea level (black line) for the years between 1954 and 2005.
39
Table 5. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses)
for the relationships between the MICs of Bodianus frenchii from the south and
lower west coasts and mean annual (financial year) and mean seasonal sea surface
temperatures (Ā°C) in waters off those coasts and mean annual and mean seasonal
Fremantle sea level (cm).
South West
Financial SST Annual 0.32 (0.0216*) 0.37 (0.0075**)
Spring 0.06 (0.6725) 0.30 (0.0355*)
Summer 0.20 (0.1532) 0.39 (0.0042**)
Autumn 0.22 (0.1172) 0.32 (0.0204*)
Financial FSL Annual 0.18 (0.2309)
Spring -0.01 (0.9334)
Summer 0.18 (0.2154)
Autumn 0.11 (0.4387)
Table 6. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses)
for the relationships between the MICs of Bodianus frenchii from the south and
lower west coasts and mean monthly sea surface temperatures (Ā°C) in those
waters, and between the MIC of Bodianus frenchii from the lower west coast and
mean monthly Fremantle sea level (cm). * denotes significant correlations at 0.05,
** denotes significant correlation at 0.01.
South West
SST July 0.14 (0.3136) 0.30 (0.0316*)
August 0.10 (0.4827) 0.27 (0.0575)
September 0.09 (0.5398) 0.29 (0.0419*)
October 0.11 (0.4502) 0.41 (0.0026**)
November 0.21 (0.1331) 0.29 (0.0409*)
December 0.15 (0.2769) 0.19 (0.1793)
January 0.14 (0.3360) 0.37 (0.0081**)
February 0.23 (0.1054) 0.35 (0.0110*)
March 0.25 (0.0683) 0.29 (0.0357*)
April 0.19 (0.1769) 0.29 (0.0366*)
May 0.16 (0.2487) 0.30 (0.0347*)
June 0.25 (0.0712) 0.28 (0.0450*)
FSL July -0.01 (0.9494)
August 0.05 (0.7482)
September 0.15 (0.2832)
October 0.15 (0.2934)
November 0.15 (0.2832)
December 0.05 (0.7330)
January 0.22 (0.1306)
February 0.14 (0.3399)
March 0.11 (0.4582)
April 0.16 (0.2555)
May 0.03 (0.8158)
June 0.14 (0.3224)
40
Generalised additive mixed models employing a smoother for age at which
the increment was formed and a random intercept for each individual
demonstrated that there is a significant relationship between mean annual SST and
otolith growth of B. frenchii on both the south (P = 0.003, Figure 9b) and lower
west (P = 0.0006, Figure 9c) coasts. Although, the relationship has a higher
significance on the lower-west coast, there was a higher estimated effect for the
south coast. Thus, for B. frenchii on the south coast, it is estimated that for any
given year, every 1Ā°C increase in SST the width of the increment corresponding to
that year will be 1.25Āµm (SE Ā± 0.42) wider, while the width of this increment in
the otolith of an individual off the lower west coast is estimated to be 1.18Āµm (SE
Ā± 0.35) wider (Table 7).
Both the south (-14.5 Ā± 2.9) and lower west (-14.6 Ā± 2.8) coasts had
similar estimates for the smoother of age (Table 7). This equates to the otolith
increment width decreasing at a rate of -14.5 and -14.6 Ī¼m for the south and lower
west coasts, respectively, for every increasing year of age of the fish. The GAMM
combines both the effects of the predictor and the smoother of the age to model
the growth of B.frenchii otolith increments. This can be better visualized using 3D
models (south coast: Figure 10b, lower-west coast: Figure 10c).
41
Table 7. Results of generalised additive mixed models of the effect of increasing
sea surface temperature on the growth of otoliths of Bodianus frenchii on the
south and lower west coasts.
South coast
Random effects Variance SD
FishID (Intercept) 6.483 2.546
Penalized component of age smooth 30832.695 175.592
Residual 11.912 3.451
Fixed effects Estimate SE t P
Intercept -3.7223 7.6025 -0.490 0.625
Annual (financial year) SST 1.2538 0.4214 2.975 0.003 **
Unrealized component of age smooth -14.5241 2.8546 -5.088 <2e-16 ***
Lower west coast
Random effects Variance SD
FishID (Intercept) 14.93 3.864
Penalized component of age smooth 79763.31 282.424
Residual 15.78 3.973
Fixed effects Estimate SE t P
Intercept -4.113 7.018 -0.586 0.557917
Annual (financial year) SST 1.183 0.346 3.419 0.000648***
Unrealized component of age smooth -14.623 2.786 -5.248 <2e-16 ***
Fremantle sea level
Pearsonā€™s correlation coefficients demonstrated that the MIC for the B. frenchii
population on the lower west coast was not correlated with mean annual, mean
seasonal or mean monthly FSL (Tables 5, 6). In contrast, on the basis of GAMM,
the MIC for B. frenchii on the lower west coast was positively related to mean
annual FSL (Table 8). This model also estimated that, for every given year, every
1 cm increase in mean annual Fremantle sea level will result in a 6.34Āµm increase
in the width of the increment for that year, for individuals on the lower-west coast.
42
Table 8. Results of generalised additive mixed models of the effect of increasing
Fremantle sea level (a proxy for Leeuwin Current strength) on the growth of
otoliths of Bodianus frenchii on the lower west coast.
t Coast
Random effects Variance SD
FishID (Intercept) 14.48 3.805
Penalized component of age smooth 67003.44 258.850
Residual 15.64 3.955
Fixed effects Estimate SE t P
Intercept 14.875 1.891 7.866 8.44e-15 ***
Financial Annual FSL 6.339 2.377 2.667 0.00777 **
Unrealized component of age smooth -14.596 2.783 -5.244 <2e-16 ***
Visualizing GAMMs
GAMMs combine both the effects of the predictor (i.e. the respective
environmental variables = x), the smoother of the age (z) to model the growth of
B.frenchii otolith increments (y = MIC). This can be better visualized using three
dimensional visualizations of the models (south coast SST model: Figure 10a,
lower west coast SST model: Figure 10b, and lower west coast FSL model: Figure
10c).
43
Figure 4. Three dimensional visualizations of the generalised additive mixed models displaying the relationship between a) the otolith growth of
Bodianus frenchii on the south coast and SST and otolith growth of B. frenchii on the lower west coast and b) SST and c) FSL. x, environmental
variable; y, mean increment chronology; z, smoother of age.
44
Discussion
Selection of otoliths and sample sizes
Bodianus frenchii is a protogynous labrid of exception longevity, attaining a
maximum age of 78 years (Cossington et al. 2010). For this study, in order to
maximize the chronology length, which increases the strength and validity of
potential correlations with environmental variables, the otoliths of the oldest
individuals were preferentially selected. The final number of otoliths, whose
increment widths were measured for the analyses, were selected, largely, on the
basis of the clarity of increment boundaries across a broad region of the dorsal
side of the sectioned otolith allowing for multiple transects along which
increments were measured (Table 2). Similar selection criteria have been used for
freshwater, lake trout Salvelinus namaycush and Selincuo naked carp
Gymnocypris selincuoensis (Black et al. 2013b, Tao et al. 2015) and marine
teleosts aurora rockfish Sebastes aurora, northern rock sole Lepidopsetta
polyxystra, yellowfin sole Limanda aspera and Alaska plaice Pleuronectes
quadrituberculatus (Matta et al. 2010, Thompson & Hannah 2010). Clarity of
growth increment boundaries is an important feature to be taken into account
when considering a species for otolith-based chronology studies. When
investigating relationships between otolith (fish) growth and environmental
variables, imprecise increment measurements may lead to a result of no
correlation or, worse, false correlations with those variables.
For this study, the clearest sectioned otoliths from the oldest individuals
were retained for increment measurement. The resultant sample sizes were of 29
and 24 for the south and lower west coast, respectively, were consistent with the
45
majority of other otolith increment width chronology studies in which the otoliths
from fewer than 50 individuals were used. A post hoc power analysis (Appendix,
Power Analysis) corroborated that a sample size of 50 is sufficient to gain
adequate statistical power for our B. frenchii chronologies. Although there is no
research standard to determine how many samples should be used when
constructing otolith-based biochronologies, sample size is likely to vary based on
the question being asked and or the geographic area over which samples are
collected. For example, Gillanders et al. (2012) used the otoliths from as few as
16 Luderick Girella tricuspidata collected over a restricted region from waters off
northern New Zealand, while Morrongiello and Thresher (2015) employed the
otoliths of 6143 Tiger Flathead Platycephalus richardsoni collected from six
fishing regions spanning eight degrees of latitude. As the samples of B. frenchii
used in the current study are from distantly populations located waters on two
coasts, the interpretation of the results have been restricted to generalities on the
influences of environmental variables on the otolith growth of individuals in those
two populations and how conditions on these two coasts drive differences in the
response in otolith growth.
Cross dating and synchrony among individuals
Cross-dating is the process of ensuring that each increment is assigned correctly to
the year of formation (Fritts 1976, Cook & Kairiukstis 1990, Maxwell et al.
2011). This is accomplished by visually inspecting the time series and looking for
conspicuously wide and/or narrow increments (Black et al. 2005). This process is
relatively easily applied in those species that are particularly sensitive to
fluctuations in environmental conditions which also is captured by conspicuous
46
variations in growth increment widths in their otoliths (e.g. Matta et al. 2010;
Coulson et al 2014; Tao et al. 2015). As in the case of A. gouldii (Rountrey et al.
2014), it may not always be possible in the case of those species, such as of B.
frenchii, whose increments have mean widths of 42-46 Āµm. In addition, B.
frenchii and A. gouldii are known as complacent, in that they do not show
significant response to any environmental change. In the case of such species, the
ability to visually align increments on the basis of their widths is not possible.
Statistical cross dating provides an avenue for the increment width time
series of individuals of such complacent, and non-complacent, species to be
statistically checked. The results of cross dating, such as the interseries correlation
and mean sensitivity, provide an indication of the quality and level of synchrony
within the increment width data. The interseries correlation (ISC), is the mean
correlation of the standardised individual series with the mean of all the others.
Values of 0.32 and 0.17 for the south and lower west coast populations indicate
that individuals of B. frenchii on the south coast respond similarly to fluctuations
in environmental conditions resulting in greater synchrony amongst their
increment widths. The ISC values for both B. frenchii populations are very low in
comparison to those values obtained in other studies, such as 0.76 for grey
snapper Lutjanus griseus (Black et al. 2011a) and 0.88 for largemouth bass
Micropterus salmoides (Rypel 2009). In the case of the south coast B. frenchii
population, the ISC value is markedly higher than 0.11 obtained for A. gouldii in
the same waters (Rountrey et al. 2014), but far less than 0.64 and 0.62 obtained
for P. laevigatus and L. inops from shallow, inshore waters along the same coast
(Coulson et al. 2014). This perhaps indicates that B. frenchii are less complacent
towards fluctuating environmental variables than A. gouldii and that, in shallow
47
water environments where variables such as water temperature exhibit greater
extremes, the individuals of species living in such waters are likely to all respond
in a similar way. This is also demonstrated by very high ISC values (0.49-0.88)
for fish in shallow, freshwater environments (Rypel 2009, Black et al. 2013b, Tao
et al. 2015).
The mean sensitivity is a measure of the range of increment width
variation. The very low values of 0.14 and 0.11 for south and lower west coast
populations indicate that the increment widths of B. frenchii do not differ
markedly between years. While these values are low are at the lower end of the
spectrum in comparison to the results from other studies, mean sensitivity values
for otolith increment width studies are typically <0.25. Like B. frenchii, the
increment widths of hapuku Polyprion oxygeneios, a deep water species found off
southern Western Australia, do not vary markedly (mean sensitivity = 0.14)
between years of formation (Nguyen et al. 2015).
Detrending
Detrending is a process, which standardises growth ring widths in chronologies to
a mean of one. The aim of detrending is to remove the age-related growth declines
in the increment width time series while preserving as much environmentally-
induced variability as possible to better illustrate climate driven anomalies in
increment widths of (Cook 1985). The type of detrending that is applied to the
increment width data will influence the final MIC and potentially any correlations
with the selected environmental variables. One of the most common types of
detrending used in otolith biochronologies is the cubic smoothing spline (Black et
al. 2005, Black et al. 2008b). In this study, three detrending methods were applied
48
and compared to determine the optimal detrending method for B. frenchii. Based
on visual inspection of detrended chronologies along with expressed population
signals it was decided that double detrending is the best detrending method for B.
frenchii. This method was also used for developing the otolith chronology for the
Western Blue Groper Achoerodus gouldii, another long-lived labrid species that
co-occurs with B. frenchii. Double detrending involves detrending the increment
width data, firstly, with a smoothing spline and then, secondly, with a negative
exponential function techniques (Rountrey et al. 2014). As separate, stand-alone
detrending methods, double detrending and cubic smoothing splines were visually
very similar and produced similar expressed population signal values. However,
upon closer inspection, the cubic smoothing spline flattened the individual
increment series for the individuals off the lower-west coast, particularly in the
early years of the chronology. The application of a negative exponential function
in conjunction with a smoothing spline greatly improved the chronology as the
double detrended chronology followed the trends in the increment widths of the
individuals off the south coast.
Synchrony between regions
Cossington et al. (2010) found regional differences between the B. frenchii
populations on the south coast and lower west coast, with those individuals on the
south coast growing faster and attaining a larger size and mass at age and
maturing earlier in life compared to those individuals on the lower west coast.
Despite these biological differences and the that fact that these two populations
are spatially separated (~1200km) and occur on coasts that are markedly different
in terms of their reef types and the environmental conditions, the response of
49
otolith growth of the individuals in those two populations to regional specific SST
was remarkably similar (Figure 9a). This demonstrates, particularly of B. frenchii
on the lower west coast close to its northern limit of distribution, that historical
increases in water temperature have only positively influenced otolith growth.
Modeling of future otolith growth of A. gouldii off the south coast of Western
Australia by Rountrey et al. (2014) demonstrated that predicted future increases in
water temperate throughout this century are only expected to positively influence
otolith growth. While Morrongiello et al. (2015) demonstrated that, throughout
the majority of its geographic range on the south east coast of Australia, the
growth of the temperate tiger flathead, P. richardsoni, responded positively to
temperature, at the northern limit of its distribution, warming waters are having a
negative impact on the growth of this species (Morrongiello & Thresher 2015).
The results from Morrongiello et al. (2015) suggest that, if future climate change
was to continue to result in increasing water temperature, the otolith growth of
typically cool-temperate species such as B. frenchii is likely to respond negatively
those future increases. Given that B. frenchii is restricted to a narrow latitudinal
range, with Rottnest Island being close to their northern limit, and recent marked
increases in water temperature off the lower west coast (Pearce & Feng 2013), the
population in this region is susceptible to future increases in water temperatures.
Financial year vs. calendar year
The main growing season for fish species in south-western Australia and, indeed,
other regions of the Southern Hemisphere at similar latitudes, extends from
approximately mid-spring (October) to early autumn (March). Therefore, the
growing period extends from the end of one calendar year to the beginning of
50
another and thus otolith growth occurs over the period that straddles two calendar
years. In contrast, the main growing period in the Northern Hemisphere falls in
the middle of the calendar year and thus the growth increment is formed over the
period of an individual calendar year. Thus, relationships between otolith
increment width chronologies constructed for fish species in the Northern
Hemisphere and environmental variables in those regions on an annual scale
(Matta et al. 2010, Black et al. 2011b, Black et al. 2013b) are correlations on over
the same temporal scale. If this same approach is taken for studies carried out in
the Southern Hemisphere, increment widths would be correlated with
environmental variables that occur after the increment has begun forming. It
therefore appears logical that, in the Southern Hemisphere, increment widths
should be correlated environmental variables on a financial year (July to June)
scale which would encompass the main growing period (October to March).
Coulson et al. (2014) used this approach when investigating relationship
between the growth of the otoliths two species of Platycephalidae and SST in
waters off southern Western Australia. While correlation tests between the MIC
for both species and mean annual (financial year) SST were significant, this was
not the case when mean annual SST was calculate on a calendar year scale. Other
studies in southern Western Australian and New Zealand waters have
demonstrated that there is a positive relationship between otolith growth and SST
in those months between mid-spring and late autumn of the previous calendar year
(Gillanders et al. 2012, Rountrey et al. 2014, Nguyen et al. 2015). However, these
researchers did not acknowledge the use of SST on a financial year scale and
instead referred to such relationships as lagged correlations with SST in months
and or seasons of the previous year. It is clear from the results of the current study
51
and previous research that, when investigating relationships between otolith
growth and environmental parameters in waters of the Southern Hemisphere,
those parameters should be considered on a financial year scale.
