characterisation of the light environment and biophysical parameters of seagrass using remote

155
Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote Sensing Novi Susetyo Adi BSc. (Hons), MSc. A thesis submitted for the degree of Doctor of Philosophy at The University of Queensland in 2015 School of Geography, Planning and Environmental Management

Upload: others

Post on 11-Sep-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

Characterisation of the Light Environment and Biophysical Parameters

of Seagrass Using Remote Sensing Novi Susetyo Adi

BSc. (Hons), MSc.

A thesis submitted for the degree of Doctor of Philosophy at

The University of Queensland in 2015

School of Geography, Planning and Environmental Management

Page 2: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

ii

Seagrass is ecologically important submerged aquatic vegetation that serves as one of the major

sources of primary production in shallow waters. Despite their high productivity, seagrass habitats

are under threat from anthropogenic and global climate change influences and therefore

understanding their dynamics and environmental controls is essential. It is fortunate that seagrasses

have distinct optical signatures observable from space by satellite sensors, allowing mapping and

monitoring of seagrass habitats in spatially continuous and multi-temporal modes. Light

availability, along with seagrass leaf area index (LAI), biomass and productivity are important

parameters for characterising the condition of seagrass habitat. The aim of this research was to

investigate these parameters by integrating field measurement, laboratory analysis and remote

sensing to estimate seagrass light climate, LAI, biomass and gross primary productivity. The

research addressed the following three objectives with investigations based in a section of

subtropical seagrass in Moreton Bay, Australia: (a) to investigate light quality and quantity in the

seagrass environment using in situ optical measurements and remote sensing, (b) to map seagrass

LAI and biomass using WorldView-2 satellite data, and (c) to estimate seagrass gross primary

productivity using a combination of models and remote sensing data.

Abstract

To address research objective one, light quality was assessed using daily water optical

measurements at two sites in Moreton Bay. Light quantity was investigated by estimating the

seagrass surface area from satellite image-based photosynthetically active radiation (PAR),

photosynthetically utilised radiation (PUR) and percentage of light relative to surface light (% SI)

parameters. The result showed green light dominated the light climate at the seagrass meadows of

the measurement site. A blue light limitation of seagrass in Wanga Wallen bank was indicated, as

there was a rapid decrease in blue light contributions relative to green and red light, moving from

the more dense inshore seagrass site to the less-dense offshore seagrass site. The majority of the

seagrass surface area in the areas assessed was successfully mapped based on satellite image-based

light quantity parameters of PAR, PUR and % SI, providing insights into the interaction between

light parameters and the spatial distribution of seagrass.

The second objective developed a method to estimate seagrass LAI and biomass from image data

by first examining the relationship between seagrass LAI and biomass, and reflectance using in situ

data. Regression models were then developed and the most accurate one was applied to two high-

spatial resolution multi-spectral WorldView-2 image data to estimate seagrass LAI and biomass

from satellite image reflectance. Analysis using in situ data revealed strong correlation between the

green band and LAI but weak correlation between reflectance and biomass. Significant

Page 3: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

iii

relationships found between LAI and biomass confirmed the two parameters had an allometric

relationship. Based on these results, winter and summer LAI maps were estimated from the

WorldView-2 green bands and the corresponding biomass maps were later generated from the LAI

maps. The accuracy of the image-based LAI and biomass maps was found acceptable (LAI = 62%

for the June 2012 map and 73% for the February 2013 map; Biomass = 71% for the June 2012 map

and 60% for the February 2013 map).

The third research objective estimated seagrass productivity using image-based parameters

produced in the previous chapters. The estimation utilised the biomass-based method and a method

that used the environmental and physiological parameters of seagrass (termed as a process-based

method). The two methods were successfully applied to produce productivity maps with values that

fell within the range of seagrass productivity literature values. Seasonal analysis using the biomass-

based method applied to the winter and summer satellite images revealed no significant differences

in the average productivities, probably due to the contrasting growth patterns between the dominant

seagrass species Zostera muelleri and the other species that cancelled out any seasonal pattern.

This research presented a set of methods combining in situ measurement, laboratory analysis and

remote sensing approaches to map seagrass light climate, LAI, biomass and primary productivity

over spatial scales < 150km2. In addition to seagrass parameters commonly estimated from remote

sensing such as the extent, cover and species present, the mapping approaches and products of this

thesis demonstrated that the light climate (in terms of PAR, PUR and % SI) and biophysical

parameters (LAI, biomass and productivity) of seagrass could be derived or estimated using remote

sensing. In terms of the regional context, this thesis provided new insights into the condition of

seagrass in the Eastern Banks, Moreton Bay, by conducting the first investigation into the spatial

and temporal dynamics of light climate, LAI and productivity in the area using a remote sensing

approach. Refining the methods presented in this thesis and combining them with hydrographic

modelling and information would be beneficial for a more comprehensive understanding of the

temporal and spatial dynamics of seagrass ecosystems.

Page 4: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

iv

This thesis is composed of my original work and contains no material previously published or

written by another person except where due reference has been made in the text. I have clearly

stated the contribution by others to jointly-authored works that I have included in my thesis.

Declaration by author

I have clearly stated the contribution of others to my thesis as a whole, including statistical

assistance, survey design, data analysis, significant technical procedures, professional editorial

advice and any other original research work used or reported in my thesis. The content of my thesis

is the result of work I have carried out since the commencement of my research higher degree

candidature and does not include a substantial part of work that has been submitted to qualify for

the award of any degree or diploma in any university or other tertiary institution. I have clearly

stated which parts of my thesis, if any, have been submitted to qualify for another award.

I acknowledge that an electronic copy of my thesis must be lodged with the University Library and,

subject to the General Award Rules of The University of Queensland, immediately made available

for research and study in accordance with Copyright Act 1968.

I acknowledge that copyright of all material contained in my thesis resides with the copyright

holder(s) of that material. Where appropriate I have obtained copyright permission from the

copyright holder to reproduce material in this thesis.

Page 5: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

v

Peer-reviewed papers

Publications during candidatures

1. Characterising Light Climate in Seagrass Habitat Using Field Measurement and Remote

Sensing. (Status: In preparation for Submission to Estuarine, Coastal and Shelf Science)

2. Estimating Seagrass Leaf Area Index and Biomass Using WorldView-2 Data. (Status: In

Preparation for submission to Remote Sensing)

3. Application of Remote Sensing Data for Estimating Seagrass Productivity. (Status: In

Preparation for submission to Remote Sensing in Ecology and Conservation)

Conference Presentation

1. Adi, N.S., Phinn, S.R., Roelfsema, C.M. 2013. Estimating The Diffuse Attenuation Coefficient

From Moderate Spatial Resolution, Multi-Spectral Satellite Data in Seagrass Environment.

International Geoscience and Remote Sensing Symposium (IGARSS). 21–26 July 2013,

Melbourne, Australia.

2. Adi, N.S., Phinn, S.R., Roelfsema, C.M., Samper-Villareal, J. 2013. Integrating Field and

Remote Sensing Approaches for Mapping Seagrass Leaf Area Index. Asian Conference On

Remote Sensing (ACRS), 20-24 October 2013, Bali, Indonesia.

No publications included

Publications included in this thesis

Page 6: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

vi

Advisers:

Contributions by others to the thesis

Prof Stuart Phinn

(School of Geography, Planning and Environmental Management, The University of Queensland)

Dr Chris Roelfsema

(School of Geography, Planning and Environmental Management, The University of Queensland)

Prof Catherine Lovelock

(School of Biological Sciences, The University of Queensland)

A/Prof Sophie Dove

(School of Biological Sciences, The University of Queensland)

The advisers Prof Stuart Phinn, Dr Chris Roelfsema, Prof Catherine Lovelock and A/Prof Sophie

Dove contributed to the research direction and research design of this thesis and provided comments

and corrections.

None

Statement of parts of the thesis submitted to qualify for the award of another degree

Page 7: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

vii

My thesis, presented here, is the culmination of the work of two-and-a-half years at the

School of Geography, Planning and Environmental Management (GPEM), at the University of

Queensland, which would not have been possible without the unwavering academic and personal

support of many people whose contribution I would like to acknowledge here.

Acknowledgements

First, I would like to sincerely thank my advisers at GPEM, Professor Stuart Phinn and Dr

Chris Roelfsema, who helped me overcome a difficult situation in my life in 2012 and kindly

welcomed me to their lab and research group, leading to the completion of this thesis. Thank you

very much Stuart and Chris for your excellent professional knowledge, advice, guidance and

patience during my study time at GPEM. I give thanks to my advisers from the School of Biological

Sciences, UQ, Associate Professor Sophie Dove for hosting me in her lab for two years and for

opening my mind to the world of climate change studies; and Professor Catherine Lovelock for

allowing me to work in her lab for seagrass sample analysis. I extend my thanks also to Professor

Ove Hoegh-Guldberg for reading the first version of my PhD proposal and for allowing me to study

in his lab when I first arrived in Australia. I also would like to express my gratitude to AusAID for

financially supporting my study in Australia and to GPEM for additional financial support for some

academic events, such as conference and field works.

There are some people I would like to thank for their friendship, help and support during my

two years at the Coral Reef Ecosystems Lab: James Fang, Catalina Reyes Nivia, Dorothea Bender,

Robert Mason, Aaron Chai, Giovanni Bernal Carrillo and Annamieke Van Den Heuvel. I know

some of you have graduated or may have left UQ or Australia. I look forward to our paths crossing

again someday.

I would like to acknowledge the contribution of the people who helped me during my field

work: Mitchell Lyons, Eva Kovacs, Megan Saunders, Chris Roelfsema and Jimena Samper-

Villarreal. Special thanks goes to Jimena, Megan and Eva for “saving my life” during the field work

of February 2013. Jimena's help and support were also influential during my long days in Cath's lab

analysing seagrass samples. I am grateful to the staff of the Moreton Bay Research Station, UQ,

particularly Kevin Townsend, Martin Wayne and Kathryn Crouch for helping and hosting me

during my field and lab works in North Stradbroke Island. I had the privilege of interacting with

bright and warm personalities from the Biophysical Remote Sensing Group, GPEM, whose shared

ideas, help and support enriched my PhD experience and I would like to thank them: Muhammad

Page 8: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

viii

Kamal, Rodney Borrego Avecedo, Robert and Tess Canto, Sabrina Wu, Javier Platino, Eva Kovacs,

Ralph Trancoso, Jiban Deb, Kasper Johansson, Martina Reichstetter, Michael Hewson, Jo Edkins

and the rest . . .

The Indonesian community in Brisbane was essential for bringing Indonesian warmth,

cheerfulness and kindness when I needed them, particularly during my hard times. Thanks to Bu

Yanti, Ashadi “mas bro”, Mas Iwan, Mas Dendi, Pak Bambang, Bu Tris, Pak Kamal, Pak Tunjung,

Mas Agus, Pak Basuki and Bu Ndari, Pak Yusuf, Kang Beben, Mas Heri, Pak Candra, Mas Rino,

Mbak Rini, Mbak Agnes, Mas Farid, Mbak Frida, Pak Radja and all the members of the Indonesian

badminton club, Smash Bersalin. A very special thanks goes to my mum and sister, Mirna, who

always believed in me and prayed for me. I can't count how many times you all said "don't be

discouraged, keep trying and praying to God", your encouragement meant a lot to me.

Last in this list, but first in my heart, I express my appreciation to my beloved wife, Lia

Kusumawati, for her unconditional support, encouragement and unwavering confidence in my

decision to pursue PhD study and in my ability to finish it. Nothing would make me happier now

than coming back to our home and painting more colours in our lives.

Page 9: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

ix

seagrass leaf area index, seagrass biomass, diffused attenuation coefficient, photosynthetically

active radiation, photosynthetically utilised radiation, seagrass gross productivity, eastern banks,

moreton bay, reflectance, remote sensing.