Relationships between MIC and environmental variables
Unlike other disciplines of sclerochronology, such as those based on the
growth bands in corals or the shells of bivalves, the results from studies
employing fish otoliths must take into consideration the fact that fish are mobile
and potentially inhabit a number of environments throughout their life cycle when
interpreting results (Morrongiello et al. 2012, Ong et al. 2015). Site fidelity and
home range estimates for the closely related California sheephead Semicossyphus
pulcher demonstrate that this species spends ~ 90% of total residence time within
a 600 m core area (Topping et al. 2006). Similarly, 10 out the 11 acoustically
tagged individuals of the co-occurring labrid, the western blue groper Achoerdus
gouldii, largely remained in a 1 km long by 40 m wide strip of coastal reef for the
12 month study period (Bryars et al. 2012). In addition, Fairclough et al. (2011)
demonstrated through otolith microchemistry analysis that, on the mid-west coast
of Western Australia, the movement of juvenile or adult baldchin groper
Choerodon rubescens occurs at relatively small spatial scales. The ability,
therefore, to detect responses in the otolith growth of B. frenchii to variations in
selected environmental variables is improved because of the likelihood that the
individuals, whose otoliths were employed in this study, have been resident in
those regions for the majority of their lives.
Because teleosts are polkiotherms, their body temperature is thus regulated
by the temperature of their surrounding environment, which has been shown to
52
positively influence otolith growth in a number of freshwater and marine species
(e.g. Rypel 2009; Matta et al. 2010; Gillanders et al. 2012; Black et al. 2013b),
including those species that inhabit south-western Australian waters (Coulson et
al. 2014; Nguyen et al. 2015). The results of the GAMMs demonstrated that,
while there was a significant, positive relationship between the MIC for the south
coast population with mean annual SST, the relationship with this variable was
more significant on the lower west coast (south P = 0.003, lower west P =
0.0006, Table 7). However, there was a higher estimated effect otolith increment
growth in increasing annual SST for the south coast population of B. frenchii
(estimate for increase of 1Ā°C (Ā±SE) = 1.25 Āµm Ā± 0.42) compared to the lower west
population (estimate for increase of 1Ā°C (Ā±SE) = 1.18 Āµm Ā± 0.35). Cossington et
al. (2010) stated that the cool south coast is more favourable for this species. As
expected, there is a higher estimated effect for otolith growth in increasing in this
area. The extent to which SST fluctuates from year to year is much greater on
along the lower west coast than the south coast (Figure 1) and, since ~ 1970, there
has been an increasing trend in the mean annual SST on that former coast.
The relationship between the MIC for B. frenchii on the south coast and
SST for is similar to the findings of Rountrey et al. (2014) for the co-occurring
A. gouldii, who found a weak, but significant influence of regional SST on the
otolith growth of A. gouldii in the south coast of Western Australia. Similarly,
Coulson et al. (2014) demonstrated that increasing water temperature has resulted
in increased otolith growth of two platycephalid species in inshore waters on the
same coast.
53
The effects of the Leeuwin Current in waters off the lower west coast of
Western Australia are well documented and Fremantle sea level (FSL) is
commonly used as a proxy for the Leeuwin Current through El Nino Southern
Oscillation cycles (Pearce & Phillips 1988, Caputi et al. 1996, Caputi et al. 2001,
Feng et al. 2008). Although there were no significant Pearson's correlations
between the MIC for the lower west coast population of B. frenchii and FSL, the
GAMM showed a highly significant relationship (P = 0.008) with an estimated
4.62 Āµm (SE = Ā± 2.05) increase in otolith increment size for every one cm
increase in Fremantle sea level. Nguyen et al. (2015) found that the Leeuwin
Current had a significant influence on the growth of hapuku Polyprion oxygeneios
in south-western Western Australia through its effect on primary productivity
eventually leading to increase prey resources for P. oxygeneios. The strength of
the Leeuwin Current weakens, as does its influence on the marine environment, as
it flows east along the south coast. It is thus not surprising that Coulson et al.
(2014) found no relationship between the otolith growth of two platycepahlids and
sea level in inshore waters along the south coast of Western Australia. The
influence of the Leeuwin Current on the otolith growth of B. frenchii is likely a
reflection of a combination of oceanographical and biological influences by the
strengthening of this current in this region.
Conclusion
In conclusion, two otolith increment-width chronologies were constructed for
Bodianus frenchii from the south and lower-west coasts of Western Australia.
While the otoliths for B. frenchii possess very clear growth increments, the lack of
conspicuous variability in widths required to undertake cross dating made this
54
process unsuccessful. The lack of synchrony in the trends displayed by the
increment widths between individuals demonstrates the insensitivity, as reflected
by the interseries correlation values, of B. frenchii to environmental fluctuations.
Despite the weak signal in the otolith increment widths, significant relationships
with regional SST and Leeuwin Current strength were present. Generalized
additive mixed modeling demonstrated that SST was an important driver of
otolith growth of B. frenchii on the south coast, while on the lower west coast, the
effects of a strong Leeuwin Current influence the otolith growth of B. frenchii in
these waters.
55
References
AbrĆ moff MD, MagalhĆ£es PJ, Ram SJ (2004) Image processing with ImageJ.
Biophotonics International 11:36-42
Black BA, Boehlert GW, Yoklavich MM (2005) Using tree-ring crossdating
techniques to validate annual growth increments in long-lived fishes.
Canadian Journal of Fisheries and Aquatic Sciences 62:2277-2284
Black BA, Boehlert GW, Yoklavich MM (2008a) Establishing climate-growth
relationships for yelloweye rockfish (Sebastes ruberrimus) in the northeast
Pacific using a dendrochronological approach. Fisheries Oceanography
17:368-379
Black BA, Gillespie DC, MacLellan SE, Hand CM (2008b) Establishing highly
accurate production-age data using the tree-ring technique of crossdating:
a case study for Pacific geoduck (Panopea abrupta). Canadian Journal of
Fisheries and Aquatic Sciences 65:2572-2578
Black BA (2009) Climate-driven synchrony across tree, bivalve, and rockfish
growth-increment chronologies of the northeast Pacific. Marine Ecology
Progress Series 378:37-46
Black BA, Schroeder ID, Sydeman WJ, Bograd SJ, Lawson PW (2010)
Wintertime ocean conditions synchronize rockfish growth and seabird
reproduction in the central California Current ecosystem. Canadian
Journal of Fisheries and Aquatic Sciences 67:1149-1158
Black BA, Allman RJ, Schroeder ID, Schirripa MJ (2011a) Multidecadal otolith
growth histories for red and gray snapper (Lutjanus spp.) in the northern
Gulf of Mexico, USA. Fisheries Oceanography 20:347-356
56
Black BA, Schroeder ID, Sydeman WJ, Bograd SJ, Wells BK, Schwing FB
(2011b) Winter and summer upwelling modes and their biological
importance in the California Current Ecosystem. Global Change Biology
17:2536-2545
Black BA, Matta ME, Helser TE, Wilderbuer TK (2013a) Otolith biochronologies
as multidecadal indicators of body size anomalies in yellowfin sole
(Limanda aspera). Fisheries Oceanography 22:523-532
Black BA, Biela VR, Zimmerman CE, Brown RJ (2013b) Lake trout otolith
chronologies as multidecadal indicators of high-latitude freshwater
ecosystems. Polar Biology 36:147-153
Boehlert GW (1985) Using objective criteria and multiple regression models for
age determination in fishes. Fishery Bulletin 83:103-117
Briffa KR, Jones PD, Schweingruber FH, KarlƩn W, Shiyatov SG (1996) Tree-
ring variables as proxy-climate indicators: problems with low-frequency
signals. Climatic variations and forcing mechanisms of the last 2000
years. Springer, Berlin Heidelberg
Bryars S, Rogers P, Huveneers C, Payne N, Smith I, McDonald B (2012) Small
home range in southern Australia's largest resident reef fish, the western
blue groper (Achoerodus gouldii): implications for adequacy of no-take
marine protected areas. Marine and Freshwater Research 63:552-563
Bunn AG (2008) A dendrochronology program library in R (dplR).
Dendrochronologia 26:115-124
Bunn AG (2010) Statistical and visual crossdating in R using the dplR library.
Dendrochronologia 28:251-258
57
Bunn AG, Korpela M, Biondi F, Campelo F, MĆ©rian P, Qeadan F, Zang C (2015)
Dendrochronology Program Library in R.
Campana SE, Neilson JD (1985) Microstructure of fish otoliths. Canadian Journal
of Fisheries and Aquatic Sciences 42:1014-1032
Campana SE (2001) Otoliths, increments, and elements: keys to a comprehensive
understanding of fish populations? Canadian Journal of Fisheries and
Aquatic Sciences 58:30-38
Campana SE, Thorrold SR (2001) Otoliths, increments, and elements: keys to a
comprehensive understanding of fish populations? Canadian Journal of
Fisheries and Aquatic Sciences 58:30-38
Caputi N, Fletcher W, Pearce A, Chubb C (1996) Effect of the Leeuwin Current
on the recruitment of fish and invertebrates along the Western Australian
coast. Marine and Freshwater Research 47:147-155
Caputi N, Chubb C, Pearce A (2001) Environmental effects on recruitment of the
western rock lobster, Panulirus cygnus. Marine and Freshwater Research
52:1167-1174
Carter R, Woodroffe C (1994) Coastal evolution: an introduction. Cambridge
University Press
Cleveland WS (1993) Visualizing data. Hobart Press
Cook ER, Peters K (1981) The smoothing spline: a new approach to standardizing
forest interior tree-ring width series for dendroclimatic studies. Tree Ring
Bulletin 41:45-53
Cook ER (1985) A time series analysis approach to tree ring standardisation
(Dendrochronloogy, Forestry, Dendroclimatology, Autoregressive
Process). Phd Thesis, Univeristy of Arizona,
58
Cook ER, Kairiukstis LA (1990) Methods of dendrochronology: applications in
the environmental sciences. Springer Science & Business Media
Cossington S, Hesp SA, Hall NG, Potter IC (2010) Growth and reproductive
biology of the foxfish Bodianus frenchii, a very long-lived and monandric
protogynous hermaphroditic labrid. Journal of Fish Biology 77:600-626
Coulson PG, Black BA, Potter IC, Hall NG (2014) Sclerochronological studies
reveal that patterns of otolith growth of adults of two co-occurring species
of Platycephalidae are synchronised by water temperature variations.
Marine Biology 161:383-393
Davis-Foust SL (2012) Long-term changes in population statistics of Freshwater
Drum (Aplodinotus grunniens) in Lake Winnebago, Wisconsin, using
otolith growth chronologies and bomb radiocarbon age validation. Doctor
of Philosophy, University of Wisconsin,
Douglass AE (1920) Evidence of climatic effects in the annual rings of trees.
Ecology 1:24-32
Douglass AE (1941) Crossdating in dendrochronology. Journal of Forestry
39:825-831
Fairclough DV, Edmonds JS, Lenanton RC, Jackson G, Keay IS, Crisafulli BM,
Newman SJ (2011) Rapid and cost-effective assessment of connectivity
among assemblages of Choerodon rubescens (Labridae), using laser
ablation ICP-MS of sagittal otoliths. Journal of Experimental Marine
Biology and Ecology 403:46-53
Feng M, Biastoch A, Bƶning C, Caputi N, Meyers G (2008) Seasonal and
interannual variations of upper ocean heat balance off the west coast of
Australia. Journal of Geophysical Research: Oceans (1978ā€“2012) 113
59
Feng M, McPhaden MJ, Xie SP, Hafner J (2013) La Nina forces unprecedented
Leeuwin Current warming in 2011. Scientific reports 3:1277
Fritts H (1976) Tree rings and climate. Elsevier
Gillanders BM, Black BA, Meekan MG, Morrison MA (2012) Climatic effects on
the growth of a temperate reef fish from the Southern Hemisphere: a
biochronological approach. Marine Biology 159:1327-1333
Grissino-Mayer HD (2001) Evaluating crossdating accuracy: a manual and
tutorial for the computer program COFECHA. Tree-Ring Research
57:205-221
Holmes RL (1983) Computer-assisted quality control in tree-ring dating and
measurement. Tree-ring bulletin 43:69-78
Howard RK (1989) The structure of a nearshore fish community of Western
Australia: diel patterns and the habitat role of limestone reefs.
Environmental Biology of Fishes 24:93-104
Kastelle CR, Helser TE, Black BA, Stuckey MJ, C. Gillespie D, McArthur J,
Little D, D. Charles K, Khan RS (2011) Bomb-produced radiocarbon
validation of growth-increment crossdating allows marine paleoclimate
reconstruction. Palaeogeography, Palaeoclimatology, Palaeoecology
311:126-135
Kendrick G (1999) Western Australia. In: Andrew N (ed) Under southern seas.
University of New South Wales Press, Sydney
Lehodey P, Grandperrin R (1996) Influence of temperature and ENSO events on
the growth of the deep demersal fish alfonsino, Beryx splendens, off New
Caledonia in the western tropical South Pacific Ocean. Deep Sea Research
Part I: Oceanographic Research Papers 43:49-57
60
List JH, Terwindt JH (1995) Large-scale coastal behavior, Vol 126. DIANE
Publishing
Matta EM, Black BA, Wilderbuer TK (2010) Climate-driven synchrony in otolith
growth-increment chronologies for three Bering Sea flatfish species.
Marine Ecology Progress Series 413:137-145
Maxwell RS, Wixom JA, Hessl AE (2011) A comparison of two techniques for
measuring and crossdating tree rings. Dendrochronologia 29:237-243
Meekan M, Wellington G, Axe L (1999) El Nino-Southern Oscillation events
produce checks in the otoliths of coral reef fishes in the GalƔpagos
Archipelago. Bulletin of Marine Science 64:383-390
Melvin TM, Briffa KR (2008) A ā€œsignal-freeā€ approach to dendroclimatic
standardisation. Dendrochronologia 26:71-86
Miller JA, Wells BK, Sogard SM, Grimes CB, Cailliet GM (2010) Introduction to
proceedings of the 4th International Otolith Symposium. Environmental
Biology of Fishes 89:203-207
Morrissey MB (2011) Exploiting natural history variation: looking to fishes for
quantitative genetic models of natural populations. Ecology of Freshwater
Fish 20:328-345
Morrongiello JR, Crook DA, King AJ, Ramsey DSL, Brown P (2011) Impacts of
drought and predicted effects of climate change on fish growth in
temperate Australian lakes. Global Change Biology 17:745-755
Morrongiello JR, Thresher RE, Smith DC (2012) Aquatic biochronologies and
climate change. Nature Climate Change 2:849-857
61
Morrongiello JR, Thresher RE (2015) A statistical framework to explore
ontogenetic growth variation among individuals and populations: a marine
fish example. Ecological Monographs 85:93-115
Nguyen HM, Rountrey AN, Meeuwig JJ, Coulson PG, Feng M, Newman SJ,
Waite AM, Wakefield CB, Meekan MG (2015) Growth of a deep-water,
predatory fish is influenced by the productivity of a boundary current
system. Scientific Reports 5
Ong JJ, Nicholas Rountrey A, Jane Meeuwig J, John Newman S, Zinke J, Gregory
Meekan M (2015) Contrasting environmental drivers of adult and juvenile
growth in a marine fish: implications for the effects of climate change.
Scientific Reports 5:10859
Pannella G (1971) Fish otoliths: daily growth layers and periodical patterns.