Keywords

ANZSRC code: 090905, Photogrammetry and Remote Sensing, 80%

Australian and New Zealand Standard Research Classifications (ANZSRC)

ANZSRC code: 050206, Environmental Monitoring, 10%

ANZSRC code: 060205, Marine and Estuarine Ecology, 10%

FoR code: 0909

Fields of Research (FoR) Classification

Geomatic Engineering, 80%

FoR code: 0502

Environmental Science and Management, 10%

FoR code: 0602

Ecology: 10%

Page 10: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

x

Table of Contents

Abstract ............................................................................................................................................... ii

Declaration by author ....................................................................................................................... iv

Publications during candidatures ..................................................................................................... v

Publications included in this thesis ................................................................................................... v

Contributions by others to the thesis............................................................................................... vi

Statement of parts of the thesis submitted to qualify for the award of another degree ............. vi

Acknowledgements........................................................................................................................... vii

Keywords ........................................................................................................................................... ix

Australian and New Zealand Standard Research Classifications (ANZSRC) ............................ ix

Fields of Research (FoR) Classification .......................................................................................... ix

Table of Contents ............................................................................................................................... x

List of Figures .................................................................................................................................. xiii

List of Tables ................................................................................................................................. xviii

List of Abbreviations used in the thesis ......................................................................................... xx

Chapter 1: Introduction and Significance of the Research ............................................................ 1

1.1. INTRODUCTION AND OVERVIEW OF THE RESEARCH CONTEXT ....................................... 1

1.2. BACKGROUND ................................................................................................................ 3

1.2.1. Mapping light in seagrass environments: Extending the principles of ocean optics

to the remote sensing approach. ....................................................................................... 3

1.2.2. Remote sensing applications for mapping seagrass biophysical properties. .......... 5

1.2.3. Estimating productivity using remote sensing data: A review of applications to

terrestrial, ocean surface and coastal environment. ......................................................... 7

1.3. PROBLEM STATEMENT .................................................................................................. 11

1.4. AIM .............................................................................................................................. 12

1.5. OBJECTIVES .................................................................................................................. 12

1.6. THESIS OUTLINE ........................................................................................................... 12

Chapter 2: Research Approach ...................................................................................................... 14

Page 11: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xi

2.1. GENERAL OVERVIEW OF RESEARCH METHODS ............................................................ 14

2.2. STUDY SITE .................................................................................................................. 16

2.3. IMAGE DATA ................................................................................................................ 18

2.4. FIELD DATA COLLECTION ............................................................................................ 19

Chapter 3: Characterising the Underwater Light Environment in Seagrass Habitat Using

Field Survey and Remote Sensing Approaches ............................................................................. 20

MAIN FINDINGS ................................................................................................................... 20

3.1. INTRODUCTION ............................................................................................................. 20

3.2. DATA AND METHODOLOGY .......................................................................................... 23

3.2.1. Field data collection and processing of Kd (λ) and PAR (λ). .............................. 23

3.2.2. Satellite image acquisition, image processing and seagrass surface area

calculation. ..................................................................................................................... 26

3.2.3. Comparison with in situ data and empirical algorithm. ....................................... 30

3.2.3.1. Estimating KdPAR from Kd490 –KdPAR empirical model. ............................... 30

3.2.3.2. Estimating KdPAR from local Kd490–KdPAR algorithm. .................................. 31

3.3. RESULTS AND DISCUSSION ........................................................................................... 31

3.3.1. In situ measurement of Kd (λ), PAR and PAR blue, green, red. ......................... 31

3.3.2. Potential seagrass surface area based on PAR, % SI and PUR maps. ................. 39

3.3.3. Comparison with in situ data. ............................................................................... 46

3.3.4. Comparison with Kd490–KdPAR empirical model. ................................................ 46

3.3.5. Comparison with local Kd490–KdPAR algorithm. .................................................. 47

3.4. CONCLUSIONS, PERSPECTIVES AND FUTURE WORK ..................................................... 50

Chapter 4: Estimating Seagrass Leaf Area Index and Biomass Using WorldView-2

Satellite Data ..................................................................................................................................... 53

MAIN FINDINGS ................................................................................................................... 53

4.1. INTRODUCTION ............................................................................................................. 53

4.2. DATA AND METHODOLOGY .......................................................................................... 55

4.2.1. Field data collection and processing. ................................................................... 55

4.2.1.1. Seagrass samples collection and analysis. ..................................................... 55

4.2.1.2. In situ reflectance data. .................................................................................. 57

4.2.2. Band selection using in situ reflectance data. ...................................................... 57

4.2.3. Image processing and model development. ......................................................... 58

4.2.4. Reliability statistics for seagrass LAI and biomass estimation. ........................... 60

Page 12: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xii

4.3. RESULTS AND DISCUSSION ........................................................................................... 61

4.3.1. Seagrass biomass and LAI from field sampling .................................................. 61

4.3.2. Correlation between LAI, biomass and in situ reflectance data. .......................... 62

4.3.3. Correlation between LAI and biomass. ................................................................ 67

4.3.4. Models for estimating biomass and LAI from image data. .................................. 68

4.3.4.1. Model for LAI. ................................................................................................ 68

4.3.4.2. Model for biomass. ......................................................................................... 71

4.3.5. Mapping seagrass LAI and biomass using WorldView-2 image data. ................ 73

4.3.6. Accuracy assessment of the satellite-based LAI and biomass. ............................ 78

4.4. CONCLUSIONS, PERSPECTIVES AND FUTURE WORK ..................................................... 82

Chapter 5: Application of Remote Sensing Data for Estimating Seagrass Productivity .......... 86

MAIN FINDINGS ................................................................................................................... 86

5.1. INTRODUCTION ............................................................................................................. 86

5.2. DATA AND METHODOLOGY .......................................................................................... 89

5.2.1. Outline. ................................................................................................................. 89

5.2.2. Image data and processing. .................................................................................. 89

5.2.3. Process-based productivity. .................................................................................. 90

5.2.4. Biomass-based productivity. ................................................................................ 91

5.2.5. Seasonal analysis. ................................................................................................. 92

5.3. RESULTS AND DISCUSSION ........................................................................................... 93

5.3.1. Process-based seagrass productivity. ................................................................... 93

5.3.1.1. Literature estimates of seagrass maximum productivity (Pmax), light-

saturation threshold for photosynthesis (Ik) and photosynthetic efficiency (α). ......... 93

5.3.1.2. Output of process-based seagrass productivity analysis. .............................. 95

5.3.2. Biomass-based seagrass productivity and seasonal analysis ............................... 99

5.3.3. Conclusions, Perspectives and Future Work ...................................................... 106

Chapter 6: Conclusions, Research Limitations and Future Work ............................................ 109

6.1. SUMMARY .................................................................................................................. 109

6.2. CONTRIBUTIONS AND OUTCOMES ............................................................................... 110

6.3. LIMITATIONS AND FUTURE WORK .............................................................................. 112

List of References ........................................................................................................................... 115

Page 13: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xiii

List of Figures

Figure 2-1. Conceptual diagram and flowchart of the proposed study, with the subject matter of

individual chapters and their interconnection identified. ............................................................ 14

Figure 2-2. Data processing frameworks for A (chapter 3), B (chapter 4) and C (chapter 5) .... 15

Figure 2-3. Moreton Bay, East Coast Australia (ALOS AVNIR-2 14th July 2010) showing the

boundary of the Eastern Banks, the main study area, in the white rectangle. ............................. 18

Figure 3-1. Details of the study area and sampling sites. (a) Moreton Bay, East Coast Australia

(ALOS AVNIR-2 14th July, 2010) with the boundary of Eastern Banks, the main study area, in

the white rectangle. The solid green circles are sampling points of the diffuse attenuation

coefficient of Phinn and Dekker (2004) used in section 3.3.7. for testing the local Kd490–KdPAR

algorithm developed in this study. The red line is a transect traversing several parts of the study

area used to test the seasonal robustness of the local Kd490–KdPAR algorithm. (b) Extent of the

Eastern Banks seagrass in June 2012 (white rectangle in (a)) analysed from the database of the

Biophysical Remote Sensing Group of The University Queensland, which was also published in

Roelfsema et al. (2014). (c) The detailed area of the yellow rectangle in (a) and (b) where in situ

water optical measurements were conducted in February 2013 (solid yellow rectangles in (c)).

..................................................................................................................................................... 24

Figure 3-2. Measurement of downwelling irradiance using the profiling radiometer aboard the

R/V Scarus. Profiling was conducted by lowering the radiometer at several depths from which

the diffuse attenuation coefficient was then calculated. The measurement was conducted at

Wanga Wallen banks (Figure 3.1 c). ........................................................................................... 25

Figure 3-3. Illustration of satellite-based parameters with different pixel sizes used in this

study. ........................................................................................................................................... 27

Figure 3-4. Equipment setting for measuring seagrass leaf optical properties. a = ASD®

spectrometer; b = ASD® integrating sphere; c = ASD® internal light source. .......................... 29

Figure 3-5. Regression coefficient (R2) from exponential fitting between irradiance profile and

depth from which Kd was obtained. Figure (a) is for point 1 and figure (b) is for point 2. ........ 32

Figure 3-6. Kd (λ) profiles for (a) point 1 and (b) point 2. The blue lines and arrow indicate Kd

(λ) profiles over the course of the day. T symbol indicates time of measurement. The average of

Page 14: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xiv

Kd (λ) profiles for both points (red lines in (a) and (b)) were plotted together in (c) for ease

of comparison. ............................................................................................................................. 33

Figure 3-7. Bottom PAR expressed as blue, green and red wavelengths calculated from the

daily average of radiometer measurements at seagrass depth (the deepest radiometer profiling).

..................................................................................................................................................... 35

Figure 3-8. Sampling locations of diffuse attenuation of Phinn and Dekker (2004) at Amity

point (black solid triangle), and of this study (black solid cross) at Wanga Wallen banks, plotted

on KdPAR image of HICO data acquired on 3 February, 2013 (left), and on the natural composite

of ALOS AVNIR-2 14 July 2010 (right). ................................................................................... 37

Figure 3-9. Spectral absorptance of phytoplankton chlorophyll as measured spectrometrically

using the method of Parsons et al. (1984). The water samples were collected in Wanga Wallen

Banks during fieldwork in February 2013, at the time interval of 30–60 min (T1–T5). ............ 38

Figure 3-10. Seagrass surface area as determined based on PAR, PUR and % SI maps for (a)

summer, (b) autumn, (c) winter, and (d) spring, plotted against ALOS-AVNIR-2 imagery of 14

July, 2010 as the background image. (e) is the reference seagrass map used in this study

(Roelfsema et al., 2014) .............................................................................................................. 41

Figure 3-11. Spectral absorptance of Si (Syringodium isoetifolium), Cs (Cymodacea serrulata),

Zc (Zostera muelleri), Ho (Halophila ovalis ) and Hs (Halophila ovalis ) collected from the

Wanga Wallen Banks during the field campaign in February 2013 and measured using ASD®

integrating sphere (Figure 3.4). Estimated carotenoids absorptance area in the green band (Kirk,

2010) is indicated. ....................................................................................................................... 44

Figure 3-12. Pixel values with the depth range of 3.30–3.32 m plotted on the seagrass map of

the Eastern Banks (e.g., Figure 3.1c). The depth range value was determined from the

relationship between KdPAR and the maximum depth of colonisation of Duarte Figure 3.12.

Pixel values with the depth range of 3.30–3.32 m plotted on the seagrass map of the Eastern

Banks (e.g., Figure 3.1c). The depth range value was determined from the relationship between

KdPAR and the maximum depth of colonisation of Duarte (1991) (equation 12). ....................... 45

Figure 3-13. Calculation of Kd (λ) using empirical model of Austin and Petzold (1986) for (a)

point 1 and (b) point 2. ................................................................................................................ 47

Page 15: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xv

Figure 3-14. Local KdPAR algorithm developed by regressing in situ KdPAR against Kd490 data.

..................................................................................................................................................... 48

Figure 3-15. Performance of KdPAR algorithm of this study compared with published KdPAR

algorithms (equation 8–11b). The input of the in situ Kd490 data are from Phinn and Dekker

(2004) (solid green circles Figure 3.1 a). .................................................................................... 48

Figure 3-16. The comparison between calculated KdPAR using the local KdPAR algorithm and

the original Lee's KdPAR for all seasons. The analysis was performed using HICO data from

point transect traversing various area in the Eastern Banks (See Figure 3.1a). .......................... 49

Figure 4-1. Seagrass sampling points of June 2012 and February 2013 used in this study plotted

on seagrass map of June 2012 (Roelfsema et al., 2014) and ALOS AVNIR-2 of 14 July, 2010,

along with the names and boundaries of major individual banks in the Eastern Banks. ............ 56

Figure 4-2. Seagrass Biomass and LAI spatial patterns from field sampling in the clockwise

direction : biomass 2012, biomass 2013, LAI 2012 and LAI 2013, plotted on seagrass areal

distribution of June 2012 (dark green area) (e.g Roelfsema et al., 2014). .................................. 61

Figure 4-3. Plots showing the relationships between in situ LAI and WorldView-2 satellite

band coastal (a), blue (b), green (c), yellow (d) and red (e). ....................................................... 64

Figure 4-4. Average diffuse attenuation coefficients (Kd) measured in situ using a radiometer

device during a field trip in 2013 for two points close to the seagrass sample collection sites.

The measurement was conducted by lowering a radiometer to a number of depths in a 30 min–

1hr time interval. Kd was then calculated using standard light exponential principle of Beer's

law. The relative positions of blue, green and red wavelengths were indicated. ........................ 65

Figure 4-5. Light energy as measured using radiometer device at seagrass depth for two points

close to seagrass sample collection sites, expressed as blue, green and red lights. ..................... 66

Figure 4-6. Regression plot between in situ LAI and log Rb 3. The data used for the regression

were in situ LAI and WorldView-2 data of 2012. The complete results of the regression

analysis are given in Table 4.4. ................................................................................................... 69

Figure 4-7. Regression plots between LAI and biomass for community level of (a) 2012 and (b)

2013 data sets and (c) species Cymodocea serrulata of 2012 data set. The complete regression

result are given in Table 4.5. ....................................................................................................... 72

Page 16: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xvi

Figure 4-8. LAI and biomass maps of June 2012 (left) and February 2013 (right) calculated by

applying equation in Figure 4.5 and equation in Table 4.5 for community level (the firs list) to

WorlView-2 satellite data. ........................................................................................................... 74

Figure 4-9. Graphics of total surface area of each LAI class in each individual banks for June

2012 and February 2013 as calculated by applying equation in Figure 4.5 to WorlView-2

satellite data. ................................................................................................................................ 75

Figure 4-10. Graphics of total surface area of each biomass class in each individual banks for

June 2012 and February 2013 as calculated by applying equation in Table 4.5 for community

level (the firs list) to WorlView-2 satellite data. ......................................................................... 76

Figure 4-11. Plots between observed and estimated LAI of (c) 2012 and (d) 2013 and biomass

of (c) 2012 and (d) 2013 data sets. The observed values are in situ data that were not used for

model development. .................................................................................................................... 79

Figure 4-12. Transect line traversing various depth, biomass and values used for correlation

analysis. The depth variation is shown on the right-side panel. Seagrass area is delimited by

white line. Solid gray areas show land masses. ........................................................................... 80

Figure 5-1. ALOS AVNIR-2 (14 July 2010) coverage of the study area in the Eastern Banks,

Australia, with the names and boundaries of the individual banks indicated. ............................ 92

Figure 5-2. Maps of classified seagrass process-based productivity calculated using equation 1

(modified from Charles-Edwards et al., 1986) (left) and equation 2 (Jassby and Platt, 1976)

(right). The maps were produced from the summer image data set (e.g., the LAI input map was

processed from WorldView-2 image of 13 February, 2013 and the PAR map was from the

average of summer months of 2003–2012). ................................................................................ 97

Figure 5-3. Graphics of total surface area of productivity classes in each individual banks

calculated using the process-based methods. Process-based 1 was calculated using equation 1

(modified from Charles-Edwards et al., 1986) and process-based 2 was based on eq uation 2

(Jassby and Platt, 1976). .............................................................................................................. 98

Figure 5-4. Maps of classified seagrass productivity of June 2012 (left) and February 2013

(right) estimated using biomass-based method (equation 3). .................................................... 103

Page 17: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xvii

Figure 5-5. Graphics of total surface area of productivity classes in each individual banks

calculated using the biomass-based methods for June 2012 (winter) and February 2013

(summer). .................................................................................................................................. 104

Figure 5-6. Transect line traversing seagrass area in the study site plotted on the image-based

depth map of WorldView-2 (left) with its corresponding depth shown in the right side panel.