Science 173:1124-1127
Pearce A, Phillips B (1988) ENSO events, the Leeuwin Current, and larval
recruitment of the western rock lobster. Journal du Conseil: ICES Journal
of Marine Science 45:13-21
Pearce A, Feng M (2007) Observations of warming on the Western Australian
continental shelf. Marine and Freshwater Research 58:914-920
Pearce A, Feng M (2011) The" marine heat wave" off Western Australia during
the summer of 2010/11. Western Australian Fisheries and Marine
Research Laboratories
Pearce AF, Feng M (2013) The rise and fall of the ā€œmarine heat waveā€ off
Western Australia during the summer of 2010/2011. Journal of Marine
Systems 111-112:139-156
62
Pereira DL, Bingham C, Spangler GR, Conner DJ, Cunningham PK, Secor D,
Dean J, Campana S (1995) Construction of a 110-year biochronology from
sagittae of freshwater drum (Aplodinotus grunniens). Recent developments
in fish otolith research University of South Carolina Press, Columbia:177-
196
Popper AN, Ramcharitar J, Campana SE (2005) Why otoliths? Insights from inner
ear physiology and fisheries biology. Marine and Freshwater Research
56:497-504
Rayner NA (2003) Global analyses of sea surface temperature, sea ice, and night
marine air temperature since the late nineteenth century. Journal of
Geophysical Research 108
Rose TH, Smale DA, Botting G (2012) The 2011 marine heat wave in Cockburn
Sound, southwest Australia. Ocean Science 8:545-550
Rountrey AN (2009) Life Histories of Juvenile Woolly Mammoths from Siberia:
Stable Isotope and Elemental Analyses of Tooth Dentin. The University of
Michigan,
Rountrey AN, Coulson PG, Meeuwig JJ, Meekan M (2014) Water temperature
and fish growth: otoliths predict growth patterns of a marine fish in a
changing climate. Global Change Biology 20:2450-2458
Rypel AL (2009) Climate growth relationships for largemouth bass (Micropterus
salmoides) across three southeastern USA states. Ecology of Freshwater
Fish 18:620-628
Sanderson P, Eliot I, Hegge B, Maxwell S (2000) Regional variation of coastal
morphology in southwestern Australia: a synthesis. Geomorphology
34:73-88
63
Tao J, Chen Y, He D, Ding C (2015) Relationships between climate and growth
of Gymnocypris selincuoensis in the Tibetan Plateau. Ecology and
Evolution 5:1693-1701
Thompson JE, Hannah RW (2010) Using cross-dating techniques to validate ages
of aurora rockfish (Sebastes aurora): estimates of age, growth and female
maturity. Environmental Biology of Fishes 88:377-388
Topping DT, Lowe CG, Caselle JE (2006) Site fidelity and seasonal movement
patterns of adult California sheephead Semicossyphus pulcher (Labridae):
an acoustic monitoring study. Marine Ecology Progress Series 326:257-
267
Tukey JW Exploratory data analysis: as part of a larger whole. Proc Proceedings
of the 18th conference on design of experiments in Army research and
development I Washington, DC. DTIC Document
Tukey JW (1977) Exploratory data analysis. Reading, Ma 231:32
Wells FE, Keesing JK (1990) Population characteristics of the abalone Haliotis
roei on intertidal platforms in the Perth metropolitan area. Journal of the
Malacological Society of Australia 11:65-71
Wells FE, Walker D, Wells FE, Walker D (1993) Introduction to the marine
environment of Rottnest Island, Western Australia. In: Lethbridge R,
Lethbridge R (eds) The Marine Flora and Fauna of Rottnest Island,
Western Australia, Book 1. Western Australian Museum, Perth
Wigley TM, Briffa KR, Jones PD (1984) On the average value of correlated time
series, with applications in dendroclimatology and hydrometeorology.
Journal of Climate and Applied Meteorology 23:201-213
Wood S, Scheipl F, Wood MS (2014) Package ā€˜gamm4ā€™.
64
Zuur A, Ieno EN, Walker N, Saveliev AA, Smith GM (2009) Mixed effects
models and extensions in ecology with R. Springer
Zuur AF, Ieno EN, Elphick CS (2010) A protocol for data exploration to avoid
common statistical problems. Methods in Ecology and Evolution 1:3-14
65
Appendix
List of figures
A1. Dot plots for individual chronology time series for the 29 Bodianus frenchii
from the south coast. Each plot represents the individual points for each individual
with their identifying fish number at the top of each plot..................................... 67
A2. Dot plots for individual chronology time series for 24 Bodianus frenchii from
the lower west coast. Each plot represents the individual points for each individual
with their identifying fish number at the top of each plot..................................... 68
A3. Plots to visually assess homogeneity. Residuals vs. fitted values for the MICs
of a) south coast and c) lower west coast and box plots of individual Bodianus
frenchii increment width time series for the b) south coast and d) west coast...... 69
A4. QQ plots and residual histograms used to assess normality in the increment
width data for the B. frenchii from a, b) the south and c, d) lower west coasts.... 70
A5. Plot displaying the autocorrelation function (ACF) calculated for the
increment width data for B. frenchii on the a) south and b) lower west coasts at
lags of 0 to 30 years. The dotted line indicates the statistical significance at alpha
= 0.05. ................................................................................................................... 72
66
Data Exploration
Outliers
An "outlier" is an observation that has a relatively large or small value
(conventionally more than three standard deviations away from the mean)
compared to the majority of the data (Zuur et al. 2010). Typically, a box plot or
Cleveland dot plot (Cleveland 1993) is used to visually assess whether outliers are
present. If outliers are highlighted, the corresponding data need to be investigated.
To assess the outliers in the chronology dataset, the data was visually, inspected
for outliers. There does not appear to be any outliers in either the south (A1) or
the lower west coast samples (A2).
A1. Dot plots for individual chronology time series for the 29 Bodianus
frenchii from the south coast. Each plot represents the individual points for each
individual with their identifying fish number at the top of each plot.
67
A2. Dot plots for individual chronology time series for 24 Bodianus
frenchii from the lower west coast. Each plot represents the individual points for
each individual with their identifying fish number at the top of each plot.
68
Homogeneity
A basic assumptions for parametric statistical analyses, such as linear
regression, is that the data have equal variance or homogeneity. This can be
assessed using a residuals vs. fitted values plot (A3 a, c), or individual boxplots
(A3 b, d) to assess the spread of residuals. The spread of variances in the
B. frenchii chronology data look as though there might be a slight conical shape
(A3 a, c), caused most likely by a decrease in otolith increment width with
increasing age. This suggests that there is a variance structure present within the
data.
A3. Plots to visually assess homogeneity. Residuals vs. fitted values for
the MICs of a) south coast and c) lower west coast and box plots of individual
Bodianus frenchii increment width time series for the b) south coast and d) west
coast.
69
Normality
Normality, another assumption for parametric statistics, is assessed using
QQ plots (A4 a, c) or inspecting the spread of the residuals, to see if they are
normally distributed (A4 b, d). The normality visualization plots for the
chronology data show that the data are not normally distributed, as the data points
in QQ plots in both the south (A4 a) and lower west coast (A4 c) are up to 10
standard deviations away from the mean and the residual histograms are highly
skewed to the right (A4 b, d). Therefore, the assumption of normality for any
parametric modeling using the B. frenchii increment width data is violated.
A4. QQ plots and residual histograms used to assess normality in the
increment width data for the B. frenchii from a, b) the south and c, d) lower west
coasts.
ELLENBOYLEN_THESIS_FOXFISH
ELLENBOYLEN_THESIS_FOXFISH
ELLENBOYLEN_THESIS_FOXFISH
ELLENBOYLEN_THESIS_FOXFISH
ELLENBOYLEN_THESIS_FOXFISH
ELLENBOYLEN_THESIS_FOXFISH

More Related Content

What's hot

Abrupt variations in_south_american_mons
Abrupt variations in_south_american_monsAbrupt variations in_south_american_mons
Abrupt variations in_south_american_mons
GeorgeaMelo1
Ā 
Montane and Austin
Montane and AustinMontane and Austin
Montane and Austin
Marcel Montane
Ā 
GEOG5839.22, Paleofloods
GEOG5839.22, PaleofloodsGEOG5839.22, Paleofloods
GEOG5839.22, Paleofloods
Scott St. George
Ā 
7_PDFsam_FBA_NEWS_70_WINTER_2016
7_PDFsam_FBA_NEWS_70_WINTER_20167_PDFsam_FBA_NEWS_70_WINTER_2016
7_PDFsam_FBA_NEWS_70_WINTER_2016
Alex Seeney
Ā 
Irwin Prairie gw flow regime and drainage pp
Irwin Prairie gw flow regime and drainage ppIrwin Prairie gw flow regime and drainage pp
Irwin Prairie gw flow regime and drainage pp
Jared Rogers
Ā 

What's hot (20)

Streamflow variability of dryland rivers
Streamflow variability of dryland rivers Streamflow variability of dryland rivers
Streamflow variability of dryland rivers
Ā 
Noah, Joseph, And High-Resolution Paleoclimatology
Noah, Joseph, And High-Resolution PaleoclimatologyNoah, Joseph, And High-Resolution Paleoclimatology
Noah, Joseph, And High-Resolution Paleoclimatology
Ā 
Climate Change Effects on Water Resources and Aquatic Ecosystems
Climate Change Effects on Water Resources and Aquatic EcosystemsClimate Change Effects on Water Resources and Aquatic Ecosystems
Climate Change Effects on Water Resources and Aquatic Ecosystems
Ā 
Donā€™t call it a comeback: Studying ancient floods to prepare for future hazards
Donā€™t call it a comeback: Studying ancient floods to prepare for future hazardsDonā€™t call it a comeback: Studying ancient floods to prepare for future hazards
Donā€™t call it a comeback: Studying ancient floods to prepare for future hazards
Ā 
Climate change and its impact on the fisheries in Lake Kivu, East Africa
Climate change and its impact on the fisheries in Lake Kivu, East AfricaClimate change and its impact on the fisheries in Lake Kivu, East Africa
Climate change and its impact on the fisheries in Lake Kivu, East Africa
Ā 
IPO Antecedent Precipitation Presentation
IPO Antecedent Precipitation PresentationIPO Antecedent Precipitation Presentation
IPO Antecedent Precipitation Presentation
Ā 
Large-scale dendrochronology and low-frequency climate variability
Large-scale dendrochronology and low-frequency climate variabilityLarge-scale dendrochronology and low-frequency climate variability
Large-scale dendrochronology and low-frequency climate variability
Ā 
AAG Tampa 2014
AAG Tampa 2014AAG Tampa 2014
AAG Tampa 2014
Ā 
The decadal character of northern California's winter precipitation
The decadal character of northern California's winter precipitationThe decadal character of northern California's winter precipitation
The decadal character of northern California's winter precipitation
Ā 
Eops 2017 8_28
Eops 2017 8_28Eops 2017 8_28
Eops 2017 8_28
Ā 
Abrupt variations in_south_american_mons
Abrupt variations in_south_american_monsAbrupt variations in_south_american_mons
Abrupt variations in_south_american_mons
Ā 
Montane and Austin
Montane and AustinMontane and Austin
Montane and Austin
Ā 
Jardine, t.d. et al, 2015
Jardine, t.d. et al, 2015Jardine, t.d. et al, 2015
Jardine, t.d. et al, 2015
Ā 
Stephens_et_al
Stephens_et_alStephens_et_al
Stephens_et_al
Ā 
Expecting the unexpected: The relevance of old floods to modern hydrology
Expecting the unexpected: The relevance of old floods to modern hydrologyExpecting the unexpected: The relevance of old floods to modern hydrology
Expecting the unexpected: The relevance of old floods to modern hydrology
Ā 
Projections of Future Tropical Cyclone Activity
Projections of Future Tropical Cyclone ActivityProjections of Future Tropical Cyclone Activity
Projections of Future Tropical Cyclone Activity
Ā 
GEOG5839.22, Paleofloods
GEOG5839.22, PaleofloodsGEOG5839.22, Paleofloods
GEOG5839.22, Paleofloods
Ā 
7_PDFsam_FBA_NEWS_70_WINTER_2016
7_PDFsam_FBA_NEWS_70_WINTER_20167_PDFsam_FBA_NEWS_70_WINTER_2016
7_PDFsam_FBA_NEWS_70_WINTER_2016
Ā 
Consequences Of Stand Age And Structure On Forest Water Yield
Consequences Of Stand Age And Structure On Forest Water YieldConsequences Of Stand Age And Structure On Forest Water Yield
Consequences Of Stand Age And Structure On Forest Water Yield
Ā 
Irwin Prairie gw flow regime and drainage pp
Irwin Prairie gw flow regime and drainage ppIrwin Prairie gw flow regime and drainage pp
Irwin Prairie gw flow regime and drainage pp
Ā 

Similar to ELLENBOYLEN_THESIS_FOXFISH

Evidence for impacts by jellyfish on north sea
Evidence for impacts by jellyfish on north seaEvidence for impacts by jellyfish on north sea
Evidence for impacts by jellyfish on north sea
ratupura
Ā 
Eelgrass Poster-Amber
Eelgrass Poster-AmberEelgrass Poster-Amber
Eelgrass Poster-Amber
Amber Wolf
Ā 
Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...
Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...
Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...
Innspub Net
Ā 
DuballPoster_CERFconference2015_PortlandOR
DuballPoster_CERFconference2015_PortlandORDuballPoster_CERFconference2015_PortlandOR
DuballPoster_CERFconference2015_PortlandOR
Chelsea Duball
Ā 
Changes in climate conditions affect the phenology of seabirds 2
Changes in climate conditions affect the phenology of seabirds 2Changes in climate conditions affect the phenology of seabirds 2
Changes in climate conditions affect the phenology of seabirds 2
chris benston
Ā 
Assessing the Sardine Multispecies Fishery of the Gulf of California
Assessing the Sardine Multispecies Fishery of the Gulf of CaliforniaAssessing the Sardine Multispecies Fishery of the Gulf of California
Assessing the Sardine Multispecies Fishery of the Gulf of California
AI Publications
Ā 
Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...
Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...
Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...
Phanor Montoya-Maya
Ā 
Behavioral responses to annual temperature variationalter th.docx
Behavioral responses to annual temperature variationalter th.docxBehavioral responses to annual temperature variationalter th.docx
Behavioral responses to annual temperature variationalter th.docx
taitcandie
Ā 
morris and glasgow 2001 wb 113-202-210 AMRE
morris and glasgow 2001 wb 113-202-210 AMREmorris and glasgow 2001 wb 113-202-210 AMRE
morris and glasgow 2001 wb 113-202-210 AMRE
Jamin Glasgow
Ā 
Bahamas Research Paper
Bahamas Research PaperBahamas Research Paper
Bahamas Research Paper
Paige Barrett
Ā 
Received 26 February 2003Accepted 10 June 2003Published .docx
Received 26 February 2003Accepted 10 June 2003Published .docxReceived 26 February 2003Accepted 10 June 2003Published .docx
Received 26 February 2003Accepted 10 June 2003Published .docx
sodhi3
Ā 
Alex_Brown_MRes_thesis_compiled_21_Aug_2006
Alex_Brown_MRes_thesis_compiled_21_Aug_2006Alex_Brown_MRes_thesis_compiled_21_Aug_2006
Alex_Brown_MRes_thesis_compiled_21_Aug_2006
Alex Brown
Ā 

Similar to ELLENBOYLEN_THESIS_FOXFISH (20)

Evidence for impacts by jellyfish on north sea
Evidence for impacts by jellyfish on north seaEvidence for impacts by jellyfish on north sea
Evidence for impacts by jellyfish on north sea
Ā 
Eelgrass Poster-Amber
Eelgrass Poster-AmberEelgrass Poster-Amber
Eelgrass Poster-Amber
Ā 
Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...
Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...
Comparative study of reproduction cycle of mangrove oyster (Crassostrea gasar...
Ā 
DuballPoster_CERFconference2015_PortlandOR
DuballPoster_CERFconference2015_PortlandORDuballPoster_CERFconference2015_PortlandOR
DuballPoster_CERFconference2015_PortlandOR
Ā 
A Quantitative Study of the Productivity of the Foraminifera in the Sea
A Quantitative Study of the Productivity of the Foraminifera in the SeaA Quantitative Study of the Productivity of the Foraminifera in the Sea
A Quantitative Study of the Productivity of the Foraminifera in the Sea
Ā 
Changes in climate conditions affect the phenology of seabirds 2
Changes in climate conditions affect the phenology of seabirds 2Changes in climate conditions affect the phenology of seabirds 2
Changes in climate conditions affect the phenology of seabirds 2
Ā 
Assessing the Sardine Multispecies Fishery of the Gulf of California
Assessing the Sardine Multispecies Fishery of the Gulf of CaliforniaAssessing the Sardine Multispecies Fishery of the Gulf of California
Assessing the Sardine Multispecies Fishery of the Gulf of California
Ā 
Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...
Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...
Montoya-maya - 2009 - DYNAMICS OF LARVAL FISH AND ZOOPLANKTON IN SELECTED SOU...
Ā 
Fishery-induced Changes in Pacific Ecosystem
Fishery-induced Changes in Pacific EcosystemFishery-induced Changes in Pacific Ecosystem
Fishery-induced Changes in Pacific Ecosystem
Ā 
EOPS_2018_07_16.pdf
EOPS_2018_07_16.pdfEOPS_2018_07_16.pdf
EOPS_2018_07_16.pdf
Ā 
Phytoplankton communities and their fatty acids in winter
Phytoplankton communities and their fatty acids in winterPhytoplankton communities and their fatty acids in winter
Phytoplankton communities and their fatty acids in winter
Ā 
Artigo_Bioterra_V23_N2_08
Artigo_Bioterra_V23_N2_08Artigo_Bioterra_V23_N2_08
Artigo_Bioterra_V23_N2_08
Ā 
Length Frequency Distribution, Length-Weight Relationship and Condition Facto...
Length Frequency Distribution, Length-Weight Relationship and Condition Facto...Length Frequency Distribution, Length-Weight Relationship and Condition Facto...