Seagrass area of June 2012 from Roelfsema et al. (2014) is bounded by white line. Solid gray

areas show land masses. ............................................................................................................ 105

Page 18: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xviii

List of Tables

Table 2-1. General Characteristics of Satellite Imageries Used in this Study ............................ 19

Table 3-1. Acquisition Dates of Satellite Data Used to Derive PAR and PUR Maps ................. 26

Table 3-2. The Ranges of Values Used to Generate PAR, % SI and PUR Maps as Depicted in

Figure 3.10 a, b, c and d. ............................................................................................................ 41

Table 3-3. Comparison Between Satellite and In Situ Radiometer Data .................................... 46

Table 4-1. Spectral Index Examined in This Study Calculated Using In Situ Reflectance Data 58

Table 4-2. Correlation Coefficients (r) of the LAI and Biomass Against the Resampled In Situ

Reflectance Data ......................................................................................................................... 63

Table 4-3. Correlation Coefficients (r) of LAI Against Biomass for Community (Total Species)

and Per Species of 2012 and 2013 data set ................................................................................. 68

Table 4-4. Regression Models, Coefficients of Determination (R2) and Probability

Corresponding to F Value (Pr > F, only available for linear model) for LAI ........................... 69

Table 4-5. Regression Models, Coefficients of Determination (R2) and Probability

Corresponding to F value (Pr > F) for Biomass ....................................................................... 71

Table 4-6. Comparison of Average Biomass per Individual Banks ............................................ 78

Table 4-7. The result of Pearson correlation analysis between depth, and LAI and biomas. All

values are significant at p < 0.05. ............................................................................................... 81

Table 5-1. Photosynthetic Parameters (Pmax, Ik, and α) Extracted and Modified from Lee et al.

(2007a) ........................................................................................................................................ 94

Table 5-2. Descriptive Statistics for Pmax, Ik, and α Given in Table 5.1. .................................. 94

Table 5-3. Average Seagrass Productivity Values of Individual Banks ...................................... 96

Table 5-4. Average Seagrass Productivity Values of Individual Banks .................................... 100

Table 5-5. Literature Estimates of the Productivity of the Five Seagrass Species Present in The

Eastern Banks ............................................................................................................................ 100

Page 19: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xix

Table 5-6. Literature Estimates of the Productivity of Seagrass Communities Containing One or

More Seagrass Species Present in The Eastern Banks ............................................................. 101

Table 5-7. The result of Pearson correlation analysis between productivity, PAR and depth. All

values are significant at p < 0.05. The productivity, PAR and depth values included in the

analysis were from transect line depicted in figure 5.6. ........................................................... 105

Page 20: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

xx

• ALOS-AVNIR-2 : Advanced Land Observing Satellite-Advanced Visible and Near

Infrared type 2

List of Abbreviations used in the thesis

• AOPs : Apparent Optical Properties

• ETM : Enhanced Thematic Mapper

• GPP : Gross Primary Production/Productivity

• HPLC : High Performance Liquid Chromatography

• HICO : Hyperspectral Imager for the Coastal Ocean

• IOPs : Inherent Optical Properties

• LAI : Leaf Area Index

• MODIS : Moderate Resolution Imaging Spectroradiometer

• MERIS : Medium Resolution Imaging Spectrometer

• NPP : Net Primary Production/Productivity

• PAR : Photosynthetically Active Radiation

• % SI : Percentage of Irradiance relative to Surface Irradiance

• PUR : Photosynthetically Utililised Radiation

• SAV : Submerged Aquatic Vegetation

• SeaDAS : SeaWIFS Data Analysis System

• SeaWIFS : Sea-Viewing Wide Field-of-View Sensor

• TM: : Thematic Mapper

• WV-2 : WorldView-2 sensor/imagery

Page 21: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

1

Chapter 1: Introduction and Significance of the Research

1.1. Introduction and Overview of the Research Context

Seagrasses are aquatic angiosperm that form grass-like meadows, which in addition to

mangrove and coral reefs, serve as ecologically critical habitats. Seagrass meadows are nursery

grounds for many commercial and recreational fish species and act as a filter for coastal waters,

taking up nutrients and contaminants from the water and causing suspended sediment to settle

(Short et al., 2006). Seagrass ecosystems are important in the mitigation of climate change because

of their role as a natural sink for atmospheric CO2 – a shallow water “biological pump” – and may

be responsible for the significant role of the ocean's annual net CO2 uptake (Duarte et al., 2010).

The term “blue carbon” was coined for a research area addressing the role of important wetland and

coastal ecosystems such as mangrove, salt marsh and seagrasses in binding anthropogenic CO2

(Nelleman et al., 2009). Considering these important functions and the global decline of seagrass

(e.g. Orth et al., 2006; Waycott et al., 2009) it is critical to understand the environmental limiting

factor(s) and the behaviour of the biophysical parameters of seagrass for the sustainable

management of this habitat (Duarte et al., 2008; Bjork et al., 2008).

Light availability has been identified as essential factors controlling the temporal and spatial

distribution, productivity and biomass of seagrass (Preen, 1992; Denison et al., 1993; Zimmerman,

2006; Lee et al., 2007a; Bjork et al., 2008). In terms of in-water surface irradiance (% SI) seagrasses

have high light requirements, ranging from 2.5% (Erftemeijer and Robin Lewis III, 2006) to 40%

(Collier and Waycott 2009; Bach et al., 1998; Grice et al., 1996), depending on species or genus.

The role of light becomes more important for seagrass inhabiting shallow coastal areas where the

light climate is sensitive to sediment re-suspension and river run-off (Preen, 1992; Longstaff, 2003;

O’Brien et al., 2011). For conservation purposes it is advisable to monitor suitable light parameters

with which the underwater light environment of seagrass can be assessed (Anastasiou, 2009). One

important underwater light parameter is Kdλ, the spectral diffuse attenuation coefficient used to

assess water clarity and light availability (i.e., PAR, photosynthetically available radiation) at the

depth of seagrass (Gallegos, 1994; Kirk, 2010). Combined with the absorptance property of

seagrass leaf, PAR can be analysed to derive PUR (photosynthetically useable radiation), which

determines the fraction of radiant energy of PAR absorbed by a given plant for a given wavelength

(Morel, 1978). PUR emphasises the importance of the portion of light utilised for photosynthesis.

While regular point-based light monitoring programs can provide in situ data to further the

understanding of long-term water quality patterns at specific sampling locations, their extrapolation

Page 22: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

2

to other locations or times may be problematic due to infrequent measurements at pre-defined

discrete sampling locations and spatial heterogeneity of the optical properties of water.

In addition to light parameters, leaf area index (LAI), biomass and productivity are

important biophysical parameters characterising the process and function of seagrass habitat.

Physically, LAI relates more to the canopy structure of seagrass and represents a measure of

photosynthetic area, which in turn determines biomass. Consequently, LAI is often used as an input

parameter for modelling of seagrass productivity (e.g., Plus et al., 2001; Zimmerman, 2003,

McPherson et al., 2011). A high rate of oxygen production, which is a by-product of photosynthesis,

is produced by high seagrass biomass, supporting life in coastal environment by supplying oxygen

to the surrounding waters (Terrados and Borum, 2004). For marine biota that feeds on seagrass, the

rate of seagrass primary productivity determines seagrass habitat capability to support grazing and

at the same time, is an estimate parameter for the impact of grazing on seagrass (Preen, 1992).

Seagrass habitat is among the most productive aquatic ecosystems. Productivity is

commonly expressed as gram carbon or dry weight produced per area per day. Seagrass

productivity stands out compared with other aquatic and terrestrial plants with its annual total net

primary productivity (NPP) reaching 0.49 Pg C · m-2 · year-1 for an area of 0.6 x 106 km2 (Mateo et

al., 2006). Seagrass LAI, biomass and productivity are usually measured in situ using field methods

(Erftemeijer et al., 1993; Short et al., 2001). For LAI and biomass, destructive methods of

harvesting the seagrass still dominate. Productivity is commonly measured in the field or laboratory

using simple methods such as biomass accumulation, leaf marking, plastochron interval techniques

and lepidochronological analysis or more complex metabolic methods such as oxygen evolution in

open or closed systems and carbon-14 tracer techniques (Erftemeijer et al., 1993; Short et al., 2001)

The developments in sensor technology and data storage have made remote sensing

technology a promising tool for the study of coastal and marine environments, with the primary

advantage of its synoptic, continuous and repetitive measurements (review by Mumby et al., 2004

and Eakin et al., 2010). This type of measurement can complement the typical point-based field

programs to map, model and monitor coastal and marine physical environments and the biophysical

parameters of coastal and marine habitats on a large scale. Although satellite-based light and

productivity parameters have been in operational use for oceanic waters (e.g., ocean surface primary

production), there are few studies addressing their application to seagrass environments. This

includes mapping the light environment of seagrass and analysing how it may affect seagrass spatial

Page 23: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

3

distribution. Addressing these issues using remote sensing methods would be advantageous for

mapping and monitoring light climate in seagrass and for understanding the spatial interaction

between light and seagrass. Likewise, despite the importance of LAI, biomass and productivity for

characterising the condition of seagrass habitat, there are few published works on remote sensing

focussing on those parameters that consider the inherent relationship between reflectance and the

parameters of interest, the inter-relation among biophysical parameters and the temporal variation

of the parameters concerned. The present study aims to fill these gaps by integrating in situ

reflectance measurements, laboratory-based analysis and multi-temporal image analysis to

investigate light climate, LAI, biomass and the productivity of seagrass habitats in the Eastern

Banks, Australia.

1.2. Background

1.2.1. Mapping light in seagrass environments: Extending the principles of ocean

optics to the remote sensing approach.

As an aquatic plant, seagrass has an abundance of the water required for growth, a

favourable environmental condition not shared by terrestrial plants. When light, another essential

environmental factor for plant growth, penetrates the water its quality and quantity reduces due to

scattering and absorption by the water itself and other light-attenuating substances (Kirk, 2010).

With the addition of its inefficient carbon concentrating mechanism for photosynthesis, this results

in seagrass having high light requirements of 2.5% (Halophila decipiens) (Erftemeijer and Robin

Lewis III, 2006) to 40% (Zostera sp and Cymodocea sp) (Collier and Waycott, 2009; Bach et al.,

1998; Grice et al., 1996), depending on species or genera. In addition, even within one species

Longtaff (2003) observed that annual percentages of surface light for Zostera capricorni (now

Zostera muelleri) of four locations in Moreton Bay were 35, 31, 36, 16, and 6. This variation was

caused by spatial and temporal variation of light reduction in the area (Longstaff, 2003).

Consequently, knowledge of the behavior of light in coastal waters would be beneficial to

understanding seagrass in term of its distribution, density and productivity.

Preisendorfer (1976) and Jerlov (1976) presented the fundamental understanding on the

nature of light propagation into the water column and introduced the terms inherent optical

properties (IOPs) and apparent optical properties (AOPs). IOPs are the optical properties of water

defined by the substances of the water itself and are commonly expressed as absorptance coefficient

(aλ) and scattering coefficient (bλ). IOPs are like a fingerprint for a particular type of water; if the

Page 24: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

4

water substances change to different types, then its IOPs change accordingly. Both IOPs and the

light field (e.g., sun zenith angle) determine AOPs. In the context of using remote sensing methods

for aquatic environmental application, one important AOP is the remote sensing reflectance (Rrs)

defined as the ratio of water leaving radiance (Lw) and downwelling irradiance (Ed). The water

content is commonly expressed using a three component model: phytoplankton, suspended matter

and coloured dissolved organic matter (cdom) (IOCCG, 2000; Sathyendranath et al., 1989). The

proportion of these three components present in a water mass defines the water’s aλ and bλ and

consequently, its AOPs. It is now commonly accepted that the radiative transfer equation for

explaining light behavior in water has been numerically solved (Mobley, 1994). This means that

calculation to achieve numerical closure, for example between IOPs and AOPs, can be accurately

performed. This success leads to the extraction of IOPs and water-attenuating substances from

remote sensing reflectance using various models of empirical, semi-analytical or analytical (Dekker

et al., 2006).

Some important light parameters for explaining submerged aquatic vegetation (SAV)

distribution are the diffuse attenuation coefficient (Kdλ), photosynthetically available radiation

(PAR) and photosynthetically usable radiation (PUR) (Zimmerman, 2003, 2006). Kdλ defines the

reduction rate of light into the water and since only the visible range (400–700 nm or PAR) is

relevant for seagrass photosynthesis, it is commonly expressed as KdPAR. Using KdPAR and depth

information PAR at SAV depth can be calculated. PUR is a conceptual extension of PAR defining

the amount of PAR absorbed by SAV for photochemical processes, for example, photosynthesis.

PUR is calculated by multiplying PAR and leaf absorptance (AL). KdPAR and PAR are commonly

the routine products of ocean colour satellites, some are still operational (e.g., MODIS, HICO) and

some are in archive products (e.g., MERIS, SeaWIFS) at resolutions between 90 m (HICO) and 9

km (SeaWIFS). In addition to KdPAR, another product commonly used for studying ocean

environment is Kd490, which is the diffuse attenuation coefficient at wavelength 490 nm. These Kd

products are calculated using either empirical or semi-analytical methods (Lee et al., 2005a, 2005b).

An empirical method is used to produce Kd from established correlation between the satellite

reflectance and in situ Kd. Other researchers also related Kd to chlorophyll concentration as an

intermediate step to calculate Kd (e.g., Morel et al., 2007). Semi-analytical methods employ

the numerical solution of the radiative transfer equation previously mentioned by inverting

satellite reflectance to IOPs (e.g., aλ and bλ) and then calculating Kdλ from these parameters

(Lee et al., 2005a, 2005b).

Page 25: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

5

Using these light parameter products, a wide variety of studies on marine environments have

been conducted, for example, studies on ocean surface productivity (review by Carr et al., 2006)

and water clarity status in the Great Barrier Reef, Australia (Weeks et al., 2010). Remote sensing

studies on the light climate of shallow coastal environments – a typical environment for seagrass

habitat – were scarce. If any, analyses were conducted at a global scale (Gattuso et al., 2006) using

satellite resolution of > 1 km. This lack of analyses for shallow coastal waters may be due to the

spatially complex and highly dynamic coastal environments not resolvable by common ocean

colour satellite resolution and issues concerning an appropriate algorithm to deal with turbid waters

(e.g., case 2 waters). Addressing the need for investigating light in seagrass environments requires

an understanding of in situ light quality and quantity, higher spatial resolution of satellite image

data and suitable remote sensing algorithms for deriving water column diffuse attenuation.

1.2.2. Remote sensing applications for mapping seagrass biophysical properties.

Remote sensing technology provides repeatable coverage, over areas of varying size, which

may not be obtainable using point-based sampling measurements (Green et al., 1996, 2000; Mumby

et al., 2004; Eakin et al., 2010). The repetitive observation capability of imaging sensors over the

same area (temporal resolution) provides the opportunity to monitor the marine environment and

changes in its composition, structure and function. Early applications of remote sensing methods for

coastal habitat change detection include monitoring marine habitat change in Bahrain (Zainal et al.,

1993) and in the Florida Keys (Dustan et al., 2001), using the first generations of Landsat data.

Other studies have identified the species of seagrass from its spectral reflectance (Fyfe, 2003;

Thorhaug et al., 2007). Fyfe (2003) concluded species differentiation was possible using

hyperspectral sensors and some studies have successfully mapped seagrass species using broad

band data from Landsat 5 TM and 7 ETM (Dekker et al., 2005) and Quickbird-2 (Phinn et al., 2008;

Lyons et al., 2011). The same success stories were also shared with seagrass biomass mapping by

correlating in situ biomass data and multispectral remote sensing data from Landsat TM

(Armstrong, 1993), Quickbird-2 (Phinn et al., 2008) and IKONOS (Knudby and Nordlund, 2011).

The list continues with the application of remote sensing to seagrass cover and species (e.g., Phinn

et al., 2008; Lyons et al., 2011, 2012, Roelfsema et al., 2014).

Similar to the extraction of light parameters from remote sensors, three methods of

empirical, semi-analytical and analytical also apply to seagrass (Dekker et al., 2006). Frequently

used methods for mapping seagrass biomass and cover are empirical, using regression analysis

between those parameters and remote sensing reflectance (Armstrong, 1993; Phinn et al., 2008;

Page 26: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

6

Lyons et al., 2011, 2012). To improve species mapping, a semi-analytical method can be used by

obtaining a priori knowledge of seagrass spectral reflectance corresponding to the satellite bands

used (Dekker et al., 2006). This knowledge can be collected using in situ information or spectral

libraries. The analytical method can be considered the most accurate, requiring that all variables

affecting the remote sensor signal be quantified to isolate extraction of the parameter(s) of interest.

In addition to cover and species, important seagrass biophysical parameters commonly

derived from remote sensing are leaf area index (LAI), the single-side leaf area per unit ground area

(Green et al., 1997) and biomass. Biomass and LAI are important factors for characterising plant

canopy structure, process and growth. LAI represents the total photosynthetic area of the plant

canopy per unit area of sediment, which has been used as a diagnostic variable for crop growth rate,

radiation intensity and above-ground height (Solana-Arellano et al., 2003). Biomass and LAI have

also been used for modelling seagrass photosynthesis (Plus et al., 2001; Zimmerman et al., 2003,

McPherson et al., 2011). While several studies have attempted to map biomass from remote sensing

using various methods, only Dierssen et al. (2003) and Wicaksono and Hafitz (2013) have explicitly

attempted to map seagrass LAI using remote sensing data. Despite some differences in their

procedure, their studies used an empirical method relating in situ LAI and remote sensing data.

Dierrsen et al. (2003) developed their LAI algorithm from the established correlation between in

situ LAI and measured underwater reflectance and then applied it to experimental airborne

hyperspectral images for low-light spectroscpy (Ocean PHILLS) with 1.25 m resolution.

Wicaksono and Hafitz (2013) correlated their in situ LAI data to principal-component-analysis

transformed bands of ALOS AVNIR-2 and ASTER images. While the method used by Dierssen et

al. (2003) can be considered as ideal for extracting image-based LAI by involving rigorous water

optical correction, in situ bottom reflectance measurement and the use of hyperspectral imagery,

their method was applied to mono-typic turtlegrass that has a relatively larger size compared with

seagrass commonly found, for example, in the Eastern Banks, Moreton Bay, Australia.

Regarding the use of remote sensing to derive vegetation LAI and biomass, empirical

methods still dominate, commonly implemented by regressing in situ LAI or biomass and image

reflectance (Gray and Song, 2012; Wulder et al., 1998). A more rigorous method for producing

satellite-based LAI or biomass map has been developed for terrestrial vegetation by inverting the

canopy reflectance models (e.g., Myneni et al., 1997; Peddle et al., 2004). This method uses a

radiative transfer approach for forest canopy by taking into account solar elevation, diffuse and

direct radiation, and the spectral quality of shaded and sunlit portions of the canopy and understory

Page 27: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

7

(Wulder et al., 1998). Similar models for seagrass have also been developed (Zimmerman, 2003;

Hedley and Enrique, 2010) but extension of their use to remote sensing applications requires further

research. The analytical method has a strong physical basis and is not restricted to a specific biome

type but there are issues with its parameterisation and its mathematic solutions may not be unique

(Gray and Song, 2012). On the other hand, the empirical method has limitations in which the

relationship between image reflectance and LAI may reach saturation at moderate-to-high LAI

values. The empirical method may also lack in temporal dimension, as the resultant LAI and

biomass maps are single representations in time. However, the empirical method is the one most

frequently and successfully used to derive LAI from image data (Nemani et al., 1993). Considering

that LAI and biomass are important variables for characterising seagrass ecosystems, improvements

for extracting these parameters from remote sensing and understanding their spatial and temporal

distribution would be advantegous for remote sensing studies and seagrass conservation.

1.2.3. Estimating productivity using remote sensing data: A review of applications to

terrestrial, ocean surface and coastal environment.

Estimation of net primary productivity (NPP) is achieved using two main ecosystem

modelling approaches in terrestrial ecosystems: (a) physiological process-based models, which use

detailed biophysical and meteorological inputs, and (b) diagnostic models, which use remotely

sensed data (Olofsson et al., 2008). The second approach is based on the light-use efficiency

concept developed by Monteith (1972, 1977) who found a conservative ratio between absorbed

photosynthetically active radiation (APAR) and NPP. The concept was later improved to model

NPP as a function of incident PAR, the fraction of PAR absorbed by vegetation (FPAR) and light-

use efficiency (ε, amount of O2 evolved per absorbed light) (Olofsson et al., 2008). This model is

attractive for large-scale applications, since the parameters are obtainable from satellite data

resulting from research that tested the extraction of each parameter from satellite data. For example,

Liang et al. (2006) developed a new method to estimate instantaneous incident PAR from MODIS

satellite data, based on the look-up table approach. Wang et al. (2010) recently improved the

method to estimate daily-integrated incident land PAR by applying both LUT (Look-up table) and

sinusoidal interpolation methods to MODIS satellite data. Their result showed the LUT method was

more accurate over the sinusoidal interpolation method and the calculated daily-integrated PAR was

comparable to the PAR of field measurements and of Geostationary Operational Environmental

(GOES) Satellites PAR products (Wang et al., 2010).

Page 28: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

8

The calculation of FPAR from remotely sensed imagery is relatively straightforward since

FPAR is strongly correlated to the variation in the red and near-infrared bands, which are used in

the normalised difference vegetation index (NDVI), a routine product of satellite data (Olofsson and

Eklundh, 2007). One challenging task for Monteith’s model is the extraction of ε, the light-use

efficiency term (equal to, ϕ, quantum yield) (Dubinsky et al., 1984; Hochberg and Atkinson, 2008),

from remote sensing data. ε is a measure of O2 evolved per photon absorbed. (Dubinsky et al., 1984;

Hochberg and Atkinson, 2008). Some studies found that the assumption of ε, as a constant for

Monteith’s original model, did not hold and ε was found to vary between vegetation types and in

response to environmental factors (Sims et al., 2006). Despite initial success in estimating ε from

the photochemical reflectance index (PRI) using spectral data in terrestrial vegetation (Gamon et al.,

1992, 1997), this relationship was found later to vary with the function of vegetation types and the

time of measurement (Sims et al., 2006). Consequently, some NPP models assign different ε for

different vegetation types, using the look-up table approach where ε is first set to the maximum

value and then degraded, based on environmental stress factors (review by Sims et al., 2006).

However, apart from ongoing research in formulating ideal ε, or any other parameter of the

improved Monteith’s model, land NPP has been an operational remote sensing product and is

widely used either to define terrestrial CO2 flux for climate change modelling, or to measure

products of vegetation growth such as crop yield and forest production.

Satellite-based ocean primary productivity models are based on the relationship between

phytoplankton photosynthesis and available light, either in the form of a photosynthesis-light model

(e.g., Jassby and Platt, 1976) or a chlorophyll-based model (Behrenfeld and Falkowski, 1997).

Behrenfeld and Falkowski (1997) and Falkowski and Raven (2007) classified numerous

mathematical models for calculating vertically-integrated primary production in aquatic ecosystems

into four categories: (a) wavelength-resolved models (sensu Behrenfeld and Falkowski, 1997) or

full spectral models (sensu Falkowski and Raven, 2007), (b) wavelength-integrated models, (c)

time-integrated models, and (d) depth-integrated models. Although these models vary in

mathematical expression or their underlying assumptions, they generally estimate oceanic primary

production as a function of surface phytoplankton biomass (a photo-adaptive variable), euphotic

depth (irradiance dependent) and day length. Surface phytoplankton biomass estimation is

commonly approached using ocean surface chlorophyll, which is a routine product from ocean

colour satellites such as MODIS and SeaWIFS. It is further explained that in spite of model

variability, two primary controlling factors on both the accuracy and possible improvements of the

models are the parameterisation of input chlorophyll data (surface phytoplankton biomass) and the

Page 29: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

9

light photosynthetic efficiency (photoadaptive variable) (Behrenfeld and Falkowski, 1997;

Campbell et al., 2002; Carr et al., 2006).

The diverse variety of ocean-surface productivity models and the different ocean regions for

which they are specifically developed necessitates testing and comparisons between models. These

studies concluded that some models performed differently in different regions, for example, in the

Southern Ocean they gave the most diverse results between models and required improvement to

accommodate extremes in sea surface temperature and chlorophyll concentrations (Campbell et al.,

2002; Carr et al., 2006). Recent studies by Saba et al. (2010) compared 36 ocean productivity

models with nearly two decades of in situ productivity data at the Bermuda Atlantic Time series

Study (BATS) and the Hawaii Ocean Time series (HOT) stations. They found that satellite-based

ocean productivity data was more in agreement with in situ productivity data when chlorophyll a,

used as input for the model, was derived from the HPLC (High Performance Liquid

Chromatography) method, rather than from fluorometric or satellite data. This result implied that

the availability of time series HPLC-based chlorophyll a, or high quality satellite-based chlorophyll

a, which is comparable to the HPLC-based method, could improve the performance of satellite-

based ocean primary productivity models.