Length Frequency Distribution, Length-Weight Relationship and Condition Facto...
Ā 
2016-Etroplus sur-IJFAStudies
2016-Etroplus sur-IJFAStudies2016-Etroplus sur-IJFAStudies
2016-Etroplus sur-IJFAStudies
Ā 
Behavioral responses to annual temperature variationalter th.docx
Behavioral responses to annual temperature variationalter th.docxBehavioral responses to annual temperature variationalter th.docx
Behavioral responses to annual temperature variationalter th.docx
Ā 
morris and glasgow 2001 wb 113-202-210 AMRE
morris and glasgow 2001 wb 113-202-210 AMREmorris and glasgow 2001 wb 113-202-210 AMRE
morris and glasgow 2001 wb 113-202-210 AMRE
Ā 
Bahamas Research Paper
Bahamas Research PaperBahamas Research Paper
Bahamas Research Paper
Ā 
Ijciet 10 01_065
Ijciet 10 01_065Ijciet 10 01_065
Ijciet 10 01_065
Ā 
Received 26 February 2003Accepted 10 June 2003Published .docx
Received 26 February 2003Accepted 10 June 2003Published .docxReceived 26 February 2003Accepted 10 June 2003Published .docx
Received 26 February 2003Accepted 10 June 2003Published .docx
Ā 
Alex_Brown_MRes_thesis_compiled_21_Aug_2006
Alex_Brown_MRes_thesis_compiled_21_Aug_2006Alex_Brown_MRes_thesis_compiled_21_Aug_2006
Alex_Brown_MRes_thesis_compiled_21_Aug_2006
Ā 

ELLENBOYLEN_THESIS_FOXFISH

  • 1. Ellen Boylen June 2015 This thesis is presented as part of the requirements for the Degree of Bachelor of Science in Marine Science with Honours at Murdoch University Factors affecting the growth of the Foxfish Bodianus frenchii on the south and lower west coasts of Western Australia
  • 2. 1 Declaration I declare that the work presented here is my own research conducted from March to October 2014, and has not been submitted for the award of any other degree at another tertiary institution. Elle Boylen June 2015 Cover photograph: "Curious Foxfish at Rottnest Island, Western Australia" Taken by Ellen Boylen, 2014
  • 3. 2 Abstract The Foxfish Bodianus frenchii is a long-lived labrid that is endemic to Western Australia, where it lives in shallow coastal reefs on the lower west and cooler south coasts. A hybrid of sclerochronological techniques, Pearsonā€™s correlation coefficient and additive mixed modelling was used to determine whether the pattern of growth of the otoliths of B. frenchii was related to temperature over the year, seasonally and or to the warmer months of the year. Emphasis was placed on testing the hypothesis that the strength of the relationships between otolith growth and temperature variables is greater on the lower west than south coast as actual temperatures and temperature range in any given month are greater on that coast and the temperatures varied more between years. Focus was also placed on determining whether the pattern of otolith growth of B. frenchii on the lower west coast was related to sea level height on that coast, an indicator of the strength of the Leeuwin Current. No attempt was made to relate otolith growth with sea level height on the south coast as the Leeuwin current flows well offshore on that coast. Successive growth increments in the transverse sections of the otoliths of 53 B. frenchii, comprising adjacent translucent and opaque zones, were measured and aligned according to year, using the increment width for the year of capture as the anchor. Visual examination of the time series of individual increment widths for all otoliths of individuals for each coast separately was undertaken to determine whether the widths of the increments, i.e. narrow or wide, of each otolith corresponded. The widths of the sequential increments of the different otoliths did not match, a common feature of otoliths in which all increment widths are very narrow. Further attempts to cross date the increment widths in the time series for each otolith, i.e. aimed at matching corresponding increment widths, was thus undertaken using a statistical approach. This
  • 4. 3 was also not entirely successful. Subsequent analyses thus continued to use the year of capture as the anchor aligning the otolith width of the different years for individuals on both coasts. The next step was detrending, whereby age-related declines in otolith growth are removed, while at the same time preserving any climate signals present in the increment width time series. Various detrending techniques have been used in otolith-based increment width studies. A visual assessment of the results of different types of detrending procedures demonstrated that the double detrending method provided the best fit for the B. frenchii increment width data. The detrended time series for the increment widths for each otolith of B. frenchii was used to calculate a mean index chronology (MIC) for otoliths from individuals from both the south and lower west coast populations. These MICs were then employed to determine whether the trends in the otolith increment widths were related to the environmental variables given earlier. Note that year in this study corresponds to the July (mid-winter) of one year to June (early-winter) of the following year so that each year encompasses the main growth period, i.e. spring to autumn. As each year encompasses the last two months of winter of one year and the first month of winter of the next year, it is inappropriate to test for differences between winters. The pattern of change in MICs with age for B. frenchii on the lower west and south coasts were similar, demonstrating that although actual temperatures varied appreciably between the two coasts, the pattern of growth was influenced by the similar overall trends. On the basis of Pearsonā€™s correlation coefficient, the MIC for B. frenchii on the south coast was positively correlated with mean annual SST (r = 0.32, P = 0.02) and, to a slightly greater extent, for this species on the lower west coast (r = 0.37, P = 0.007).
  • 5. 4 The MIC for B. frenchii on the south coast was not related, however, to SST when considered in the context of seasons or months. In contrast, the MIC for B. frenchii on the lower west coast was related individually to spring, summer and autumn (r = 0.39, P = 0.004) and thus to the times of year when temperatures were greatest and most growth would be expected to occur. The relationships between MICs and temperature were next explored using generalized additive mixed modeling (GAMM), which permitted the age of each fish at each increment and a random individual effect to be incorporated. The GAMM results demonstrated that the MICs for B. frenchii were positively correlated with mean annual SST for individuals on both the lower west and south coasts, with the relationship being stronger for the former coast. These results parallel those of the Pearsonā€™s correlation coefficients. Pearsonā€™s correlation coefficients demonstrated that the MIC for B. frenchii was not correlated with mean annual FSL for the population on the lower west coast and the same was true for the MIC on both a seasonal and monthly basis. On the basis of GAMM, the MIC for B. frenchii on the lower west coast was strongly (P = 0.008) related to sea level height, contrasting the result of the Pearsonā€™s correlation coefficient. The results, derived from Pearsonā€™s correlation coefficient, and by the use of GAMM validate the hypothesis that otolith growth in B. frenchii are correlated with water temperature and, to a slightly greater extent, on the lower west than south coasts. They also demonstrate a relationship between otolith growth and sea level on the lower west coast.
  • 6. 5 Acknowledgements I would like to thank my primary supervisors Peter Coulson and Ian Potter for their guidance and patience. I would also like to thank my "unofficial" supervisors Adrian Hordyk and Norm Hall for their endless technical expertise and time. Thank you to my parents Greg and Deb for their unwavering support and belief in my abilities. Special thanks to Justine Arnold and Nicholas Zebegew for the emotional support and last minute advice. Without all of you, I would not be where I am today.
  • 7. 6 Table of Contents Declaration........................................................................................................................1 Abstract .............................................................................................................................2 Acknowledgements...........................................................................................................5 Table of Contents..............................................................................................................6 List of Figures ...................................................................................................................8 List of Tables & Equations .............................................................................................10 Introduction.....................................................................................................................12 Objectives........................................................................................................19 Materials & Methods.......................................................................................................21 Sample collection............................................................................................21 Otolith preparation, imaging and increment measurement.............................22 Cross dating, detrending and otolith chronology construction .......................25 Analysis...........................................................................................................27 Data Exploration .........................................................................................29 Environmental Data ........................................................................................29 Results.............................................................................................................................31 Biological data ................................................................................................31 Cross dating and synchrony among individuals within ..................................31 Detrending.......................................................................................................35 Synchrony between regions ............................................................................38 Correlations with environmental variables .....................................................38 Sea Surface Temperature ............................................................................38 Fremantle sea level......................................................................................42
  • 8. 7 Discussion .......................................................................................................................45 Selection of otoliths and sample sizes.............................................................45 Cross dating and synchrony among individuals .............................................46 Detrending.......................................................................................................48 Synchrony between regions ............................................................................49 Financial year vs. calendar year......................................................................50 Relationships between MIC and environmental variables..............................52 References.......................................................................................................................56 Appendix.........................................................................................................................66 List of figures ..................................................................................................66 Data Exploration .............................................................................................67 Outliers........................................................................................................67 Homogeneity...............................................................................................69 Normality ....................................................................................................70 Independence...............................................................................................71 Motivation for Modeling.............................................................................73 Power Analysis................................................................................................73
  • 9. 8 List of Figures Figure 1. Mean annual values for a) sea surface temperatures (Ā°C) at Rottnest Island on the lower west coast and at Esperance on the south coast of Western Australia and b) sea level (cm) at Fremantle. .................................................................................. 18 Figure 2. Map of south-western Australia showing the location of where samples of Bodianus frenchii were collected from waters near Rottnest, on the lower west coast, and Esperance, on the south coast. Red dots denote the approximate location for which sea surface temperature data was obtained. .......................................................... 22 Figure 3. a) The sectioned otolith of a 51 year old Bodianus frenchii and a b) higher magnification image of part of the dorsal side of the otolith were increment measurements were undertaken. In a) white dots indicate the first and every tenth opaque zone and vertical black lines are indicative of the axis followed when measuring increments. In b) horizontal white lines show the position of the outer edge of each opaque zone........................................................................................................... 24 Figure 4. A visual representation of the successive increment widths of the 29 Bodianus frenchii individuals from the south coast used to assist in visually identifying synchronous patterns in those widths. Numbers on the left and right y-axis are identifying codes for individual fish. .................................................................... 33 Figure 5. A visual representation of the successive increment widths of the 24 Bodianus frenchii individuals from the lower west coast used to assist in visually identifying synchronous patterns in those widths. Numbers on the left and right y-axis are identifying codes for individual fish. .................................................................... 33 Figure 6. A visual representation of the statistical cross-dating of the standardized raw increment time series for the 29 individual Bodianus frenchii from the south coast. Blue correlates well (p-values less or equal to the user-set critical value) while potential dating problems are indicated by the red segments (p-values greater than the user set
  • 10. 9 critical value). Green lines show segments that do not completely overlap the time period and thus have no correlations calculated (Bunn 2010). ............................. 34 Figure 7. A visual representation of the statistical cross-dating of the standardized raw increment time series for the 24 individual Bodianus frenchii from the lower west coast. Blue correlates well (p-values less or equal to the user-set critical value) while potential dating problems are indicated by the red segments (p-values greater than the user set critical value). Green lines show segments that do not completely overlap the time period and thus have no correlations calculated (Bunn 2010). ............................. 34 Figure 8. Comparisons of the a, b) raw chronologies, c, d) negative exponential and e, f) cubic smoothing spline g, h) double detrending methods for individuals of Bodianus frenchii (grey lines) from the south (blue lines) and lower west (red lines) coasts of Western Australia.................................................................................................. 37 Figure 9. a) A comparison of the mean index chronologies (MIC) for Bodianus frenchii from the south (blue line) and lower west coasts (red line) and a comparison of the MIC for B. frenchii from the b) south and c) lower west coasts and mean annual sea surface temperature (black line) in those regions and comparison of the MIC for Bodianus frenchii from the lower west and d) mean annual Fremantle sea level (black line) for the years between 1954 and 2005. .............................................................................. 39 Figure 10. Three dimensional visualizations of the generalised additive mixed models displaying the relationship between a) the otolith growth of Bodianus frenchii on the south coast and SST and otolith growth of B. frenchii on the lower west coast and b) SST and c) FSL. x, environmental variable; y, mean increment chronology; z, smoother of age..................................................................................................................... 44
  • 11. 10 List of Tables & Equations Table 1. Examples of the different detrending methods employed in otolith based increment width chronology studies on fish species with varying times series length and sample size. ........................................................................................................... 15 Table 2. Length and age range, method of capture and sex of the individuals of Bodianus frenchii from the south and lower west coasts of Western Australia used in the this study. .............................................................................................................. 32 Table 3. Time series range, mean increment size, interseries correlation, mean sensitivity and standard deviation in the individual time series for Bodianus frenchii chronologies for the south and lower west coasts and both coasts combined. n = sample size......................................................................................................................... 32 Table 4. The expressed population signal (EPS) and š‘Ÿ values for the raw chronologies, and the increment widths for Bodianus frenchii on the south and lower west coasts of Western Australia after the negative exponential, spline and double detrending methods were applied. ......................................................................................................... 36 Table 5. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses) for the relationships between the MICs of Bodianus frenchii from the south and lower west coasts and mean annual (financial year) and mean seasonal sea surface temperatures (Ā°C) in waters off those coasts and mean annual and mean seasonal Fremantle sea level (cm). ...................................................................................................................... 40 Table 6. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses) for the relationships between the MICs of Bodianus frenchii from the south and lower west coasts and mean monthly sea surface temperatures (Ā°C) in those waters, and between the MIC of Bodianus frenchii from the lower west coast and mean monthly Fremantle sea level (cm). * denotes significant correlations at 0.05, ** denotes significant correlation at 0.01.................................................................................................. 40
  • 12. 11 Table 7. Results of generalised additive mixed models of the effect of increasing sea surface temperature on the growth of otoliths of Bodianus frenchii on the south and lower west coasts................................................................................................... 42 Table 8. Results of generalised additive mixed models of the effect of increasing Fremantle sea level (a proxy for Leeuwin Current strength) on the growth of otoliths of Bodianus frenchii on the lower west coast............................................................ 43 Equation 1. Generalised additive mixed model formula for B. frenchii biochronologies ............................................................................................................................... 28
  • 13. 12 Introduction The otoliths, ear bones of fish, consist of calcium carbonate and an organic matrix, which are used for sensory purposes such as balance and hearing (Campana & Neilson 1985, Popper et al. 2005). Of the three pairs of otoliths, including the lapilli and asterisci, the largest, the sagittae, have been routinely used by researchers to age individual fish (Campana 2001). The density at which calcium carbonate material is deposited, which varies during the year (Pannella 1971), results in the production of alternating opaque (slow winter growth, densely deposited material) and translucent (faster summer growth, sparsely deposited material) growth zones in the otolith. One of the unique properties of otoliths is the consistency of which calcium carbonate and an organic matrix is deposited (Campana & Thorrold 2001). As this material is deposited regularly, even during times of stress or starvation and after somatic growth has largely ceased, the growth zones in otoliths are very valuable for determining the age of fish (Morrongiello et al. 2012). Otoliths contain more information than simply the age of a fish. Early researchers showed that the distances between successive daily growth bands in otoliths could vary by manipulating the environmental conditions under which fish were held in aquaria (Pannella 1971, Campana & Neilson 1985), and that historical annual time series of growth could be obtained by measuring the widths of annuli in otoliths, as a proxy for annual fish growth (Boehlert 1985, Pereira et al. 1995, Lehodey & Grandperrin 1996). More recently, trends in the widths of successive growth increments have been used to investigate how inter-annual fluctuations in environmental variables, such as water temperature and the strength of major ocean current systems, influence the growth of fish species (Black et al. 2011a, Black et al. 2013a, Rountrey et al. 2014, Morrongiello & Thresher 2015). It is hypothesized that variations in increment widths