For coastal habitats, there are few studies using remote sensing data for estimating coral reef

productivity (e.g., Atkinson and Grigg, 1984; Ahmad and Neil, 1994; Andrefouet and Payri, 2001;

Brock et al., 2006). It was only the work by Hochberg and Atkinson (2008) that went beyond area

scaling-up of reef productivity, by applying the light-absorptance model of Monteith (1972, 1977)

to a coral reef environment, on a pixel-by-pixel basis. Using this method, continuous values of

spatial reef productivity were produced, which they claimed were in the range of standard reef

productivity values detailed by Kinsey (1985). Anomalous lower productivity values in the deeper

fore-reef area were attributed to the single/standard light-use efficiency employed, which was

expected to increase with depth (Hochberg and Atkinson, 2008). Besides PAR, two other key

parameters of Hochberg and Atkinson’s work are fraction of PAR absorbed by the reef community

(FPAR) and ε, light-use efficiency, a measure of O2 evolved per photon absorbed. They approached

FPAR as a product of light absorptance (A) derived from the inversion of remote sensing

reflectance (R) such that A = 1 - R. The underlying assumption of their method was that the light

not reflected by the benthic community was used for photosynthesis. The pigments responsible for

the photosynthetic processes residing in photosynthetic organisms absorb specific wavelength

Page 30: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

10

(energies) of light and reflect all others, which are then converted to heat or fluorescence

(Falkowski and Raven, 2007; Al-Najjar, 2010).

Morel (1978) proposed a concept of light utilisation for marine photosynthesis processes by

introducing terms of PAR (photosynthetically available radiation), PUR (photosynthetically usable

radiation) and PSR (photosynthetically stored radiation). PUR is defined as the fraction of PAR that

can be absorbed by algae and is dependent on the photosynthetic absorption coefficient of the algae

and the spectral composition of the underwater light. PSR describes a particular fraction of the

absorbed energy used in the photosynthetic process, for conversion into chemical energy in the

form of organic matter (Morel, 1978). In relation to Hochberg and Atkinson (2008), while

reflectance plays a role as a discriminating factor for defining bottom-type categories (e.g.,

Hochberg et al., 2003) it may not indicate the portion of light used for photosynthetic processes

(PSR). Estimation of photosynthesis using PAR could overestimate it, as only a portion of PAR is

absorbed by photosynthetic pigments (Zimmerman, 2003). This so-called PUR is a function of PAR

and leaf optical properties.

For seagrass environments, Dierssen et al. (2010) were the first to estimate seagrass net

primary productivity, in the Great Bahama Bank, using an empirical algorithm applied to SeaWIFS

imagery. Their NPP algorithm was produced by regressing the derived bottom reflectance at

555 nm against NPP from either leaf growth measurement, or from NPP estimated from LAI. As

the nature of the algorithm is empirical, it may not be accurate for the generation of NPP at other

sites, unless some carefully defined assumptions can be justified (e.g., similarity of seagrass

environment).

With recent advances in radiative transfer modelling, it is now possible to calculate

parameters unobtainable from in situ or even laboratory measurement of seagrass (e.g., canopy

KdPAR and light absorptance) from the newly developed, three-dimensional radiative transfer

models (Hedley and Enrique, 2010). Previously, seagrass leaf absorptance was able to be measured

only using spectrometer technique (e.g. Carruthers and Walker, 1997; Perez-Llorens and Niell,

1993 in Plus et al., 2001; Enriquez et al., 2005). Considering that bio-optical or mathematical

modelling for seagrass (e.g., Zimmerman, 2003) or vegetation productivity (e.g., Charles-Edwards

et al, 1986) exists and light parameters (e.g., KdPAR, PAR) and biophysical parameters influencing

productivity (e.g., LAI and biomass) can be derived from remote sensing data, it may be possible to

estimate and model seagrass productivity using remote sensing.

Page 31: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

11

1.3. Problem Statement

Light availability, LAI, biomass and productivity are among the most important parameters

for characterising and modelling seagrass habitat (Zimmerman, 2003; Lee et al., 2007a), but remote

sensing studies facilitating the representation of these parameters on a large scale are still either

limited, or face technical and application challenges. Given the high value of seagrass as a primary

producer, understanding the light availability that is vital for seagrass photosynthesis is of

considerable environmental and economic concern. The light requirement of seagrass is often

expressed in percentage difference relative to surface in-water light, termed as percentage surface

irradiance (% SI). With this measure some seagrass management areas set minimum % SI targets

for total light (PAR) reaching the bottom as a primary indicator for seagrass monitoring

(Anastasiou, 2009). However, as light may vary horizontally and vertically there is a knowledge

gap on how to present light in a spatially explicit representation from which large-scale light

measurements, monitoring and analysis can be performed efficiently and effectively. Previous

review studies on the light requirements of seagrass found that seagrass growth rates dropped

markedly below PAR of 5 mol photons · m-2 · day-1 (Gattuso et al., 2006; Lee et al., 2007a),

implying that seagrass spatial distribution may follow certain range of PAR values, which may be

identified using satellite-based light parameters (e.g., PAR, PUR and % SI). To date, only the

published work of Gattuso et al. (2006) attempts to estimate potential surface area of benthic

habitat, including seagrass, from PAR and % SI using SeaWIFS satellite data at global scale.

However, as mentioned in their paper recommendation section, their study did not consider

temporal analysis and in situ validation and they used a chlorophyll-based algorithm to derive

KdPAR, which tended to underestimate KdPAR in coastal waters (Lee et al., 2005a, 2005b; Sauquin et

al., 2013; Zhao et al., 2013). Investigating the light climate in seagrass inhabiting spatially complex

and highly dynamic coastal environments, at local and regional scale, requires higher satellite

spatial resolution, a robust remote sensing algorithm for case 2 waters, in situ light measurements as

validation and multi-temporal satellite analysis.

Although some authors have initiated the mapping of seagrass LAI, biomass and

productivity using remote sensing methods, empirical methods relating image reflectance and

parameters of interest were still dominant. However, this method may produce spurious results as

causal relationships may not exist between image reflectance and the parameters studied. For

benthic habitat applications, atmospheric signals and water column effects can also interfere with

image reflectance, making the empirical method more challenging. A refinement of the remote

sensing method for studying LAI and biomass should include an examination of the relationship

Page 32: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

12

between reflectance and seagrass LAI or biomass using in situ reflectance data, the examination of

allometric relationships between seagrass LAI and biomass, and multi-temporal analysis to

accommodate temporal variation of the two parameters (e.g., Kocak et al., 2011; Preen et al., 1992).

In the case of seagrass productivity, improvement could be made by employing suitable bio-optical

or mathematical models with which image-based light, biomass and LAI parameters can be

integrated to estimate seagrass productivity. The innovation and significance of this attempt, if

successful, is that the estimation of productivity using remote sensing may be extendable temporally

and spatially as the method does not rely on a regional or local empirical relationship between

productivity and remote sensing reflectance.

1.4. Aim

The aim of this research is to characterise the spatial and temporal distribution of light

climate, LAI, biomass and productivity of seagrass in the Eastern Banks, Australia by integrating in

situ measurements, laboratory analyses, bio-optical modelling and remote sensing methods.

1.5. Objectives

Objective 1: To determine light quality and quantity in seagrass environments and

the relationship between seagrass surface area and light parameters using field and remote

sensing data.

Objective 2: To develop a method for mapping the spatial and temporal distribution

of seagrass LAI and biomass using WorldView-2 satellite data.

Objective 3: To estimate the spatial and temporal distribution of seagrass gross

primary productivity using the combination of modelling and remote sensing.

1.6. Thesis Outline

Chapter 1. Introduction

This chapter introduces the research context of the thesis, describes the importance of light,

LAI, biomass and productivity in seagrass environments and addresses the remote sensing approach

for studying these parameters and possible refinements. The chapter consists of an opening

introduction, background, problem statement, aim, objectives and thesis outline.

Page 33: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

13

Chapter 2. Research Approach

This chapter presents the general overview of the research methods including a flowchart

and conceptual diagram, a brief description of the study site and the field/in situ data collected, as

well as remote sensing data used.

Chapter 3. (Objective 1, Paper 1)

The thesis began with characterising the KdPAR, PAR and PUR of the seagrass environment

in the study site using a combination of in situ measurement and remote sensing methods. This

stage was designed to investigate light quality, map light quantity and determine whether light

parameters could be used to explain the spatial distribution of seagrasses in the study site.

Chapter 4. (Objective 2, Paper 2)

The next stage was to develop a method to map seagrass LAI and biomass from remote

sensing data. In this stage, a systematic method using field, laboratory and remote sensing based

approaches was used to produce satellite-based LAI and biomass maps.

Chapter 5. (Objective 3, Paper 3)

The last stage was a synthesis to integrate the results from previous chapters for mapping

seagrass productivity. This last chapter uses light parameters, LAI and biomass produced from the

previous chapters as inputs for seagrass productivity modelling.

Chapter 6. Conclusions, Contributions and Research Limitations

This chapter presents the conclusions and summary of the main research outcomes and

contributions based on the research objectives. The research limitations and directions for future

works are also given.

Page 34: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

14

Chapter 2: Research Approach

2.1. General Overview of Research Methods

The present study aims to address the following research questions applied to the Eastern

Banks, Moreton Bay, Australia:

1. How can the KdPAR, PAR and PUR of the spatially-complex seagrass environment be

mapped using the remote sensing method?

2. Can the satellite-based light parameters explain the spatial distribution of seagrass in

the study site?

3. How accurately can seagrass LAI and biomass be modelled and mapped from remote

sensing data?

4. How can remote sensing data and the existing bio-optical model be integrated to

estimate seagrass gross primary productivity?

Figure 2-1. Conceptual diagram and flowchart of the proposed study, with the subject matter of individual chapters and their interconnection identified.

Investigation of Kd, PAR and PUR using field survey and remote sensing approaches (Chapter 3)

Estimation of seagrass productivity using the integration of a bio-optical model and remote sensing methods (Chapter 5)

Mapping seagrass LAI and biomass using high resolution remote sensing data (Chapter 4)

Remote sensing data

Seagrass productivity models

Seagrass core sample analysis

= seagrass underwater light environment. = seagrass biophysical parameters = seagrass productivity

Page 35: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

15

Figure 2-2. Data processing frameworks for A (chapter 3), B (chapter 4) and C (chapter 5)

Chapter 3 focuses on the characterisation of diffuse attenuation coefficient (KdPAR),

photosynthetically available radiation (PAR), bottom-to-surface ratio of PAR (%SI) and

photosynthetically usable radiation (PUR) using integration of field survey and remote sensing

approaches. Light quality and quantity were determined using in situ irradiance profiling

measurements during tidal cycles. Using the combination of PAR from MERIS satellite, KdPAR

from HICO satellite, bathymetry from WorldView-2 satellite and laboratory-based seagrass leaf

Page 36: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

16

absorptance coefficient, satellite-based light maps of PAR, % SI and PUR were produced and used

to determine potential seagrass surface area. Satellite-based PAR and % SI were then compared

with their corresponding in situ data. The performance of empirical, semi-analytical and locally

developed algorithms of KdPAR were also compared.

Chapter 4 develops a systematic method for mapping seagrass LAI and biomass using

WorldView-2 satellite data. It started with an examination of the relationship between reflectance

and LAI and biomass using in situ reflectance collected simultaneously with seagrass samples.

Regression models for estimating LAI and biomass from reflectance were then developed, based on

analysis using in situ measurements. The allometric relationship between in situ seagrass LAI and

biomass was also examined. An accuracy assessment of the resultant satellite-based LAI and

biomass was conducted at the end of the chapter along with temporal analysis of LAI and biomass

for the study site.

Chapter 5 develops a method to estimate seagrass gross primary productivity from remote

sensing data, using the results produced in the previous chapters. Instead of using the site-specific

empirical relationship between spectral data and in situ productivity, this chapter attempts to apply

suitable model(s) to remote sensing data. Productivity was determined using two methods: the

biomass-based model that converts biomass to production rate, and a model that requires PAR, LAI,

canopy extinction coefficient, photosynthetic efficiency and maximum production (termed as

process-based model). Seasonal analysis of seagrass productivity was also conducted.

Overall, the study attempted to fill the knowledge gap on the local perspective of the spatial

and temporal distribution of the light climate, LAI, biomass and productivity of seagrass habitat by

using remote sensing as the primary tool.

Due to the nature of the research topic, it is not possible to organise the thesis in a common

traditional linear format. Figure 2.1. provides a flowchart of the methods used to meet the objectives

and describes the interconnection between various stages in this study. The flowchart also serves as

a conceptual diagram for the present study.

2.2. Study Site

The study site is the seagrass dominated region of the Eastern Banks, Moreton Bay,

Australia (27° 23' 24.38" S, 153° 22' 33.06" E) (Figure 2.2.) situated at the south-east corner of the

state of Queensland, on the east Australian coast. Six seagrass species are abundant in the area :

Halophila spinulosa, Halophila ovalis, Halodule uninervis, Zostera muelleri, Cymodocea serrulata,

and Syringodium isoetifolium, with Zostera muelleri present as the most dominant species in the

Eastern Banks (Preen et al., 1992). Seagrass occurs with specific zonation in the five individual

Page 37: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

17

banks present in the area: Amity, Chain, Maroom, Moreton and Wanga Wallen banks (Hyland et

al., 1989). Most of the dense seagrass in Moreton Bay was found in shallow intertidal areas with a

depth of < 1 m of State Chart Datum, whereas subtidal communities were found at depths of < 3 m

of water at low water (Hyland et al., 1989). Seagrass was absent at depths > 10 m (Hyland et al.,

1989). Due to the abundance of seagrass and the transformation of the western region into the

anthropogenic waterfront of Brisbane's eastern suburbs, the Eastern Banks is the major habitat for

dugong in Moreton Bay (Preen, 1992). The Eastern Banks is largely ocean-influenced and well

flushed by ocean waters entering the banks from various channels, while the western part of

Moreton Bay is dominated by terrestrial inputs from rivers (Figure 2.2.). This geographical setting

determines different sediment characteristics in the area where quartz sand predominates in the

Eastern Banks and mud was the dominant substrate in the western region (Preen, 1992). The tidal

type in Moreton Bay is semi-diurnal (Milford and Church, 1977) with low tidal range of ±1 m

(Lyons et al., 2011). This site was selected for the present study primarily due to data availability

from previous seagrass remote sensing studies in the area (e.g., Phinn and Dekker, 2004; Phinn et

al., 2008; Roelfsema, 2009, 2014; Lyons et al., 2011, 2012) and easy accessibility for field work.

Page 38: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

18

Figure 2-3. Moreton Bay, East Coast Australia (ALOS AVNIR-2 14th July 2010) showing the boundary of the Eastern Banks, the main study area, in the white rectangle.

2.3. Image Data

Seasonal MERIS level 2 data from 2003–2012 with acceptable cloud cover were

downloaded from ocean colour website data (oceancolor.gsfc.nasa.gov), resulting in a total of

38 scenes at approximately 3 month intervals. HICO Level 1B images of 2013 with seasonal

coverage (4 scenes) were downloaded from the same website and processed using Seadas software

to produce semi-analytical Kd490 and KdPAR parameters (Lee et al., 2005a, 2005b). A season is

defined by the Australian seasons of summer (December-February), autumn (March-May), winter

(June-August) and spring (September-November). The general characteristics of the satellite

imagery used are given in Table 2.1. The acquisition dates and the detail of the pre-processing steps

are further explained in succeeding chapters.

Page 39: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

19

Table 2-1. General Characteristics of Satellite Imageries Used in this Study

MERIS seasonal average data were used for estimating PAR, for example summer PAR data

were estimated from the average of summer months from 2003–2012, etc. Ideally, PAR data were

obtained from HICO images, which would enable concurrent coverage of both PAR and Kd490 or

KdPAR. However, as HICO was launched in 2009 and many of its applications are still under

development the algorithm to extract PAR from HICO has not been developed yet. Depth data were

provided from WorldView2 satellite-based bathymetry (acquisition date 12 June, 2012) available at

the database of the Biophysical Remote Sensing Group, the University of Queensland, processed

using the Lyzenga method (Lyzenga, 1978; Lyons et al., 2011). The original Worldview-2 data of

12 June, 2012 (not processed to bathymetry) and of 13 February, 2013 were used for LAI and

biomass mapping for chapter 4.

2.4. Field Data Collection

There were three main activities conducted during fieldwork in February 2013: seagrass

sample collection (for LAI and biomass), water sampling and optical profiling, and in situ spectral

measurement. Biomass core sampling included seagrass leaf collection for the measurement of

optical properties. In situ spectral measurements were performed over points where seagrass

biomass were sampled. A more detailed procedure of water sampling and optical profiling is

described in chapter 3. Details of in the situ reflectance measurements and seagrass core collection

procedure, including sample laboratory treatment to obtain LAI and biomass, are explained in

chapter 4.

Some of the input data also comprised of in situ optical data collected in Moreton Bay by

past collaborative research projects in 2001 (Phinn and Dekker, 2004) and were primarily used for

algorithm testing in chapter 3. The data of (Phinn and Dekker, 2004) used were Kd490 and KdPAR.

The detail of the data location and processing steps is further explained in chapter 3.

AttributesMeris Hico Wv-2

Time coverage/ acquisition 2003-2012 (10 years); Seasonal 2013 (1 year); Seasonal 12-Jun-12(Summer, Autumn, Winter, Spring) (Summer, Autumn, Winter, Spring)

Spatial resolution 300 m 90 m 2 mNumber of Bands 15 87 (only 15 used in this study) 8Spectral resolution (band width) 1.8 nm 5.7 nm 50 -60 nmSpectral Range 390 - 1040 nm 400 - 900 nm 400 - 1040 nmNumber of scene (s) Used in this study 38 4 1Parameter/product used PAR, in Einstein / m-2 / day Kd490 Lee (Lee algorithm), m-1 Bathymetry (Lyzenga method),m

KdPAR Lee (Lee algorithm), m-1

Satellite

Page 40: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

20

Chapter 3: Characterising the Underwater Light Environment in Seagrass Habitat Using

Field Survey and Remote Sensing Approaches

Main Findings

• The light climate in seagrass in the study site was dominated by green light and based on its

light diffuse attenuation characteristic falls into the first coastal type of Jerlov's optical

classification that typically shows increasing attenuations at blue wavelengths due to colour

and suspended matter and at red wavelength due to water absorption.

• Seagrass distribution across one of the study sites (Wanga Wallen Banks) indicated it may be

blue light-limited due to the observed rapid decrease in blue light contributions relative to

green and red light, moving from the more-dense inshore seagrass site to the less-dense

offshore seagrass site.

• Light appeared to be one of the main controlling factors that determined the spatial distribution

of seagrass in the Eastern Banks, as indicated by similarities between seagrass surface area

determined from light parameters of photosynthetically active radiation (PAR),

photosynthetically utilised radiation (PUR) and percentage of light relative to surface light

(% SI) and published seagrass maps.

• Agreement between satellite-based estimates of PAR and % SI and their corresponding in situ

data, suggested the acceptability of the satellite-based method used in this study for

investigating the seagrass light climate in the study site.

3.1. Introduction

Available sunlight is an essential factor controlling seagrass distribution, productivity and

biomass (Preen, 1992; Denison et al., 1993; Zimmerman, 2006; Bjork et al., 2008). It is further

recognised that seagrasses have high light requirements, ranging from 10–37% of in-water at-

surface light, partly due to their inefficient carbon concentrating mechanism for photosynthesis

(Zimmerman, 2006). The importance of light in affecting the spatial distribution of seagrass

increases for seagrasses inhabiting coastal waters with varying water clarity due to sediment re-

suspension and river run-off (Preen, 1992; Longstaff, 2003; O’Brien et al., 2011).

Photosynthetically available radiation (PAR), the diffuse attenuation coefficient of PAR

(KdPAR) and photosynthetically usable radiation (PUR) are important parameters for understanding

and investigating the underwater light environment of benthic habitats, particularly for assessing the

Page 41: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

21

suitability of water bodies for supporting seagrasses (e.g., Gallegos, 1994; Dennison et al., 1993;

Anastasiou, 2009). For scientific use, it is also common to express PAR, PUR and Kd in spectral

notation, (λ), referring to a specific, or a range of wavelength. One spectral product of Kd (λ),

Kd490, can be used to characterise water clarity and light conditions at the substrate. It can also be

used to estimate photosynthetically available radiation (PAR) once the relationship between Kd490

and KdPAR is established (Saulquin et al., 2013). Additionally, the portion of sunlight used by

seagrass leaves for photosynthetic processes (PUR) can be calculated when PAR and the optical

properties data of the leaves are available (Zimmermann, 2003; Anastasiou, 2009). Both PAR and

PUR are important parameters for modelling seagrass photosynthesis (Zimmerman, 2003;

McPherson et al., 2011). In the context of seagrass conservation, maps of benthic underwater light

environments can be used to support light-based seagrass management approaches that are currently

under development in some Australian waters (Chartrand et al., 2012).

Kd (λ) is categorised as one of the apparent optical properties (AOPs), influenced by both

inherent optical properties (IOPs) such as absorption and scattering and the radiant field (e.g. sun

zenith angle) (Preisendorfer, 1976). Kirk (2010) used the term “quasi-inherent optical properties”

for Kd (λ) to emphasise its slight dependent on light geometry (e.g. sun angle or depth) and the

importance and close relationship between AOPs and IOPs. Kd (λ) is therefore a good parameter to

characterise different water masses or water optical classification (Jerlov, 1976).

The routine mapping of light and water quality parameters in the world’s oceans is

conducted using various satellite sensors, such as SeaWIFS, MODIS, and MERIS. The data are

used to calculate parameters including Kd490, KdPAR, and PAR. Three standard operational

algorithms are commonly used for the calculation of Kd490 and KdPAR, using remote sensing

reflectance (Rrs) of satellite ocean colour data (Lee et al., 2005a, 2005b; Saulquin et al., 2013, Zhao

et al., 2013). Algorithm 1 uses the blue-green ratio of Rrs to calculate Kd490 and KdPAR (Mueller,

2000). In algorithm 2, the chlorophyll concentration correlated to Kd490 and KdPAR (Morel, et al.,

2007). The last algorithm uses a semi-analytical approach in which Rrs is decomposed into water

column absorption and scattering coefficients, from which Kd490 and KdPAR are calculated (Lee et

al., 2005a, 2005b). Only the third algorithm (semi-empirical) has robust applications in both

oceanic and coastal waters, while the other two perform well only for oceanic waters (Saulquin et

al., 2013;. Zhao et al, 2013). While KdPAR was originally developed as a function of chlorophyll

concentration, recent development shows that it was derived more from the empirical relationship

Page 42: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

22

between in situ Kd490–KdPAR, which resulted in regional variations for the relationship (Morel et al.,

2007; Pierson et al., 2008; Wang et al., 2009).

To date, only Gattuso et al. (2006) have used a satellite-based light map to study the spatial

distribution of benthic habitats including seagrass. Their pioneering study was conducted at global

scale using 9 km resolution of SeaWIFS satellite data and a PAR parameter calculated from a

chlorophyll-based algorithm. Several studies have confirmed that the chlorophyll-based algorithm is

reliable only in case 1 water (oceanic water) where its optical properties are primarily determined

by phytoplankton (Lee et al., 2005a, 2005b; Saulquin et al., 2013; Zhao et al., 2013). Additionally,

as mentioned in their recommendation section, the study of Gattuso et al. (2006) did not include

field validation for their satellite-based light maps and seasonal analysis. As seagrass commonly

inhabits coastal areas where its optical properties are not only chlorophyll-dependent, further

studies are thus required for investigating spatially-complex, narrow and highly dynamic coastal

environments using a more robust algorithm (e.g., semi-analytical algorithm) and higher satellite

resolution.

For this study, the light spectral quality in seagrass-dominated areas of the Eastern Banks,

Moreton Bay, Australia was investigated using in situ data from field surveys through analysis of

Kd (λ) and PAR (λ). Next, remote sensing analysis was implemented to map light quantity

parameters in terms of PAR (λ), percentage of surface irradiance (% SI) and PUR (λ) for the study

site. This mapping section used MERIS satellite-based PAR, HICO satellite-based KdPAR,

WorldView2 satellite-based bathymetry and laboratory-based leaf absorptance (AL) data to

calculate and map PAR, % SI and PUR at the depth of seagrass. The semi-analytical KdPAR

algorithm of Lee et al. (2005a, 2005b) was used in this study. The final PAR, % SI and PUR maps

were used to estimate seagrass surface area in the study site. A published seagrass map of the study

site (Roelfsema et al., 2014) was used as a reference for the correct extent of seagrass and was

compared with the seagrass surface area determined by the light parameters. Finally, the validity of

satellite-based PAR was tested against in situ PAR data and the performance of the semi-analytical

KdPAR algorithm was compared with the KdPAR algorithm empirically developed from in situ data.

Page 43: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

23

3.2. Data and Methodology

3.2.1. Field data collection and processing of Kd (λ) and PAR (λ).

Total downwelling irradiance measurements Ed(λ) were collected at two sites at Wanga

Wallen Banks (Figure 3.1b.) on 11 and 12 February, 2013 using an upward-looking Trios Ramses

radiometer (Figure 3.2.). Point 1 (average depth 4 m) was located several meters beyond the outer

limit of the seagrass meadow while point 2 (average depth 1.9 m) was measured the next day closer

to the seagrass meadows of the Wanga Wallen Banks (Figure 3.1b.). The substrate at both points

was seagrass and sand, with less seagrass at point 1 than at point 2. The description of the relative

position of the sampling points to the seagrass meadow was obtained from either visual inspection

or the published seagrass maps (e.g., Phinn et al., 2008; Lyons et al., 2011; Roelfsema et al., 2014;

Figure 3.1c.). The radiometer instrument has a wavelength range of 320–950 nm with around 3 nm

spectral resolution. To avoid boat shadow the radiometer was attached to a deployment pole and

was positioned away from the boat and on the sunny side (Mishra et al., 2005; Veal et al., 2009)

(Figure 3.2.).

Page 44: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

24

(a) (b)

(c) Figure 3-1. Details of the study area and sampling sites. (a) Moreton Bay, East Coast Australia (ALOS AVNIR-2 14th July, 2010) with the boundary of Eastern Banks, the main study area, in the white rectangle. The solid green circles are sampling points of the diffuse attenuation coefficient of Phinn and Dekker (2004) used in section 3.3.7. for testing the local Kd490–KdPAR algorithm developed in this study. The red line is a transect traversing several parts of the study area used to test the seasonal robustness of the local Kd490–KdPAR algorithm. (b) Extent of the Eastern Banks seagrass in June 2012 (white rectangle in (a)) analysed from the database of the Biophysical Remote Sensing Group of The University Queensland, which was also published in Roelfsema et al. (2014). (c) The detailed area of the yellow rectangle in (a) and (b) where in situ water optical measurements were conducted in February 2013 (solid yellow rectangles in (c)).

Page 45: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

25

Figure 3-2. Measurement of downwelling irradiance using the profiling radiometer aboard the R/V Scarus. Profiling was conducted by lowering the radiometer at several depths from which the diffuse attenuation coefficient was then calculated. The measurement was conducted at Wanga Wallen banks (Figure 3.1 c).

The irradiance profiling involved lowering the radiometer through the water column at depth

intervals of approximately 0.5 m. The profiling was repeated every 30–60 min, with measurements

conducted from 10am–3.55pm and 8am–2.21pm for point 1 and 2, respectively.