  • 14. 13 are produced by a number of intrinsic and extrinsic factors (Miller et al. 2010, Morrongiello et al. 2012). Intrinsic factors include genetic predispositions (Morrissey 2011), biological or physiological factors, while extrinsic factors include inter- and intra-species competition and environmental influences, such as large scale oceanographic systems such as the El Nino Southern Oscillation (Lehodey & Grandperrin 1996). The aim of previous sclerochronological (sclero = hard part, chronos = time series) work employing otoliths has been to detect relationships between the trends displayed by successive increment widths in otoliths with extrinsic factors such as air temperature (Black et al. 2013b), sea surface temperature (Coulson et al. 2014), sea bottom temperature (Black 2009), wind direction (Black et al. 2011a), oceanic upwelling (Black et al. 2011b), the El Nino Southern Oscillation (Meekan et al. 1999) and hydrological regimes of freshwater environments (Morrongiello et al. 2011, Black et al. 2013b). The relationships established between growth and environmental variables can then be used to make ecological inferences about populations, and population interactions and to forecast ecological reactions to those environmental drivers (e.g. Rountrey et al. 2014). The methods for constructing biochronologies from increment width measurements obtained from otoliths in many recent studies (i.e. Black et al. 2005, 2008a; Matta et al. 2010 Gillanders et al. 2012) have been heavily borrowed from those used to construct chronologies from annual banding in the trunks of trees (Douglass 1920, 1941), which have been developed over the past 95 years. Special emphasis has been placed, by dendrochronology (dendro = tree, chronos = time series), on the importance of the process of cross dating, i.e. ensuring that each increment is assigned correctly to the year of formation (Fritts 1976, Cook & Kairiukstis 1990, Maxwell et al. 2011). Cross dating, which is traditionally carried out by visually inspecting the
  • 15. 14 individual increment series and matching conspicuously wide and or narrow increments, is possible on these otoliths for those species that display such trends (e.g. Matta et al. 2010; Coulson et al 2014; Tao et al. 2015). However, when increments are only 40-Āµm wide, visually detecting patterns of wide and narrow increments is very difficult (Rountrey et al. 2014, Nguyen et al. 2015). In these cases, statistical cross dating, a process which mimics visual cross dating using correlation analyses, has become increasingly important. In order to tease out any climate signals in otolith increment time series, detrending, i.e. the process of standardizing growth ring widths in chronologies to a mean of one, is used to remove age-related growth declines while preserving as much environmentally-induced variability as possible to better illustrate climate-driven anomalies in ring widths (Cook 1985). The most common method of detrending in otolith-based biochronology studies is a modified exponential or curvilinear method (Table 1). Other detrending methods used include smoothing splines, double detrending and regional curve standardization (Table 1). Historically, curvilinear standardization has been used in dendrochronological studies carried out in the north-west of America, where the trees exhibit exponential growth (Cook 1985). Curvilinear standardization is inappropriate in those cases where the trend in the increment time series is not perfectly exponential or linear (Cook 1985). The smoothing spline represents a highly flexible technique, which represents a major improvement in standardizing ring-width series compared to linear or curvilinear fits (Cook & Peters 1981). Double detrending is the combination of negative exponential and smoothing spline detrending techniques, where the data are detrended with a spline, then a negative exponential function, hence it is detrended twice (Rountrey et al. 2014). Regional curve standardization is a detrending technique whereby a single increment is standardized by taking the average
  • 16. 15 increment width of a calendar year over the average increment width for an age class (Briffa et al. 1996, Melvin & Briffa 2008) and preserves low frequency variance but removes the overall trends in long time series (Melvin & Briffa 2008). This present study determined the best method for detrending chronologies for the Bodianus frenchii. hom Table 1. Examples of the different detrending methods employed in otolith based increment width chronology studies on fish species with varying times series length and sample size. Detrending method Species Time series length Sample size Curvilinear standardization red snapper Lutjanus campechanus1 gray snapper Lutjanus griseus 1 lake trout Salvelinus namaycush2 freshwater drum Aplodinotus grunniens3 28 31 21 22 30 24 17 1351 Smoothing splines splitnose rockļ¬sh Sebastes diploproa4 yelloweye rockļ¬sh Sebastes ruberrimus5 49 50 50 66 Double detrending western blue groper Achoerodus gouldii6 51 56 Regional curve standardization rock flathead Platycephalus laevigatus7 longhead flathead Leviprora inops 7 yellowfin sole Limanda aspera8 14 12 21 96 120 17 Black et al. 4 2005, 5 2008a,1 2011a, 2 2013b, 8 2013a, 3 Davis-Foust, 2012; 6 Rountrey et al. 7 2014; Coulson et al. 2014 The Foxfish Bodianus frenchii is a medium-sized, shallow-water, temperate reef fish of considerable longevity, attaining a maximum age of 78 years (Cossington et al. 2010), making it the longest-lived species within the speciose Labridae family. A comprehensive study of the biology of B. frenchii collected over reefs on the lower west and south coasts of Western Australia demonstrated that this protogynous hermaphroditic labrid attains a greater length and body mass at age throughout life in waters on the cooler south coast than on the warmer lower west coast and that females mature at an earlier age on the former coast and that those females exhibit substantially greater fecundity (Cossington et al. 2010). Bodianus frenchii, endemic to temperate Australian waters, is most common off the lower west and south coasts of Western
  • 17. 16 Australia, where it co-occurs with another long-lived, but much larger labrid, the Western Blue Groper Achoerodus gouldii (Cossington et al. 2010). The otolith growth of A. gouldii, a species which attains a maximum age of 70 years, was shown to be positively and significantly correlated with sea surface temperatures (SST) off the south coast of Western Australia in the months from November to May i.e. the time of year of increasing and elevated water temperatures and thus the main growing period (Rountrey et al. 2014). This is similar to the period for which there is a positive relationship between SST and otolith growth for two species of flathead (Platycephalidae) in the same region (Coulson et al. 2014, Rountrey et al. 2014). In deep water off south-western Australia, the growth of Hapuku Polyprion oxygeneios was found to also have a significant, positive relationship with SST in winter and spring months of the previous year (Nguyen et al. 2015). In neither the study involving A. gouldii nor the two platycepahlids was there any correlation between otolith growth and the strength of the Leeuwin Current, a dominant oceanographic feature along the west Australian coast. However, Nguyen et al. (2015) found a lagged positive relationship between the growth of Hapuku Polyprion oxygeneios and Leeuwin Current strength in the previous calendar year. The lag in the effect of the Leeuwin Current on the growth of P. oxygeneios was thought to be a result of the time required for their main (> 50%) prey, squid, most likely arrow squid Nototodarus gouldi, to attain a size where they form a substantial part of the hapuku diet (Nguyen et al. 2015). The near shore waters of the lower west and south coasts of Western Australia differ markedly in their habitat structures, reef types and water depths. It is generally accepted that the habitat structure within a region affects the biological community structure and that habitat structure is driven by physical oceanography within a region (Carter & Woodroffe 1994, List & Terwindt 1995). The near-shore waters of the west
  • 18. 17 coast are protected by a semi-continuous, limestone-based reef and barrier island system, which moderates wave energy (Howard 1989, Wells et al. 1993, Sanderson et al. 2000). In contrast, granite boulder reefs and headlands dominate near-shore waters of the south coast which is fully exposed to the southern ocean and is a high wave-energy coast (Wells & Keesing 1990, Wells et al. 1993, Kendrick 1999, Sanderson et al. 2000). In addition, the waters of the south coast increase in depth far more rapidly with distance from shore than is the case with waters off the lower west coast (Kendrick 1999, Sanderson et al. 2000). Another important difference between the two coasts is their temperature regimes, with average temperature, and the range in average temperature, on the lower west coast being greater than on the south coast (Kendrick 1999, Sanderson et al. 2000) (Figure 1). An important contributor to this difference is not only lower latitude, but also the influence of the pole-ward flowing, warm waters of the Leeuwin Current on this coast. The Leeuwin Current has a strong influence on the distribution and abundance of marine flora and fauna in south western Australia (Pearce & Phillips 1988, Caputi et al. 1996), and in recent years has led to marine heat wave-like conditions in waters off the lower west coast (Pearce & Feng 2011, Rose et al. 2012, Feng et al. 2013, Pearce & Feng 2013). Fremantle sea level (FSL) has been used as a proxy for the strength of the Leeuwin Current, i.e. strong current strength results in higher FSL while weak current strength results in lower FSL, which, until recently, has been closely correlated with the recruitment of rock lobster to reefs along the lower west coast (Pearce & Phillips 1988, Caputi et al. 2001).
  • 19. 18 Figure 1. Mean annual values for a) sea surface temperatures (Ā°C) at Rottnest Island on the lower west coast and at Esperance on the south coast of Western Australia and b) sea level (cm) at Fremantle. Objectives The species that is the focus of this thesis is a long-lived, temperate fish species that lives in a region experiencing increases in temperature (Pearce & Feng 2007, Pearce & Feng 2011). Previous otolith-based biochronology studies have demonstrated that water temperature during the summer months, and thus those when the majority of growth is occurring, has an important influence on the growth of fish in Western Australia
  • 20. 19 (Coulson et al. 2014; Rountrey et al. 2014). Those studies were restricted, however, to fish populations along the south coast of Western Australia. The current study thus investigates otolith growth of B. frenchii from the lower west as well as south coast waters, it also employs the otoliths of this species caught in waters off the lower west coast, where temperatures are greater and have increased since 1970. Furthermore, the influence of the Leeuwin Current is greater on the lower west coast and occurs more inshore than on the south coast. The first aim of this study was to develop a biochronology, using growth increment widths in otoliths, for the long-lived B. frenchii on both the south and lower west coasts in south-western Australia, which could be used for the following purposes. 1) Determine whether the patterns of otolith growth of the two populations are synchronized, thereby implying that the growth of individuals of this species in these two distantly-located populations respond similarly to environmental influences. 2) Test the hypotheses that, the waters off the lower west coast of Australia exhibit greater inter-annual fluctuations in temperature than off the south coast of Western Australia, the relationship between the otolith increment width chronology for B. frenchii will be stronger in the waters of the former coast. 3) Test the hypotheses that, as the Leeuwin Current exhibits a substantial influence on waters along the west coast of Australia as far south as the lower west coast, it will have an effect on the pattern of otolith growth of B. frenchii in these waters.
  • 21. 20 Materials & Methods Sample collection Bodianus frenchii, whose otoliths were used in this current study, were collected between April 2004 and 2006 as part of an earlier study on the biological characteristics of this species in south-western Australia (Cossington et al. 2010). In that study, individuals were collected from numerous locations in marine waters between Jurien Bay at ~ 30Ā°18ā€™S, 115Ā°02ā€™E on the lower west coast and from Esperance at ~ 33Ā°51ā€™S, 121Ā°53ā€™E on the south coast (Figure 2). To enable the current study to investigate the influence of environmental variables on the growth of B. frenchii at the northern and southern extent of this species distribution for which sufficient samples were available, B. frenchii sampled from Rottnest Island at ~ 32Ā°00ā€™S, 115Ā°52ā€™E (hereafter referred to ā€˜lower west coastā€™) and Esperance (hereafter referred to ā€˜south coastā€™) were selected for otolith increment analysis. Bodianus frenchii from Rottnest Island were sampled by rod and line angling and spear fishing while SCUBA diving, whereas those from Esperance where sampled by spear fishing while snorkeling and collected from recreational line and commercial gillnet fishers (Table 2). One of the aims of this study is to investigate the influence of fluctuating environmental conditions on the growth of B. frenchii (see earlier). In order to maximize temporal resolution of the data and the greater certainty of resultant correlation tests that is afforded by using long time series (Wigley et al. 1984), i.e. a series of successive increment measurements from an individual, the sectioned otoliths from the oldest individuals from each of the two coasts were preferentially selected. In addition, only those otoliths from the oldest fish, whose increment boundaries could be clearly defined, allowing for the most precise increment measurements, were employed for otolith chronology construction.
  • 22. 21 Figure 2. Map of south-western Australia showing the location of where samples of Bodianus frenchii were collected from waters near Rottnest, on the lower west coast, and Esperance, on the south coast. Red dots denote the approximate location for which sea surface temperature data was obtained. Otolith preparation, imaging and increment measurement The otoliths employed during the current study were previously sectioned for age determination (Cossington et al. 2010). This involved embedding one otolith from each individual in clear epoxy resin before cutting a thin, ~ 300 Ī¼m thick transverse section, using a low-speed diamond saw (Buehler), through the otolith primordia. The surface of the sections were then lightly polished using fine wet and dry carborundum paper (grade 1200) and mounted on slides using DePX mounting medium with a cover slip. Multiple images of the dorsal side of the sectioned otoliths were taken at a
  • 23. 22 magnification of 10X using an Olympus DP70 12.0 megapixel digital camera mounted on an Olympus BX51 stereo microscope. These images were then stitched together using the stitch function in Leica Application Suite V. 4.3 (Leica Microsystems). The measurement of the widths of successive growth increments on the digital images of sectioned otoliths was carried out using ImageJ V. 1.47 (AbrĆ moff et al. 2004) and employing the plugin ā€˜IncMeasā€™ (Rountrey 2009). As the width of an increment at one particular location on an otolith may vary from the width of that same increment measured elsewhere, increments were measured along three to five transects in order to obtain a more precise, otolith-wide measurement of that increment. The average of these transects was used as the time series for an individual. Increments were measured to the nearest 1 Āµm along transects that were drawn perpendicularly to the otolith increments on the digital images of the sectioned otoliths, i.e. parallel to growth direction (Figure 3a). Starting at the outermost opaque zone, the outer edge of each individual opaque zone was marked until the edges of those opaque zones where no longer clearly defined, after which increment measurement ceased (Figure 3b). This often meant that the increments representing the growth undertaken in the first 5 years of life were not measured. Partial increments on the peripheral side of the outermost opaque zone were not measured.
  • 24. 23 Figure 3. a) The sectioned otolith of a 51 year old Bodianus frenchii and a b) higher magnification image of part of the dorsal side of the otolith were increment measurements were undertaken. In a) white dots indicate the first and every tenth opaque zone and vertical black lines are indicative of the axis followed when measuring increments. In b) horizontal white lines show the position of the outer edge of each opaque zone.
  • 25. 24 Cross dating, detrending and otolith chronology construction Cross dating is an important process that relies on the assumption that there are synchronous growth patterns in multiple individuals due to the influence of common environmental drivers of growth (Black et al. 2005). By matching growth patterns among individuals, missing or false increments can be identified and taken into account when assigning calendar years, ensuring that each increment is assigned to the correct year of formation (Kastelle et al. 2011). While visual cross dating has been possible for some fish species (Black et al. 2005, Gillanders et al. 2012, Coulson et al. 2014), it was not possible for B. frenchii, as was the case for Achoerodus gouldii (Rountrey et al. 2014), due to the very small increment widths, ~ 10-20 Āµm, thus making it difficult to visually detect any changes in those widths. As the date of capture of the individuals used in this study was known, this served as an anchor for each increment measurement series and accordingly back-dated. Statistical cross-dating was performed using the dendrochronology program library dplR in R (Bunn 2010) following the methods outlined in Bunn et al. (2015). This cross dating program was used to fit each series of increment measurements with a highly flexible cubic smoothing spline using a 15-year moving window. Each time series was then divided by the values predicted by the spline function, removing low-frequency variability and standardizing all measurement time series to a mean of one. The detrended time series for each individual was then correlated, at segment lengths of 20 years, with a lag of 10 years, with the average for all other standardised time series of measurements for all individuals. The mean, calculated as a Tukey's biweight robust mean, was reported as the inter-series correlation. This approach of isolating only the high frequency, serially-independent growth pattern mathematically mimicked the process of visual cross-dating (Holmes 1983, Grissino-Mayer 2001). dplR also calculated the mean sensitivity, a measure of
  • 26. 25 the relative change in increment width between successive years that ranges from a minimum of zero, i.e. two increments with the same width, to a theoretical maximum of two, i.e. a pair of increments in which one value is zero (Fritts 1976). The corresponding otoliths whose time series were highlighted in the output by the dplR program to have potential errors, were visually expected, and if no error was found the increment measurement was not adjusted. To develop the mean index chronology (MIC) for climate analysis, each set of the original growth-increment measurement time series was detrended employing double detrending method outlined in Rountrey et al. (2014). The detrending effectively removed the rapid ontogenetic decline in growth rate that occurs in early life, and removed other low frequency variation including departures from the negative exponential fit. Additional detrending techniques were also explored such as negative exponential detrending (Black 2009, Black et al. 2010, Matta et al. 2010, Black et al. 2011a, Gillanders et al. 2012) and cubic smoothing splines (Black et al. 2005). All detrending analyses were carried out employing the R package dplR (Bunn 2008) and only those years with a sample depth of > five individuals were retained to ensure a true mean is calculated (Matta et al. 2010, Black et al. 2013a).
  • 27. 26 Analysis For the analysis the chronologies of B. frenchii a hybridization of sclerochronology techniques, Pearson's correlation analysis and additive mixed-effects modeling was used, similar to Rountrey et al (2014), who investigated the relationship between otolith growth of the co-occurring A. gouldii and environmental variables on the south coast. The analysis can be broken down into three parts: 1) data exploration, 2) Pearson's correlation analyses, and 3) modeling otolith growth, via mean index chronology (MIC), using a generalized additive mixed model, with age as a smooth function and an individual random effect. Reasoning for using the GAMM approach is outlined in the Appendix. To determine a parsimonious predictive relationship between the MICs for B. frenchii on the south and lower west coasts and the mean annual sea surface temperature in waters of those regions and Fremantle sea level, generalised additive mixed models (GAMMs) were used. This resulted in three models, each model was created, using one predictor, either mean annual sea surface temperature (for both the south coast and lower-west coast) or mean annual Fremantle sea level (lower-west coast). Models were created using gamm4 in R (Wood et al. 2014).