Kd was calculated using the exponential light attenuation principal of Beer’s law :

𝐸𝑑𝑧2,𝜆 = 𝐸𝑑𝑧1,𝜆 × 𝑒(−𝐾𝑑𝜆(𝑧1−𝑧2)) (1)

where Edz1,λ and Edz2,λ are the downwelling irradiance at depth z1 and z2, respectively. In

principle, Kdλ was obtained from the exponential fitting of the irradiance profile over depth

(Bracchini et al., 2007; Veal et al., 2009). Kd was used only when the coefficient of determination

(R2) of the fitting between depths and irradiance was > 0.7 (Bracchini et al., 2007). When only two

depths were measured due to shallow water, Kd was calculated using (Palandro, 2006):

𝐾𝑑𝜆 = ln(𝐸𝑑𝑧1,𝜆/𝐸𝑑𝑧2,𝜆)2 (𝑍2−𝑍1)

(2)

Kd (λ) was then plotted and descriptively analysed in term of its difference in magnitude

and spectral shape and possible connections with seagrass light requirements.

PAR is a measurrement of incident light over the wavelength range between 400–700 nm

and was calculated using formula :

𝑃𝐴𝑅 (𝑧) = ∫ 𝐸𝑑(𝜆, 𝑧) 𝑑700 𝑛𝑚400 𝑛𝑚 𝜆 (3)

Page 46: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

26

To find out the impact of Kd (λ) on the light spectral composition at depth of seagrass, PAR

at blue (400–900 nm), green (491–600 nm) and red (640–700) wavelengths were calculated using

equation 3. This calculation was applied only to the deepest radiometer measurements at the

two points.

3.2.2. Satellite image acquisition, image processing and seagrass surface area

calculation.

The light parameters of interest in this mapping section are PAR, percentage ratio of bottom

PAR-to-surface PAR (% SI = Surface Intensity) and PUR at seagrass depth, which can be

calculated if surface PAR, KdPAR, bathymetry and the leaf absorptance coefficient (AL) are

available. PAR at seagrass depth was calculated using MERIS-based surface PAR (300 m

resolution) and HICO-based KdPAR (90 m resolution) downloaded from satellite ocean colour

website (http://oceancolour.gsfc.nasa.gov/), and WorldView-2-based bathymetry (2 m resolution)

available at the Biophysical Remote Sensing Group, at The University of Queensland. (See Table

3.1 for the acquisition dates of the satellites used for generating PAR and PUR maps).

Table 3-1. Acquisition Dates of Satellite Data Used to Derive PAR and PUR Maps

It would be ideal if all required satellite-based parameters were available at the same

optimum spatial resolution, in this case a 2 m resolution of WorldView-2 image. The different

forms of satellite data constrain this. First, the water optical property parameters (e.g., KdPAR) were

only available from ocean colour satellite data (e.g., MODIS, SeaWIFS, MERIS, HICO) among

which HICO has the highest spatial resolution (90 m). HICO is a new hyperspectral ocean colour

satellite launched in 2009. The use of HICO data for various applications is still in progress and this

study is one of the first applications of HICO data for the study of seagrass. Second, several

Satellite Summer Autumn Winter Spring22-Dec-03 8-Mar-03 28-Jul-03 14-Nov-0320-Feb-04 1-May-04 8-Aug-04 29-Sep-0429-Dec-05 18-Apr-05 26-Jul-05 18-Nov-05

MERIS 17-Jan-06 10-Apr-06 1-Jul-06 5-Oct-0615-Jan-07 17-Apr-07 2-Jul-07 25-Oct-07

7-Feb-08 27-Apr-08 9-Jul-08 19-Oct-084-Feb-09 28-Apr-09 3-Jul-09 7-Oct-094-Jan-10 4-May-10 2-Aug-10 6-Sep-10

26-Feb-11 13-Apr-11 19-Aug-11 21-Sep-11HICO 3-Feb-13 21-Apr-13 30-Jul-13 6-Oct-13WV-2 12-Jun-13

Season

Page 47: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

27

attempts to extract PAR from HICO data using SeaDAS software resulted in error messages stating

that the PAR extraction algorithm (Frouin et al., 1989) has not been extended yet for HICO data.

Since MODIS and SeaWIFS data produce a coarse resolution of PAR (500 m and 1 km for MODIS;

4 and 9 km for SeaWIFS), PAR was then extracted from MERIS data at 300 m resolution.

Unfortunately, MERIS data were only available until 2012 preventing concurrent data acquisition

with the HICO data of 2013. Use of a seasonal 10-year average of MERIS PAR data as input for the

PAR parameter (Table 3.1) addressed the shortfall. The term “seasonal” in this study referred

to summer (December-February), autumn (March-May), winter (June-August) and spring

(September-November).

Figure 3-3. Illustration of satellite-based parameters with different pixel sizes used in this study.

In order to conduct an image analysis, all satellite data are required to be at the same spatial

resolution, so all MERIS PAR and HICO KdPAR data were sub-sampled to 2 m resolution of

WorldView-2 data. This rescaling did not add information detail to the resized data but was

intended as a processing requirement only, so it was assumed that reflectance was constant across

each pixel and subdivided each one. In this study, irradiance (or PAR here) and KdPAR were

therefore assumed homogeneous within 300 m (MERIS resolution) and 90 m (HICO resolution),

respectively, while bathymetry varied over shorter distances (2 m resolution of WorldView-2)

(Figure 3.3.). In regards to the optimum spatial resolution for mapping seagrass species in the study

site, Phinn et al. (2008) suggested satellite data with a pixel size of < 10-15 m for the patchy and

MERIS PAR 300 m

HICO KdPAR

300

m

90 m

WorldView-2

Bathymetry 2 x 2 m

90 m

Page 48: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

28

heterogeneous seagrass that is characteristic of the Eastern Banks. Data at this scale did not exist to

estimate water optical properties, including KdPAR. Holden (2002) questioned the horizontal

homogeneity assumption required for water column correction but her paper did not clearly mention

the metric distance or optimum pixel size for which the assumption remained valid. Even if the

study of Holden (2002) resulted in optimum pixel size for the horizontal homogeneity assumption,

this might not be applicable to the study site in the Eastern Banks (coastal waters) as her study was

conducted in the clear coral reef waters of Bunaken Island, in Indonesia. In regard to PAR, previous

studies have attempted to assess the spatial variability of PAR but they were conducted either at a

global scale (e.g., Bishop, 1991) or using data from meteorological stations separated by > 1 km

distance (e.g., Bardosa et al., 2013; Rihiimaki and Long, 2014). While the possible impact of the

homogeneity assumption for PAR and KdPAR at their respective data resolution on the calculation is

acknowledged, its quantitative assessment is therefore unpredictable.

The MERIS PAR product is directly downloadable from the ocean colour website

(oceancolor.gsfc.nasa.gov) available at level 2, while KdPAR was generated from HICO image data

using Seadas software. MERIS PAR product was generated using the formula described in Frouin

et al. (1989) and KdPAR was produced by Seadas software using the quasi-analytical algorithm

(QAA) of Lee et al. (2005a, 2005b) embedded in the software. While the validity of chlorophyll-

based KdPAR was limited to oceanic waters, the use of KdPAR using Lee's method (Lee et al., 2005a,

2005b) was justified by previous studies indicating its application is robust for both oceanic and

coastal waters (Lee et al., 2005a, 2005b; Saulquin et al., 2013; Zhao et al., 2013). The available

WorldView-2-based bathymetry was calculated using the empirical Lyzenga method as described in

Lyons et al. (2011).

The PAR map was generated by applying the light exponential extinction equation to the

satellite-based surface PAR data using formula :

𝑃𝐴𝑅𝑏 (𝑧) = PARs. exp−kdPAR . z (4)

where PARs, KdPAR and z are surface PAR (PARs), diffused attenuation coefficient at PAR

wavelength and depth, respectively.

Percentage of PAR at seagrass depth (PARb) relative to surface PAR (PARs), termed as %

SI (percentage of Surface Irradiance), was then calculated as the percentage ratio between at-

seagrass depth and surface PAR using simple formula :

Page 49: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

29

%𝑆𝐼 = (PARbPARs� ) × 100% (5)

PUR was calculated by multiplying bottom PAR by seagrass leaf absorptance coefficient

(AL) as expressed in the following formula :

𝑃𝑈𝑅 = PARb × 𝐴𝐿(𝑃𝐴𝑅) (6)

where 𝐴𝐿(𝑃𝐴𝑅) is seagrass leaf absorptance at PAR wavelength (see texts below).

Seagrass leaves of five species (Zostera muelleri, Cymodacea serrulata, Halophila ovalis,

Halophila spinulosa and Syringodium isoetifolium) from Wanga Wallen Banks were collected

during the February 2013 field campaign. Leaf optical properties of transmission and

reflectance were measured in a laboratory using an ASD® integrating sphere equipped with ASD®

spectrometer (Figure 3.4.).

Figure 3-4. Equipment setting for measuring seagrass leaf optical properties. a = ASD® spectrometer; b = ASD® integrating sphere; c = ASD® internal light source.

AL was calculated using the following formulas (Durako, 2007):

𝐴𝐿(𝑁𝑃) = [1 − 𝑇𝐿(750)] − 𝑅𝐿(750) (7)

𝐴𝐿(𝜆) = [1 − 𝑇𝐿(𝜆)] − 𝑅𝐿(𝜆) − 𝐴𝐿(𝑁𝑃) (8)

Page 50: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

30

𝐴𝐿(𝑃𝐴𝑅) = 𝑎𝑣𝑒𝑟𝑎𝑔𝑒 [𝐴𝐿(400) − 𝐴𝐿(700)] (9)

where AL(NP) = non-photosynthetic absorptance (Zimmerman, 2003); TL(λ) = spectral

transmittance and RL(λ) = spectral reflectance, numbers = wavelength position.

Finally, a clustering procedure using the density slice technique in ENVI® software was

applied to the resultant PARb, % SI and PUR maps to estimate potential seagrass surface area in the

study site based on literature values of PAR and % SI for seagrass. The PAR literature values were

mainly from Gatusso et al. (2006) who compiled minimum light requirement for various seagrass

species (Appendix C of Gatusso et al., 2006) and from Longstaff (2003) for the dominant species

(Zostera muelleri) in Eastern Banks. The literature values of % SI were from various publications

for various species (e.g Longstaff., 2003; Gatusso et al., 2006; Zimmerman., 2006; Erftemeijer and

Robin Lewis III., 2006; Collier and Waycott., 2009; Bach et al., 1998; Grice et al., 1996). These

literature values were assumed to be transferrable to seagrasses examined in Moreton Bay.

Published seagrass maps of the study area were used as reference and comparison (e.g., Roelfsema

et al., 2014) and was depicted in Figure 3.1b and 3.10e.

3.2.3. Comparison with in situ data and empirical algorithm.

Although a rigorous accuracy assessment was not possible due to the limited number of in

situ measurement points (two sites), it would be beneficial to compare selected satellite-based

products and their corresponding in situ values. For this purpose satellite-based bottom PAR, KdPAR

and % SI were compared with in situ data obtained from radiometer measurements. Satellite-based

PAR in units of mol photons · m-2 · day-1 was converted into radiometer unit of w/m2 using

conversion factor (6.02e23 quanta/mol photon)/(86,400 sec/day x 2.77e18 quanta · s · W)

(http://www.mbari.org/bog/NOPP/par.html; Morel and Smith, 1974). In situ % SI was calculated by

the ratio of PAR at bottom profile to the above-water PAR measurement. To match the daily

product of the satellite parameters all radiometer data parameters, which were measured over the

course of the day in the time interval of 30 min–1 hr, were averaged.

3.2.3.1. Estimating KdPAR from Kd490 –KdPAR empirical model.

The relationship between Kd490 and KdPAR was explored using in situ data by applying

empirical model of Austin and Petzold (1986) (eq. 7) to reconstruct Kd (λ) in the range of PAR

spectrum (400-700 nm). The model uses value of Kd490 and two constants as the main inputs to

calculate Kd (λ) in the PAR spectrum at 5 nm interval. If the empirical model can be used to

reconstruct Kd (λ) in the PAR range, then KdPAR can be calculated by integration (Saulquin et al.,

2013). The expression of the empirical model is :

Page 51: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

31

𝐾𝑑(𝜆) = 𝐼(𝜆) + 𝑀(𝜆) ∗ 𝐾𝑑(490) (10)

where I(λ) and M(λ) are coefficients in Austin and Petzold (1986) and Kd (490) is from in

situ data.

3.2.3.2. Estimating KdPAR from local Kd490–KdPAR algorithm.

The local Kd490–KdPAR algorithm was developed using data from in situ radiometer profile

measurement. For this purpose in situ downwelling irradiance at PAR range, EdPAR, was calculated

by integrating over the wavelengths at 400 - 700 nm for each radiometer profile measurement from

which KdPAR was then calculated using equation 1. Kd490–KdPAR algorithm was produced by

implementing regression analysis between Kd490 and KdPAR for each radiometer profile

measurement. Despite its empirical nature, the implementation of local Kd490–KdPAR algorithm has

been applied to different regions with varying water optical properties, including clear and turbid

waters (Morel et al., 2007; Pierson et al., 2008; Wang et al., 2009, Saulquin et al, 2013).