  • 28. 27 To investigate the effects of variations in sea surface temperature on B. frenchii growth, we created a generalised additive mixed model in the form of: Equation 1. Generalised additive mixed model formula for B. frenchii biochronologies š‘”(š‘€š¼š¶š‘–,š‘— š‘ ) = š›½0 + š‘“1(š“š‘”š‘’š‘–,š‘—) + š‘‹š‘— + š‘§š‘– š‘‡ š‘ where: š‘”(š‘€š¼š¶š‘–,š‘— š‘ ) is a monotonic differential link function of fish otolith width, where MIC(i, j) where i is the MIC increment and j is year relating to increment; š›½0 is the intercept parameter; š‘“1(š“š‘”š‘’š‘–š‘—) is a centered twice-differentiated smooth function for fish age; š‘‹š‘— is the average annual (financial calendar) environmental variable for year j; š‘§š‘– š‘‡ š‘ is an error term; where š‘ are random effects, assumed to be distributed ~š‘{0, š·( šœƒ)} where šœƒ is a š‘ Ɨ 1 vector of variance components
  • 29. 28 Data Exploration Data exploration is the first step in data analysis (Tukey 1977, Zuur et al. 2010) with the objective of summarizing the dataset's main characteristics, often with visual methods. Data exploration is described as "graphical detective work" (Tukey 1972) of factors, such as outliers, unequal variance, or dependence structures, which affect the type of analysis which can be performed. The results of the data exploration can be found in the Appendix. Environmental Data A range of environmental variables and climate indices (i.e. bottom temperature, wind direction, oceanic upwelling, El Nino Southern Oscillation Index or Multivariate ENSO Index, and Dipole Mode Index) were considered in order to determine their influence on the growth of B. frenchii in south-western Australia. However, many of these variables were not available throughout much of the early years of the MIC on both the west and south coasts. The variables which were selected, were also the most likely to influence the growth of fish in south-western Australia, were sea surface temperature (SST) and Fremantle sea level (FSL), a measure of the strength of the Leeuwin Current. No attempt was made to relate otolith growth with sea level height on the south coast as the Leeuwin current flows well offshore on that coast. Mean monthly SST data was collected from Hadley Satellites and obtained from the Met Office Marine Data Bank (Rayner 2003). Mean monthly SST data was obtained for every month between January 1938 and December 2005 from the grid cell off the lower-west coast (-32Ā°, 116ā€™E) and south coast (-35Ā°, 119ā€™E). Fremantle sea level was obtained from the Bureau of Meteorology. Pearsonā€™s correlation coefficients (r) were tested to determine whether there was a significant relationship between the B. frenchii
  • 30. 29 MIC and mean annual, seasonal and monthly SST data for each coast and mean annual, seasonal and monthly FSL. In the present study, correlation tests between otolith increment widths and environmental variables were conducted with those variables organized on a financial year scale, rather than calendar year. The financial year is preferred to the calendar year, because it follows more closely the biological growth year of teleosts in the southern hemisphere (Coulson et al. 2014), i.e. the peak growing period is summer (December - February) which would be divided if the traditional calendar year scale was employed.
  • 31. 30 Results Biological data The ages of individuals used in this chronology study ranged from 39 - 78 years for the south coast and 41 - 62 years for the west coast, which corresponded to lengths ranging from 372 - 435 and 315 - 446 mm, respectively (Table 2). As B. frenchii is a protogynous hermaphrodite that changes sex at ~ 29 years (Cossington et al. 2010), all but one of the individuals from each coast that were used in this chronology study were male (Table 2). Cross dating and synchrony among individuals within The increments measured encompassed years between 1938 and 2005 for the south coast and 1947 and 2005 for the west coast. A sample depth of at least six occurred in those years between 1954 and 2005 and 1954 and 2004 on south and lower west coasts, respectively. Visually, there is no pattern of variation in the widths of the increments in the otoliths of B. frenchii within and between the otoliths of the individuals on both coasts (Figure 4, Figure 5). As visual cross dating was not possible due to the very small increment widths (Table 3) increasing the difficultly in identifying synchrony in conspicuously wide and/or narrow increments, statistical cross dating was employed, with the capture year used as the anchor for this process. Although there was some synchrony among a few individuals on each coast (indicated by the blue segments in Figure 6 and Figure 7) shown by the statistical cross dating, on the whole, this process also demonstrated a lack of synchrony among those same individuals (indicated by the red segments in Figure 6 and Figure 7). Despite this apparent lack of synchrony identified by statistical cross dating, the interseries correlation (ISC), a measure of how much correlation exists amongst the
  • 32. 31 increment widths (Grissino-Mayer 2001), demonstrates that there is synchrony among the standardised increment time series for those individuals of B. frenchii on the south coast, and to a lesser extent, the lower west coast (Table 3). The very low mean sensitivity values, a measure of the change in increment width between successive years, of 0.14 for the south coast and 0.01 for the lower west coast indicate that there is very little variation between successive increment widths, which is probably why visual and statistical cross dating failed to find any synchrony (Table 3). Table 2. Length and age range, method of capture and sex of the individuals of Bodianus frenchii from the south and lower west coasts of Western Australia used in the this study. Length range (mm) Age range (years) Method of capture Sex Gillnet Line Spear F M South 372-435 39-78 8 10 11 1 28 Lower west 315-446 41-62 20 4 1 23 Table 3. Time series range, mean increment size, interseries correlation, mean sensitivity and standard deviation in the individual time series for Bodianus frenchii chronologies for the south and lower west coasts and both coasts combined. n = sample size. Time series (years) Mean increment size (Āµm) Interseries correlation Mean sensitivity Standard deviation n South 33-68 46 0.32 0.14 4.87 29 Lower west 33-58 42 0.17 0.11 3.02 24 Combined coasts 31-59 44 0.19 0.12 4.15 53
  • 33. 32 Figure 4. A visual representation of the successive increment widths of the 29 Bodianus frenchii individuals from the south coast used to assist in visually identifying synchronous patterns in those widths. Numbers on the left and right y-axis are identifying codes for individual fish. Figure 5. A visual representation of the successive increment widths of the 24 Bodianus frenchii individuals from the lower west coast used to assist in visually identifying synchronous patterns in those widths. Numbers on the left and right y-axis are identifying codes for individual fish.
  • 34. 33 Figure 6. A visual representation of the statistical cross-dating of the standardized raw increment time series for the 29 individual Bodianus frenchii from the south coast. Blue correlates well (P less or equal to the user-set critical value) while potential dating problems are indicated by the red segments (P greater than the user set critical value). Green lines show segments that do not completely overlap the time period and thus have no correlations calculated (Bunn 2010). Figure 7. A visual representation of the statistical cross-dating of the standardized raw increment time series for the 24 individual Bodianus frenchii from the lower west coast. Blue correlates well (P less or equal to the user-set critical value) while potential dating problems are indicated by the red segments (P greater than the user set critical value). Green lines show segments that do not completely overlap the time period and thus have no correlations calculated (Bunn 2010).
  • 35. 34 Detrending A range of detrending methods that are most commonly used in otolith increment width-based chronology studies were employed in order to determine which of these was most suitable to apply to the B. frenchii increment width data. The negative exponential detrending method is a rigid function that removes age related growth declines in the increment series that are more prevalent in tree ring data, for which it was developed. The negative exponential detrending method does not appear to have has not completely removed the age-related decline in the otolith increment-width data for B. frenchii, as there is still a declining trend in the time series (Figure 8c, d). The cubic smoothing spline method (often termed 'spline'), which is by far the most common in dendrochronology, but not so in otolith sclerochronology, works by taking a mean of the surrounding four data points, in front, behind, above and below. Double detrending also employs a cubic-smoothing spline after the increment width data is initially detrended by applying a negative exponential function (Cook & Kairiukstis 1990). When applied to the increment width data for B. frenchii from each coast, both the spline and double detrending methods have largely removed any age related decline in the increment width data (Figure 8e, f, g, h). In addition, the detrended time series also display a homogeneous variance around a mean of 1. Furthermore, these detrending methods, while eliminating the age related decline in otolith growth, have preserved the high frequency climate effects, which is particularly evident in the south coast B. frenchii data where there is conspicuous decrease in increment width in the early 1990s (Figure 8e, g). The higher expressed population signal (EPS), a measure of how well the sample means represent the mean of the theoretical population (Wigley et al. 1984), provided by negative exponential detrending suggests that this best detrending option for the B. frenchii increment width data (Table 4). Chronologies with EPS values above
  • 36. 35 the theoretical threshold of 0.85 or high š‘ŸĢ… values are generally considered to be acceptable; however, detrending with the aim of maximizing these indicator parameters may result in the retention of aspects of an ontogenetic trend (Nguyen et al. 2015). Thus, despite the negative exponential detrending option providing the highest EPS and š‘ŸĢ… values, visually it was still shown to retain an ontogentic decline and was therefore discarded. The EPS and š‘ŸĢ… values for the spline and double detrending methods for the increment data for south coast population were identical (Table 4). As the EPS and š‘ŸĢ… values for double detrending of the increment data for lower west coast population were marginally higher than those obtained employing the spline detrending, and to enable comparisons with the previous otolith chronology study of the Western Blue Groper Achoerodus gouldii in the same region (Rountrey et al. 2014), double detrending was chosen as the detrending method for B. frenchii on the south and lower west coasts. Table 4. The expressed population signal (EPS) and š‘ŸĢ… values for the raw chronologies, and the increment widths for Bodianus frenchii on the south and lower west coasts of Western Australia after the negative exponential, spline and double detrending methods were applied. Raw Negative Exponential Spline Double Detrending South EPS 0.854 0.878 0.697 0.697 š‘ŸĢ… 0.221 0.259 0.100 0.100 Lower West EPS 0.741 0.821 0.421 0.433 š‘ŸĢ… 0.131 0.195 0.037 0.039
  • 37. 36 Figure 8. Comparisons of the a, b) raw chronologies, c, d) negative exponential and e, f) cubic smoothing spline g, h) double detrending methods for individuals of Bodianus frenchii (grey lines) from the south (blue lines) and lower west (red lines) coasts of Western Australia.
  • 38. 37 Synchrony between regions The MICā€™s for the two B. frenchii populations were significantly correlated (r = 0.15, P= 0.005), indicating that there is a high level of synchrony in the otolith growth of the individuals of these two populations (Figure 9a). This finding suggests that B. frenchii on the south and lower west coasts respond in a similar way to variations in temperature, despite the fact that these two populations are spatial separated (~1200km) and occur on coasts that are markedly different in terms of their reef types and the environmental conditions experienced. Even though synchrony in the trends of increment widths is present between the two B. frenchii populations, in order to investigate the relationships between otolith growth of individuals in those two populations and region specific environmental variables, the MICs have not been combined to generate a single MIC. Correlations with environmental variables Sea Surface Temperature On the basis of Pearsonā€™s correlation coefficient, the MIC for B. frenchii on the south coast was weakly and positively correlated with mean annual (financial year) SST (r = 0.32, P = 0.02) in those years between 1954 and 2005 (Table 5). There was a stronger positive relationship (r = 0.37, P = 0.007) between the MIC for B. frenchii on the lower west coast mean annual (financial year) SST (Table 5). This positive relationship on an annual level for the lower west coast was largely driven by the significant positive correlation (r = 0.39, P = 0.004) with mean seasonal SST for summer, the time of year when temperature is highest and when the majority of growth is expected to occur (Table 6).
  • 39. 38 Figure 9. a) A comparison of the mean index chronologies (MIC) for Bodianus frenchii from the south (blue line) and lower west coasts (red line) and a comparison of the MIC for B. frenchii from the b) south and c) lower west coasts and mean annual sea surface temperature (black line) in those regions and comparison of the MIC for Bodianus frenchii from the lower west and d) mean annual Fremantle sea level (black line) for the years between 1954 and 2005.
  • 40. 39 Table 5. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses) for the relationships between the MICs of Bodianus frenchii from the south and lower west coasts and mean annual (financial year) and mean seasonal sea surface temperatures (Ā°C) in waters off those coasts and mean annual and mean seasonal Fremantle sea level (cm). South West Financial SST Annual 0.32 (0.0216*) 0.37 (0.0075**) Spring 0.06 (0.6725) 0.30 (0.0355*) Summer 0.20 (0.1532) 0.39 (0.0042**) Autumn 0.22 (0.1172) 0.32 (0.0204*) Financial FSL Annual 0.18 (0.2309) Spring -0.01 (0.9334) Summer 0.18 (0.2154) Autumn 0.11 (0.4387) Table 6. Pearsonā€™s correlation coefficients (r) and their P values (in parentheses) for the relationships between the MICs of Bodianus frenchii from the south and lower west coasts and mean monthly sea surface temperatures (Ā°C) in those waters, and between the MIC of Bodianus frenchii from the lower west coast and mean monthly Fremantle sea level (cm). * denotes significant correlations at 0.05, ** denotes significant correlation at 0.01. South West SST July 0.14 (0.3136) 0.30 (0.0316*) August 0.10 (0.4827) 0.27 (0.0575) September 0.09 (0.5398) 0.29 (0.0419*) October 0.11 (0.4502) 0.41 (0.0026**) November 0.21 (0.1331) 0.29 (0.0409*) December 0.15 (0.2769) 0.19 (0.1793) January 0.14 (0.3360) 0.37 (0.0081**) February 0.23 (0.1054) 0.35 (0.0110*) March 0.25 (0.0683) 0.29 (0.0357*) April 0.19 (0.1769) 0.29 (0.0366*) May 0.16 (0.2487) 0.30 (0.0347*) June 0.25 (0.0712) 0.28 (0.0450*) FSL July -0.01 (0.9494) August 0.05 (0.7482) September 0.15 (0.2832) October 0.15 (0.2934) November 0.15 (0.2832) December 0.05 (0.7330) January 0.22 (0.1306) February 0.14 (0.3399) March 0.11 (0.4582) April 0.16 (0.2555) May 0.03 (0.8158) June 0.14 (0.3224)
  • 41. 40 Generalised additive mixed models employing a smoother for age at which the increment was formed and a random intercept for each individual demonstrated that there is a significant relationship between mean annual SST and otolith growth of B. frenchii on both the south (P = 0.003, Figure 9b) and lower west (P = 0.0006, Figure 9c) coasts. Although, the relationship has a higher significance on the lower-west coast, there was a higher estimated effect for the south coast. Thus, for B. frenchii on the south coast, it is estimated that for any given year, every 1Ā°C increase in SST the width of the increment corresponding to that year will be 1.25Āµm (SE Ā± 0.42) wider, while the width of this increment in the otolith of an individual off the lower west coast is estimated to be 1.18Āµm (SE Ā± 0.35) wider (Table 7). Both the south (-14.5 Ā± 2.9) and lower west (-14.6 Ā± 2.8) coasts had similar estimates for the smoother of age (Table 7). This equates to the otolith increment width decreasing at a rate of -14.5 and -14.6 Ī¼m for the south and lower west coasts, respectively, for every increasing year of age of the fish. The GAMM combines both the effects of the predictor and the smoother of the age to model the growth of B.frenchii otolith increments. This can be better visualized using 3D models (south coast: Figure 10b, lower-west coast: Figure 10c).
  • 42. 41 Table 7. Results of generalised additive mixed models of the effect of increasing sea surface temperature on the growth of otoliths of Bodianus frenchii on the south and lower west coasts. South coast Random effects Variance SD FishID (Intercept) 6.483 2.546 Penalized component of age smooth 30832.695 175.592 Residual 11.912 3.451 Fixed effects Estimate SE t P Intercept -3.7223 7.6025 -0.490 0.625 Annual (financial year) SST 1.2538 0.4214 2.975 0.003 ** Unrealized component of age smooth -14.5241 2.8546 -5.088 <2e-16 *** Lower west coast Random effects Variance SD FishID (Intercept) 14.93 3.864 Penalized component of age smooth 79763.31 282.424 Residual 15.78 3.973 Fixed effects Estimate SE t P Intercept -4.113 7.018 -0.586 0.557917 Annual (financial year) SST 1.183 0.346 3.419 0.000648*** Unrealized component of age smooth -14.623 2.786 -5.248 <2e-16 *** Fremantle sea level Pearsonā€™s correlation coefficients demonstrated that the MIC for the B. frenchii population on the lower west coast was not correlated with mean annual, mean seasonal or mean monthly FSL (Tables 5, 6). In contrast, on the basis of GAMM, the MIC for B. frenchii on the lower west coast was positively related to mean annual FSL (Table 8). This model also estimated that, for every given year, every 1 cm increase in mean annual Fremantle sea level will result in a 6.34Āµm increase in the width of the increment for that year, for individuals on the lower-west coast.
  • 43. 42 Table 8. Results of generalised additive mixed models of the effect of increasing Fremantle sea level (a proxy for Leeuwin Current strength) on the growth of otoliths of Bodianus frenchii on the lower west coast. t Coast Random effects Variance SD FishID (Intercept) 14.48 3.805 Penalized component of age smooth 67003.44 258.850 Residual 15.64 3.955 Fixed effects Estimate SE t P Intercept 14.875 1.891 7.866 8.44e-15 *** Financial Annual FSL 6.339 2.377 2.667 0.00777 ** Unrealized component of age smooth -14.596 2.783 -5.244 <2e-16 *** Visualizing GAMMs GAMMs combine both the effects of the predictor (i.e. the respective environmental variables = x), the smoother of the age (z) to model the growth of B.frenchii otolith increments (y = MIC). This can be better visualized using three dimensional visualizations of the models (south coast SST model: Figure 10a, lower west coast SST model: Figure 10b, and lower west coast FSL model: Figure 10c).
  • 44. 43 Figure 4. Three dimensional visualizations of the generalised additive mixed models displaying the relationship between a) the otolith growth of Bodianus frenchii on the south coast and SST and otolith growth of B. frenchii on the lower west coast and b) SST and c) FSL. x, environmental variable; y, mean increment chronology; z, smoother of age.
  • 45. 44 Discussion Selection of otoliths and sample sizes Bodianus frenchii is a protogynous labrid of exception longevity, attaining a maximum age of 78 years (Cossington et al. 2010). For this study, in order to maximize the chronology length, which increases the strength and validity of potential correlations with environmental variables, the otoliths of the oldest individuals were preferentially selected. The final number of otoliths, whose increment widths were measured for the analyses, were selected, largely, on the basis of the clarity of increment boundaries across a broad region of the dorsal side of the sectioned otolith allowing for multiple transects along which increments were measured (Table 2). Similar selection criteria have been used for freshwater, lake trout Salvelinus namaycush and Selincuo naked carp Gymnocypris selincuoensis (Black et al. 2013b, Tao et al. 2015) and marine teleosts aurora rockfish Sebastes aurora, northern rock sole Lepidopsetta polyxystra, yellowfin sole Limanda aspera and Alaska plaice Pleuronectes quadrituberculatus (Matta et al. 2010, Thompson & Hannah 2010). Clarity of growth increment boundaries is an important feature to be taken into account when considering a species for otolith-based chronology studies. When investigating relationships between otolith (fish) growth and environmental variables, imprecise increment measurements may lead to a result of no correlation or, worse, false correlations with those variables. For this study, the clearest sectioned otoliths from the oldest individuals were retained for increment measurement. The resultant sample sizes were of 29 and 24 for the south and lower west coast, respectively, were consistent with the
  • 46. 45 majority of other otolith increment width chronology studies in which the otoliths from fewer than 50 individuals were used. A post hoc power analysis (Appendix, Power Analysis) corroborated that a sample size of 50 is sufficient to gain adequate statistical power for our B. frenchii chronologies. Although there is no research standard to determine how many samples should be used when constructing otolith-based biochronologies, sample size is likely to vary based on the question being asked and or the geographic area over which samples are collected. For example, Gillanders et al. (2012) used the otoliths from as few as 16 Luderick Girella tricuspidata collected over a restricted region from waters off northern New Zealand, while Morrongiello and Thresher (2015) employed the otoliths of 6143 Tiger Flathead Platycephalus richardsoni collected from six fishing regions spanning eight degrees of latitude. As the samples of B. frenchii used in the current study are from distantly populations located waters on two coasts, the interpretation of the results have been restricted to generalities on the influences of environmental variables on the otolith growth of individuals in those two populations and how conditions on these two coasts drive differences in the response in otolith growth. Cross dating and synchrony among individuals Cross-dating is the process of ensuring that each increment is assigned correctly to the year of formation (Fritts 1976, Cook & Kairiukstis 1990, Maxwell et al. 2011). This is accomplished by visually inspecting the time series and looking for conspicuously wide and/or narrow increments (Black et al. 2005). This process is relatively easily applied in those species that are particularly sensitive to fluctuations in environmental conditions which also is captured by conspicuous
  • 47. 46 variations in growth increment widths in their otoliths (e.g. Matta et al. 2010; Coulson et al 2014; Tao et al. 2015). As in the case of A. gouldii (Rountrey et al. 2014), it may not always be possible in the case of those species, such as of B. frenchii, whose increments have mean widths of 42-46 Āµm. In addition, B. frenchii and A. gouldii are known as complacent, in that they do not show significant response to any environmental change. In the case of such species, the ability to visually align increments on the basis of their widths is not possible. Statistical cross dating provides an avenue for the increment width time series of individuals of such complacent, and non-complacent, species to be statistically checked. The results of cross dating, such as the interseries correlation and mean sensitivity, provide an indication of the quality and level of synchrony within the increment width data. The interseries correlation (ISC), is the mean correlation of the standardised individual series with the mean of all the others. Values of 0.32 and 0.17 for the south and lower west coast populations indicate that individuals of B. frenchii on the south coast respond similarly to fluctuations in environmental conditions resulting in greater synchrony amongst their increment widths. The ISC values for both B. frenchii populations are very low in comparison to those values obtained in other studies, such as 0.76 for grey snapper Lutjanus griseus (Black et al. 2011a) and 0.88 for largemouth bass Micropterus salmoides (Rypel 2009). In the case of the south coast B. frenchii population, the ISC value is markedly higher than 0.11 obtained for A. gouldii in the same waters (Rountrey et al. 2014), but far less than 0.64 and 0.62 obtained for P. laevigatus and L. inops from shallow, inshore waters along the same coast (Coulson et al. 2014). This perhaps indicates that B. frenchii are less complacent towards fluctuating environmental variables than A. gouldii and that, in shallow
  • 48. 47 water environments where variables such as water temperature exhibit greater extremes, the individuals of species living in such waters are likely to all respond in a similar way. This is also demonstrated by very high ISC values (0.49-0.88) for fish in shallow, freshwater environments (Rypel 2009, Black et al. 2013b, Tao et al. 2015). The mean sensitivity is a measure of the range of increment width variation. The very low values of 0.14 and 0.11 for south and lower west coast populations indicate that the increment widths of B. frenchii do not differ markedly between years. While these values are low are at the lower end of the spectrum in comparison to the results from other studies, mean sensitivity values for otolith increment width studies are typically <0.25. Like B. frenchii, the increment widths of hapuku Polyprion oxygeneios, a deep water species found off southern Western Australia, do not vary markedly (mean sensitivity = 0.14) between years of formation (Nguyen et al. 2015). Detrending Detrending is a process, which standardises growth ring widths in chronologies to a mean of one. The aim of detrending is to remove the age-related growth declines in the increment width time series while preserving as much environmentally- induced variability as possible to better illustrate climate driven anomalies in increment widths of (Cook 1985). The type of detrending that is applied to the increment width data will influence the final MIC and potentially any correlations with the selected environmental variables. One of the most common types of detrending used in otolith biochronologies is the cubic smoothing spline (Black et al. 2005, Black et al. 2008b). In this study, three detrending methods were applied
  • 49. 48 and compared to determine the optimal detrending method for B. frenchii. Based on visual inspection of detrended chronologies along with expressed population signals it was decided that double detrending is the best detrending method for B. frenchii. This method was also used for developing the otolith chronology for the Western Blue Groper Achoerodus gouldii, another long-lived labrid species that co-occurs with B. frenchii. Double detrending involves detrending the increment width data, firstly, with a smoothing spline and then, secondly, with a negative exponential function techniques (Rountrey et al. 2014). As separate, stand-alone detrending methods, double detrending and cubic smoothing splines were visually very similar and produced similar expressed population signal values. However, upon closer inspection, the cubic smoothing spline flattened the individual increment series for the individuals off the lower-west coast, particularly in the early years of the chronology. The application of a negative exponential function in conjunction with a smoothing spline greatly improved the chronology as the double detrended chronology followed the trends in the increment widths of the individuals off the south coast. Synchrony between regions Cossington et al. (2010) found regional differences between the B. frenchii populations on the south coast and lower west coast, with those individuals on the south coast growing faster and attaining a larger size and mass at age and maturing earlier in life compared to those individuals on the lower west coast. Despite these biological differences and the that fact that these two populations are spatially separated (~1200km) and occur on coasts that are markedly different in terms of their reef types and the environmental conditions, the response of
  • 50. 49 otolith growth of the individuals in those two populations to regional specific SST was remarkably similar (Figure 9a). This demonstrates, particularly of B. frenchii on the lower west coast close to its northern limit of distribution, that historical increases in water temperature have only positively influenced otolith growth. Modeling of future otolith growth of A. gouldii off the south coast of Western Australia by Rountrey et al. (2014) demonstrated that predicted future increases in water temperate throughout this century are only expected to positively influence otolith growth. While Morrongiello et al. (2015) demonstrated that, throughout the majority of its geographic range on the south east coast of Australia, the growth of the temperate tiger flathead, P. richardsoni, responded positively to temperature, at the northern limit of its distribution, warming waters are having a negative impact on the growth of this species (Morrongiello & Thresher 2015). The results from Morrongiello et al. (2015) suggest that, if future climate change was to continue to result in increasing water temperature, the otolith growth of typically cool-temperate species such as B. frenchii is likely to respond negatively those future increases. Given that B. frenchii is restricted to a narrow latitudinal range, with Rottnest Island being close to their northern limit, and recent marked increases in water temperature off the lower west coast (Pearce & Feng 2013), the population in this region is susceptible to future increases in water temperatures. Financial year vs. calendar year The main growing season for fish species in south-western Australia and, indeed, other regions of the Southern Hemisphere at similar latitudes, extends from approximately mid-spring (October) to early autumn (March). Therefore, the growing period extends from the end of one calendar year to the beginning of
  • 51. 50 another and thus otolith growth occurs over the period that straddles two calendar years. In contrast, the main growing period in the Northern Hemisphere falls in the middle of the calendar year and thus the growth increment is formed over the period of an individual calendar year. Thus, relationships between otolith increment width chronologies constructed for fish species in the Northern Hemisphere and environmental variables in those regions on an annual scale (Matta et al. 2010, Black et al. 2011b, Black et al. 2013b) are correlations on over the same temporal scale. If this same approach is taken for studies carried out in the Southern Hemisphere, increment widths would be correlated with environmental variables that occur after the increment has begun forming. It therefore appears logical that, in the Southern Hemisphere, increment widths should be correlated environmental variables on a financial year (July to June) scale which would encompass the main growing period (October to March). Coulson et al. (2014) used this approach when investigating relationship between the growth of the otoliths two species of Platycephalidae and SST in waters off southern Western Australia. While correlation tests between the MIC for both species and mean annual (financial year) SST were significant, this was not the case when mean annual SST was calculate on a calendar year scale. Other studies in southern Western Australian and New Zealand waters have demonstrated that there is a positive relationship between otolith growth and SST in those months between mid-spring and late autumn of the previous calendar year (Gillanders et al. 2012, Rountrey et al. 2014, Nguyen et al. 2015). However, these researchers did not acknowledge the use of SST on a financial year scale and instead referred to such relationships as lagged correlations with SST in months and or seasons of the previous year. It is clear from the results of the current study
  • 52. 51 and previous research that, when investigating relationships between otolith growth and environmental parameters in waters of the Southern Hemisphere, those parameters should be considered on a financial year scale. Relationships between MIC and environmental variables Unlike other disciplines of sclerochronology, such as those based on the growth bands in corals or the shells of bivalves, the results from studies employing fish otoliths must take into consideration the fact that fish are mobile and potentially inhabit a number of environments throughout their life cycle when interpreting results (Morrongiello et al. 2012, Ong et al. 2015). Site fidelity and home range estimates for the closely related California sheephead Semicossyphus pulcher demonstrate that this species spends ~ 90% of total residence time within a 600 m core area (Topping et al. 2006). Similarly, 10 out the 11 acoustically tagged individuals of the co-occurring labrid, the western blue groper Achoerdus gouldii, largely remained in a 1 km long by 40 m wide strip of coastal reef for the 12 month study period (Bryars et al. 2012). In addition, Fairclough et al. (2011) demonstrated through otolith microchemistry analysis that, on the mid-west coast of Western Australia, the movement of juvenile or adult baldchin groper Choerodon rubescens occurs at relatively small spatial scales. The ability, therefore, to detect responses in the otolith growth of B. frenchii to variations in selected environmental variables is improved because of the likelihood that the individuals, whose otoliths were employed in this study, have been resident in those regions for the majority of their lives. Because teleosts are polkiotherms, their body temperature is thus regulated by the temperature of their surrounding environment, which has been shown to
  • 53. 52 positively influence otolith growth in a number of freshwater and marine species (e.g. Rypel 2009; Matta et al. 2010; Gillanders et al. 2012; Black et al. 2013b), including those species that inhabit south-western Australian waters (Coulson et al. 2014; Nguyen et al. 2015). The results of the GAMMs demonstrated that, while there was a significant, positive relationship between the MIC for the south coast population with mean annual SST, the relationship with this variable was more significant on the lower west coast (south P = 0.003, lower west P = 0.0006, Table 7). However, there was a higher estimated effect otolith increment growth in increasing annual SST for the south coast population of B. frenchii (estimate for increase of 1Ā°C (Ā±SE) = 1.25 Āµm Ā± 0.42) compared to the lower west population (estimate for increase of 1Ā°C (Ā±SE) = 1.18 Āµm Ā± 0.35). Cossington et al. (2010) stated that the cool south coast is more favourable for this species. As expected, there is a higher estimated effect for otolith growth in increasing in this area. The extent to which SST fluctuates from year to year is much greater on along the lower west coast than the south coast (Figure 1) and, since ~ 1970, there has been an increasing trend in the mean annual SST on that former coast. The relationship between the MIC for B. frenchii on the south coast and SST for is similar to the findings of Rountrey et al. (2014) for the co-occurring A. gouldii, who found a weak, but significant influence of regional SST on the otolith growth of A. gouldii in the south coast of Western Australia. Similarly, Coulson et al. (2014) demonstrated that increasing water temperature has resulted in increased otolith growth of two platycephalid species in inshore waters on the same coast.
  • 54. 53 The effects of the Leeuwin Current in waters off the lower west coast of Western Australia are well documented and Fremantle sea level (FSL) is commonly used as a proxy for the Leeuwin Current through El Nino Southern Oscillation cycles (Pearce & Phillips 1988, Caputi et al. 1996, Caputi et al. 2001, Feng et al. 2008). Although there were no significant Pearson's correlations between the MIC for the lower west coast population of B. frenchii and FSL, the GAMM showed a highly significant relationship (P = 0.008) with an estimated 4.62 Āµm (SE = Ā± 2.05) increase in otolith increment size for every one cm increase in Fremantle sea level. Nguyen et al. (2015) found that the Leeuwin Current had a significant influence on the growth of hapuku Polyprion oxygeneios in south-western Western Australia through its effect on primary productivity eventually leading to increase prey resources for P. oxygeneios. The strength of the Leeuwin Current weakens, as does its influence on the marine environment, as it flows east along the south coast. It is thus not surprising that Coulson et al. (2014) found no relationship between the otolith growth of two platycepahlids and sea level in inshore waters along the south coast of Western Australia. The influence of the Leeuwin Current on the otolith growth of B. frenchii is likely a reflection of a combination of oceanographical and biological influences by the strengthening of this current in this region. Conclusion In conclusion, two otolith increment-width chronologies were constructed for Bodianus frenchii from the south and lower-west coasts of Western Australia. While the otoliths for B. frenchii possess very clear growth increments, the lack of conspicuous variability in widths required to undertake cross dating made this
  • 55. 54 process unsuccessful. The lack of synchrony in the trends displayed by the increment widths between individuals demonstrates the insensitivity, as reflected by the interseries correlation values, of B. frenchii to environmental fluctuations. Despite the weak signal in the otolith increment widths, significant relationships with regional SST and Leeuwin Current strength were present. Generalized additive mixed modeling demonstrated that SST was an important driver of otolith growth of B. frenchii on the south coast, while on the lower west coast, the effects of a strong Leeuwin Current influence the otolith growth of B. frenchii in these waters.
  • 56. 55 References AbrĆ moff MD, MagalhĆ£es PJ, Ram SJ (2004) Image processing with ImageJ. Biophotonics International 11:36-42 Black BA, Boehlert GW, Yoklavich MM (2005) Using tree-ring crossdating techniques to validate annual growth increments in long-lived fishes. Canadian Journal of Fisheries and Aquatic Sciences 62:2277-2284 Black BA, Boehlert GW, Yoklavich MM (2008a) Establishing climate-growth relationships for yelloweye rockfish (Sebastes ruberrimus) in the northeast Pacific using a dendrochronological approach. Fisheries Oceanography 17:368-379 Black BA, Gillespie DC, MacLellan SE, Hand CM (2008b) Establishing highly accurate production-age data using the tree-ring technique of crossdating: a case study for Pacific geoduck (Panopea abrupta). Canadian Journal of Fisheries and Aquatic Sciences 65:2572-2578 Black BA (2009) Climate-driven synchrony across tree, bivalve, and rockfish growth-increment chronologies of the northeast Pacific. Marine Ecology Progress Series 378:37-46 Black BA, Schroeder ID, Sydeman WJ, Bograd SJ, Lawson PW (2010) Wintertime ocean conditions synchronize rockfish growth and seabird reproduction in the central California Current ecosystem. Canadian Journal of Fisheries and Aquatic Sciences 67:1149-1158 Black BA, Allman RJ, Schroeder ID, Schirripa MJ (2011a) Multidecadal otolith growth histories for red and gray snapper (Lutjanus spp.) in the northern Gulf of Mexico, USA. Fisheries Oceanography 20:347-356
  • 57. 56 Black BA, Schroeder ID, Sydeman WJ, Bograd SJ, Wells BK, Schwing FB (2011b) Winter and summer upwelling modes and their biological importance in the California Current Ecosystem. Global Change Biology 17:2536-2545 Black BA, Matta ME, Helser TE, Wilderbuer TK (2013a) Otolith biochronologies as multidecadal indicators of body size anomalies in yellowfin sole (Limanda aspera). Fisheries Oceanography 22:523-532 Black BA, Biela VR, Zimmerman CE, Brown RJ (2013b) Lake trout otolith chronologies as multidecadal indicators of high-latitude freshwater ecosystems. Polar Biology 36:147-153 Boehlert GW (1985) Using objective criteria and multiple regression models for age determination in fishes. Fishery Bulletin 83:103-117 Briffa KR, Jones PD, Schweingruber FH, KarlĆ©n W, Shiyatov SG (1996) Tree- ring variables as proxy-climate indicators: problems with low-frequency signals. Climatic variations and forcing mechanisms of the last 2000 years. Springer, Berlin Heidelberg Bryars S, Rogers P, Huveneers C, Payne N, Smith I, McDonald B (2012) Small home range in southern Australia's largest resident reef fish, the western blue groper (Achoerodus gouldii): implications for adequacy of no-take marine protected areas. Marine and Freshwater Research 63:552-563 Bunn AG (2008) A dendrochronology program library in R (dplR). Dendrochronologia 26:115-124 Bunn AG (2010) Statistical and visual crossdating in R using the dplR library. Dendrochronologia 28:251-258
  • 58. 57 Bunn AG, Korpela M, Biondi F, Campelo F, MĆ©rian P, Qeadan F, Zang C (2015) Dendrochronology Program Library in R. Campana SE, Neilson JD (1985) Microstructure of fish otoliths. Canadian Journal of Fisheries and Aquatic Sciences 42:1014-1032 Campana SE (2001) Otoliths, increments, and elements: keys to a comprehensive understanding of fish populations? Canadian Journal of Fisheries and Aquatic Sciences 58:30-38 Campana SE, Thorrold SR (2001) Otoliths, increments, and elements: keys to a comprehensive understanding of fish populations? Canadian Journal of Fisheries and Aquatic Sciences 58:30-38 Caputi N, Fletcher W, Pearce A, Chubb C (1996) Effect of the Leeuwin Current on the recruitment of fish and invertebrates along the Western Australian coast. Marine and Freshwater Research 47:147-155 Caputi N, Chubb C, Pearce A (2001) Environmental effects on recruitment of the western rock lobster, Panulirus cygnus. Marine and Freshwater Research 52:1167-1174 Carter R, Woodroffe C (1994) Coastal evolution: an introduction. Cambridge University Press Cleveland WS (1993) Visualizing data. Hobart Press Cook ER, Peters K (1981) The smoothing spline: a new approach to standardizing forest interior tree-ring width series for dendroclimatic studies. Tree Ring Bulletin 41:45-53 Cook ER (1985) A time series analysis approach to tree ring standardisation (Dendrochronloogy, Forestry, Dendroclimatology, Autoregressive Process). Phd Thesis, Univeristy of Arizona,
  • 59. 58 Cook ER, Kairiukstis LA (1990) Methods of dendrochronology: applications in the environmental sciences. Springer Science & Business Media Cossington S, Hesp SA, Hall NG, Potter IC (2010) Growth and reproductive biology of the foxfish Bodianus frenchii, a very long-lived and monandric protogynous hermaphroditic labrid. Journal of Fish Biology 77:600-626 Coulson PG, Black BA, Potter IC, Hall NG (2014) Sclerochronological studies reveal that patterns of otolith growth of adults of two co-occurring species of Platycephalidae are synchronised by water temperature variations. Marine Biology 161:383-393 Davis-Foust SL (2012) Long-term changes in population statistics of Freshwater Drum (Aplodinotus grunniens) in Lake Winnebago, Wisconsin, using otolith growth chronologies and bomb radiocarbon age validation. Doctor of Philosophy, University of Wisconsin, Douglass AE (1920) Evidence of climatic effects in the annual rings of trees. Ecology 1:24-32 Douglass AE (1941) Crossdating in dendrochronology. Journal of Forestry 39:825-831 Fairclough DV, Edmonds JS, Lenanton RC, Jackson G, Keay IS, Crisafulli BM, Newman SJ (2011) Rapid and cost-effective assessment of connectivity among assemblages of Choerodon rubescens (Labridae), using laser ablation ICP-MS of sagittal otoliths. Journal of Experimental Marine Biology and Ecology 403:46-53 Feng M, Biastoch A, Bƶning C, Caputi N, Meyers G (2008) Seasonal and interannual variations of upper ocean heat balance off the west coast of Australia. Journal of Geophysical Research: Oceans (1978ā€“2012) 113
  • 60. 59 Feng M, McPhaden MJ, Xie SP, Hafner J (2013) La Nina forces unprecedented Leeuwin Current warming in 2011. Scientific reports 3:1277 Fritts H (1976) Tree rings and climate. Elsevier Gillanders BM, Black BA, Meekan MG, Morrison MA (2012) Climatic effects on the growth of a temperate reef fish from the Southern Hemisphere: a biochronological approach. Marine Biology 159:1327-1333 Grissino-Mayer HD (2001) Evaluating crossdating accuracy: a manual and tutorial for the computer program COFECHA. Tree-Ring Research 57:205-221 Holmes RL (1983) Computer-assisted quality control in tree-ring dating and measurement. Tree-ring bulletin 43:69-78 Howard RK (1989) The structure of a nearshore fish community of Western Australia: diel patterns and the habitat role of limestone reefs. Environmental Biology of Fishes 24:93-104 Kastelle CR, Helser TE, Black BA, Stuckey MJ, C. Gillespie D, McArthur J, Little D, D. Charles K, Khan RS (2011) Bomb-produced radiocarbon validation of growth-increment crossdating allows marine paleoclimate reconstruction. Palaeogeography, Palaeoclimatology, Palaeoecology 311:126-135 Kendrick G (1999) Western Australia. In: Andrew N (ed) Under southern seas. University of New South Wales Press, Sydney Lehodey P, Grandperrin R (1996) Influence of temperature and ENSO events on the growth of the deep demersal fish alfonsino, Beryx splendens, off New Caledonia in the western tropical South Pacific Ocean. Deep Sea Research Part I: Oceanographic Research Papers 43:49-57
  • 61. 60 List JH, Terwindt JH (1995) Large-scale coastal behavior, Vol 126. DIANE Publishing Matta EM, Black BA, Wilderbuer TK (2010) Climate-driven synchrony in otolith growth-increment chronologies for three Bering Sea flatfish species. Marine Ecology Progress Series 413:137-145 Maxwell RS, Wixom JA, Hessl AE (2011) A comparison of two techniques for measuring and crossdating tree rings. Dendrochronologia 29:237-243 Meekan M, Wellington G, Axe L (1999) El Nino-Southern Oscillation events produce checks in the otoliths of coral reef fishes in the GalĆ”pagos Archipelago. Bulletin of Marine Science 64:383-390 Melvin TM, Briffa KR (2008) A ā€œsignal-freeā€ approach to dendroclimatic standardisation. Dendrochronologia 26:71-86 Miller JA, Wells BK, Sogard SM, Grimes CB, Cailliet GM (2010) Introduction to proceedings of the 4th International Otolith Symposium. Environmental Biology of Fishes 89:203-207 Morrissey MB (2011) Exploiting natural history variation: looking to fishes for quantitative genetic models of natural populations. Ecology of Freshwater Fish 20:328-345 Morrongiello JR, Crook DA, King AJ, Ramsey DSL, Brown P (2011) Impacts of drought and predicted effects of climate change on fish growth in temperate Australian lakes. Global Change Biology 17:745-755 Morrongiello JR, Thresher RE, Smith DC (2012) Aquatic biochronologies and climate change. Nature Climate Change 2:849-857
  • 62. 61 Morrongiello JR, Thresher RE (2015) A statistical framework to explore ontogenetic growth variation among individuals and populations: a marine fish example. Ecological Monographs 85:93-115 Nguyen HM, Rountrey AN, Meeuwig JJ, Coulson PG, Feng M, Newman SJ, Waite AM, Wakefield CB, Meekan MG (2015) Growth of a deep-water, predatory fish is influenced by the productivity of a boundary current system. Scientific Reports 5 Ong JJ, Nicholas Rountrey A, Jane Meeuwig J, John Newman S, Zinke J, Gregory Meekan M (2015) Contrasting environmental drivers of adult and juvenile growth in a marine fish: implications for the effects of climate change. Scientific Reports 5:10859 Pannella G (1971) Fish otoliths: daily growth layers and periodical patterns. Science 173:1124-1127 Pearce A, Phillips B (1988) ENSO events, the Leeuwin Current, and larval recruitment of the western rock lobster. Journal du Conseil: ICES Journal of Marine Science 45:13-21 Pearce A, Feng M (2007) Observations of warming on the Western Australian continental shelf. Marine and Freshwater Research 58:914-920 Pearce A, Feng M (2011) The" marine heat wave" off Western Australia during the summer of 2010/11. Western Australian Fisheries and Marine Research Laboratories Pearce AF, Feng M (2013) The rise and fall of the ā€œmarine heat waveā€ off Western Australia during the summer of 2010/2011. Journal of Marine Systems 111-112:139-156
  • 63. 62 Pereira DL, Bingham C, Spangler GR, Conner DJ, Cunningham PK, Secor D, Dean J, Campana S (1995) Construction of a 110-year biochronology from sagittae of freshwater drum (Aplodinotus grunniens). Recent developments in fish otolith research University of South Carolina Press, Columbia:177- 196 Popper AN, Ramcharitar J, Campana SE (2005) Why otoliths? Insights from inner ear physiology and fisheries biology. Marine and Freshwater Research 56:497-504 Rayner NA (2003) Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. Journal of Geophysical Research 108 Rose TH, Smale DA, Botting G (2012) The 2011 marine heat wave in Cockburn Sound, southwest Australia. Ocean Science 8:545-550 Rountrey AN (2009) Life Histories of Juvenile Woolly Mammoths from Siberia: Stable Isotope and Elemental Analyses of Tooth Dentin. The University of Michigan, Rountrey AN, Coulson PG, Meeuwig JJ, Meekan M (2014) Water temperature and fish growth: otoliths predict growth patterns of a marine fish in a changing climate. Global Change Biology 20:2450-2458 Rypel AL (2009) Climate growth relationships for largemouth bass (Micropterus salmoides) across three southeastern USA states. Ecology of Freshwater Fish 18:620-628 Sanderson P, Eliot I, Hegge B, Maxwell S (2000) Regional variation of coastal morphology in southwestern Australia: a synthesis. Geomorphology 34:73-88
  • 64. 63 Tao J, Chen Y, He D, Ding C (2015) Relationships between climate and growth of Gymnocypris selincuoensis in the Tibetan Plateau. Ecology and Evolution 5:1693-1701 Thompson JE, Hannah RW (2010) Using cross-dating techniques to validate ages of aurora rockfish (Sebastes aurora): estimates of age, growth and female maturity. Environmental Biology of Fishes 88:377-388 Topping DT, Lowe CG, Caselle JE (2006) Site fidelity and seasonal movement patterns of adult California sheephead Semicossyphus pulcher (Labridae): an acoustic monitoring study. Marine Ecology Progress Series 326:257- 267 Tukey JW Exploratory data analysis: as part of a larger whole. Proc Proceedings of the 18th conference on design of experiments in Army research and development I Washington, DC. DTIC Document Tukey JW (1977) Exploratory data analysis. Reading, Ma 231:32 Wells FE, Keesing JK (1990) Population characteristics of the abalone Haliotis roei on intertidal platforms in the Perth metropolitan area. Journal of the Malacological Society of Australia 11:65-71 Wells FE, Walker D, Wells FE, Walker D (1993) Introduction to the marine environment of Rottnest Island, Western Australia. In: Lethbridge R, Lethbridge R (eds) The Marine Flora and Fauna of Rottnest Island, Western Australia, Book 1. Western Australian Museum, Perth Wigley TM, Briffa KR, Jones PD (1984) On the average value of correlated time series, with applications in dendroclimatology and hydrometeorology. Journal of Climate and Applied Meteorology 23:201-213 Wood S, Scheipl F, Wood MS (2014) Package ā€˜gamm4ā€™.
  • 65. 64 Zuur A, Ieno EN, Walker N, Saveliev AA, Smith GM (2009) Mixed effects models and extensions in ecology with R. Springer Zuur AF, Ieno EN, Elphick CS (2010) A protocol for data exploration to avoid common statistical problems. Methods in Ecology and Evolution 1:3-14
  • 66. 65 Appendix List of figures A1. Dot plots for individual chronology time series for the 29 Bodianus frenchii from the south coast. Each plot represents the individual points for each individual with their identifying fish number at the top of each plot..................................... 67 A2. Dot plots for individual chronology time series for 24 Bodianus frenchii from the lower west coast. Each plot represents the individual points for each individual with their identifying fish number at the top of each plot..................................... 68 A3. Plots to visually assess homogeneity. Residuals vs. fitted values for the MICs of a) south coast and c) lower west coast and box plots of individual Bodianus frenchii increment width time series for the b) south coast and d) west coast...... 69 A4. QQ plots and residual histograms used to assess normality in the increment width data for the B. frenchii from a, b) the south and c, d) lower west coasts.... 70 A5. Plot displaying the autocorrelation function (ACF) calculated for the increment width data for B. frenchii on the a) south and b) lower west coasts at lags of 0 to 30 years. The dotted line indicates the statistical significance at alpha = 0.05. ................................................................................................................... 72
  • 67. 66 Data Exploration Outliers An "outlier" is an observation that has a relatively large or small value (conventionally more than three standard deviations away from the mean) compared to the majority of the data (Zuur et al. 2010). Typically, a box plot or Cleveland dot plot (Cleveland 1993) is used to visually assess whether outliers are present. If outliers are highlighted, the corresponding data need to be investigated. To assess the outliers in the chronology dataset, the data was visually, inspected for outliers. There does not appear to be any outliers in either the south (A1) or the lower west coast samples (A2). A1. Dot plots for individual chronology time series for the 29 Bodianus frenchii from the south coast. Each plot represents the individual points for each individual with their identifying fish number at the top of each plot.
  • 68. 67 A2. Dot plots for individual chronology time series for 24 Bodianus frenchii from the lower west coast. Each plot represents the individual points for each individual with their identifying fish number at the top of each plot.
  • 69. 68 Homogeneity A basic assumptions for parametric statistical analyses, such as linear regression, is that the data have equal variance or homogeneity. This can be assessed using a residuals vs. fitted values plot (A3 a, c), or individual boxplots (A3 b, d) to assess the spread of residuals. The spread of variances in the B. frenchii chronology data look as though there might be a slight conical shape (A3 a, c), caused most likely by a decrease in otolith increment width with increasing age. This suggests that there is a variance structure present within the data. A3. Plots to visually assess homogeneity. Residuals vs. fitted values for the MICs of a) south coast and c) lower west coast and box plots of individual Bodianus frenchii increment width time series for the b) south coast and d) west coast.
  • 70. 69 Normality Normality, another assumption for parametric statistics, is assessed using QQ plots (A4 a, c) or inspecting the spread of the residuals, to see if they are normally distributed (A4 b, d). The normality visualization plots for the chronology data show that the data are not normally distributed, as the data points in QQ plots in both the south (A4 a) and lower west coast (A4 c) are up to 10 standard deviations away from the mean and the residual histograms are highly skewed to the right (A4 b, d). Therefore, the assumption of normality for any parametric modeling using the B. frenchii increment width data is violated. A4. QQ plots and residual histograms used to assess normality in the increment width data for the B. frenchii from a, b) the south and c, d) lower west coasts.