The performance of the local Kd490–KdPAR algorithm was tested against the performance of

the following local Kd490–KdPAR algorithms from literature using the independent in situ Kd490 data

set of Phinn and Dekker (2004) (see Figure 3.1a.) :

𝐾𝑑𝑃𝐴𝑅 = 0.0665 + 0.874.𝐾𝑑490 − 0.00121/𝐾𝑑490 (Morel et al., 2007) (11)

𝐾𝑑𝑃𝐴𝑅 = 0.6677.𝐾𝑑4900.6763 (Pierson et al., 2008) (12)

𝐾𝑑𝑃𝐴𝑅 = 0.8045.𝐾𝑑4900..9170 (Wang and Son., 2009) (13)

𝐾𝑑𝑃𝐴𝑅 = 4.6051. 𝑘𝑑490/(6.0700.𝐾𝑑490 + 3.200), for Kd490 <= 0.115 m-1 (Saulquin

et al., 2013) (14a)

𝐾𝑑𝑃𝐴𝑅 = 0.8100.𝐾𝑑4900.8256, for Kd490 > 0.115 m-1 (Saulquin et al., 2013) (14b)

3.3. Results and Discussion

3.3.1. In situ measurement of Kd (λ), PAR and PAR blue, green, red.

Kd (λ) was obtained from exponential plotting between depth and irradiance data of the

radiometer equipment. The coefficient of determination (R2) of this exponential plotting is shown

in Figure 3.5.

Page 52: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

32

Figure 3-5. Regression coefficient (R2) from exponential fitting between irradiance profile and depth from which Kd was obtained. Figure (a) is for point 1 and figure (b) is for point 2.

.

The average R2 of measurements from both day 1 and day 2 shows a similar pattern in

which correlation coefficients between downwelling irradiance and depths are high at both ends of

the light spectrum (e.g., blue and red) and decreased towards the mid-range of the spectrum (e.g.,

green). When linear regression (instead of exponential fitting) was applied to the green spectrum the

correlation coefficients increased indicating the penetration of green spectrum is higher (less

exponential) than blue and red spectrum. On average, the R2 of point 1, calculated from more depth

profiles, are higher than that of day 2 that were processed using fewer depth profiles. The depth

varied from approximately 1.7–6 m and 1.1–2.4 m for point 1 and point 2, respectively.

Figure 3.6. shows the result of Kd (λ) calculation for point 1 and 2.

Page 53: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

33

Figure 3-6. Kd (λ) profiles for (a) point 1 and (b) point 2. The blue lines and arrow indicate Kd (λ) profiles over the course of the day. T symbol indicates time of measurement. The average of Kd (λ) profiles for both points (red lines in (a) and (b)) were plotted together in (c) for ease of comparison.

Page 54: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

34

In general, there was increasing light attenuation at both blue and red wavelengths at both

study sites (Figure 3.6c.). The green spectrum at both sites showed lower attenuation than the blue

and red wavalengths. The change in the overall Kd (λ) magnitude over time could be due to the

changing sun zenith angle during the course of the day. The magnitude of Kd (λ), an apparent

water optical property, depends on inherent optical properties (e.g., absorption and scattering

coefficient) and sun zenith angle (Lee et al., 2005a, 2005b; Lee, 2009). Comparing the two sites,

light attenuation at wavelengths longer than ± 460 nm at point 1 is lower than the same spectrum at

point 2, causing the overall lower KdPAR value of point 1, as confirmed by both in situ and satellite

KdPAR data. To provide the analysis with light quality information and to account for the impact of

the Kd (λ) profiles on the underwater light climate, bottom PAR at both sites were expressed as

blue, green and red spectrum (Figure 3.7.).

Page 55: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

35

Figure 3-7. Bottom PAR expressed as blue, green and red wavelengths calculated from the daily average of radiometer measurements at seagrass depth (the deepest radiometer profiling).

It can be shown from Figure 3.7. that the light spectral proportion at each site followed its

corresponding Kd (λ) profile in Figure 3.6c. For both points, green light was the most abundant as

Kds at this wavelength were the lowest, meaning waters at both sites were the most transparant to

green light. At point 1 the proportion of blue (4.94 w/m2) and red (4.96 w/m2) lights energy was

almost the same, as Kds at these wavelenghths were very similar. The lower Kd at the blue

wavelength, than it was at the red wavelength, caused more blue light than red light at point 2.

Comparing the light quality between the two sites, one noticeable feature was the higher blue light

Page 56: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

36

attenuation at point 1 than at point 2 (Figure 3.6c.). Combined with the depth factor, this caused the

greatest reduction in blue light from point 2 to point 1 up to 201% while the reduction in green and

red lights were 65% and 125%, respectively. Kirk (2010) used the term quasi-inherent optical for

defining Kd to emphasize that the parameter reflects inherent optical properties (IOPs) and less

dependent on the radiation field (light geometry), e.g. sun angle or in this case, depth. As IOPs is 'a

finger print' of a water mass, this would mean that the spectral characteristic of Kd could be used to

characterise different water masses. Using the system devised by Jerlov (1976) that used Kd (λ)

values to optically classify waters into ocean types (I, IA, IB, II and III) and coastal types (1, 3, 5, 7

and 9), the two study sites belonged to coastal type 1. Gallegos and Kenworthy (1996) stated that

typical coastal waters showed increasing attenuations at blue wavelength, due to colour and

suspended matter and at red wavelength, due to water absorption. As this colour and suspended

matter comes mostly from land origins (e.g., Anastasiou, 2009), it is tempting to attribute the

reduction in the blue light at point 1 to land-based materials. However, point 1 is located at the

seaward edge of the seagass meadow of Wanga Wallen Bank, close to the deep Rainbow Channel,

while point 2 is located closer to the land where seagrass is more abundant (Figure 3.1b.). Phinn and

Dekker (2004) studied both inherent and apparent water optical properties of the Moreton Bay area

and optically classified the waters of the area into four classes: 1) near Case 1 oceanic waters, 2)

complex green brown waters (Case 2), 3) complex near coastal waters and 4) turbid – Case 2 waters

of riverine influence. The near Case 1 oceanic waters type was characterised by high absorption of

phytoplankton (Phinn and Dekker, 2004) and its sampling position (Amity Jetty) was ±3 km to the

north of the two study sites. Figure 3.8. depicts the position of Amity Jetty and the two study sites

plotted on the KdPAR image of HICO data acquired on 3 February 2013. From the optical

categorisation of Phinn and Dekker (2004) and the HICO KdPAR data it is suspected there is a north-

to-south “intrusion” of this near Case 1 oceanic water passing through the deep Rainbow channel, to

the Wanga Wallen Banks area. On the HICO KdPAR image this near Case 1 water is characterised

by low KdPAR values (red color in Figure 3.8.).

Page 57: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

37

Figure 3-8. Sampling locations of diffuse attenuation of Phinn and Dekker (2004) at Amity point (black solid triangle), and of this study (black solid cross) at Wanga Wallen banks, plotted on KdPAR image of HICO data acquired on 3 February, 2013 (left), and on the natural composite of ALOS AVNIR-2 14 July 2010 (right).

Therefore, while the overall higher KdPAR value of point 2 may be due to the influence of

terrestrial inputs, the higher light attenuation at the blue wavelength of point 1 was suspected due to

phytoplankton brought by the near Case 1 waters whose typical absorption is greatest in the blue

spectrum (Figure 3.9.; Bricaud et al., 1983). The HICO KdPAR image also revealed the different

characteristic of KdPAR between Rainbow Channel (lower KdPAR) and Wanga Wallen Banks (higher

KdPAR) which was reflected in the in situ data (e.g., Figure 3.6c.).

KdPAR (m-1)

Page 58: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

38

Figure 3-9. Spectral absorptance of phytoplankton chlorophyll as measured spectrometrically using the method of Parsons et al. (1984). The water samples were collected in Wanga Wallen Banks during fieldwork in February 2013, at the time interval of 30–60 min (T1–T5).

For photosynthetic processes seagrass possesses light-harvesting pigments in the form of

chlorophyll a, b, c and the carotenoids, of which chlorophyll a is the primary pigment absorbing

light at around 440 nm (blue wavelength region) and 660–680 nm (red wavelength region) ( Kirk,

2010). While green light was abundant in the study sites, seagrass is not equipped with a specific

pigment to harvest green light, though the presence of carotenoids may extend their light absorption

region beyond blue and red lights (Cummings and Zimmerman, 2003; Kirk, 2010). The average

bottom PAR as measured using radiometer were 24.27 and 50.08 w/m2 for point 1 and point 2,

respectively during the periods sampled. In terms of % SI, these values were equal to 13.50% and

30.60% for point 1 and point 2, respectively. Anastasiou (2009) studied light spectral quality at four

sites located along the seagrass deep edge of Tampa Bay, Florida. In ecological terms, the seagrass

deep edge is called ecological compensation depth (ECD), below which the carbon balance for

seagrass becomes negative and, hence, unsustainable (Gallegos and Kenworthy, 1996). Anastasiou

(2009) concluded that seagrass in his study area was blue-light limited, as all stations showed the

least amount of blue light while the % SI were still in the range for seagrass survival. According to

the published seagrass map (Figure 3.1b.; e.g., Roelfsema et al., 2014) point 1 of the study site here

is located in the seaward edge of seagrass meadow of Wanga Wallen Banks, close to the deep

Rainbow Channel. As the % SI values of the two sites still support seagrass survival (e.g.,

Erftemeijer and Robin Lewis III, 2006; Zimmerman, 2006; Papenbrock, 2012), the lower

abundance of seagrass at point 1 may be an indication of blue light dependent of seagrass.

Page 59: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

39

3.3.2. Potential seagrass surface area based on PAR, % SI and PUR maps.

The main products of this light mapping section were maps of PAR, bottom light relative to

surface light (% SI, surface irradiance) and the amount of light utilised by seagrass (PUR). The first

and second maps were then used to estimate potential seagrass area in the study site based on

literature values of PAR and % SI. The range of PAR for seagrass survival was often termed as the

minimum light requirement for seagrass (Emin). Gattuso et al.(2006) conducted a literature review

on the Emin of various benthic habitats, including seagrass. They found that the deepest depth of

seagrass colonisation was 90 m, corresponding to % SI of 11%, and the median of Emin across

different species was 5.1 mol photons · m-2 · day-1. In this study these values were used as

approximate starting points for estimating the potential seagrass surface area in the study site.

Page 60: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

40

Potential seagrass surface area

(a) (b)

(d) (c)

Page 61: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

41

Figure 3.10 a, b, c and d show the maps of potential seagrass surface area as calculated using

procedure in section 3.2.2. The difference in potential seagrass surface area among the maps in

Figure 3.10 a,b, c and d (area with the green color) is mainly due to the difference in cloud cover, or

areas in MERIS or HICO data marked as turbidity and/or atmospheric correction failure by SeaDAS

image processing software. In this context, the summer scene (a) is the clearest one among the other

scenes. The range of values used to generate the maps in Figure 3.10 a,b, c and d are given in Table

3.2. The main seagrass map reference used as comparison for Figure 3.10 a,b, c and d is the

seagrass map in Figure 3.10 e, which has been published in Roelfsema et al. (2014). Table 3-2. The Ranges of Values Used to Generate PAR, % SI and PUR Maps as Depicted in Figure 3.10 a, b, c and d.

SeasonMin Max Min Max Min Max

Summer 67.94 37.34Autumn 5 46.21 2.5 45 2 25.4Winter 33.75 18.55Spring 54.35 29.88

PAR (mol photons/m2/day) % SI PUR (mol photons/m2/day)

Figure 3-10. Seagrass surface area as determined based on PAR, PUR and % SI maps for (a) summer, (b) autumn, (c) winter, and (d) spring, plotted against ALOS-AVNIR-2 imagery of 14 July, 2010 as the background image. (e) is the reference seagrass map used in this study (Roelfsema et al., 2014)

(e)

Page 62: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

42

While seasonal PAR maps were generated with the same minimum value of

5 mol photons · m-2 · day-1, the maximum values varied seasonally following the maximum

seasonal PAR values of the MERIS PAR data. The minimum value of 5 mol photons · m-2 · day-1

also has another ecological meaning. Markager and Sand-Jensen (1994) introduced the term of

compensation irradiance for growth (Ec growth), another important light parameter that reflects

irradiance for a particular organism for which gross primary production and carbon losses (e.g.,

respiration and reproduction) are in a balanced state. Gatusso et al. (2006) further explained that

while Ec growth was experimentally obtained from long-term growth-irradiance experiments, it was

empirically the irradiance at the depth limit of benthic primary producers (such as seagrass in this

study). In addition to that, Ec growth is an integration of light requirements over a prolonged

timeframe, smoothing out seasonal variations in irradiance. From the literature study they found

Ec growth for seagrass was 5 mol photons · m-2 · day-1.

The second light parameter used to estimate seagrass surface area as depicted in Figure 3.10.

is % SI with a final range value from 2.5–45%. The seagrass area mapped using this % SI range

coincides well with the majority of the extent of seagrass reported in previous studies (Figure 3.1c.)

(Phinn et al., 2008; Roelfsema et al., 2009; Lyons et al., 2011, 2012; Roelfsema et al., 2014).

Zimmerman (2006) stated that seagrass had a light requirement of 10–37% of surface irradiance.

However, the range values of % SI of seagrass cannot be considered as exact limits, as the ranges

could vary among different species and areas. Previous studies on the light environments of

seagrass in the study site were limited and one of them was Longstaff (2003) who found that annual

percentages of surface light for Zostera capricorni (now Zostera muelleri) of South Moreton,

Pelican Banks, Waterloo Bay, Deception Bay and Bramble Bay (all within Moreton Bay), were 35,

31, 36, 16, and 6, respectively. Erftemeijer and Robin Lewis III (2006) compiled % SI of seagrass

and found a considerable range of values, from 2.5–37% of SI for the minimum light requirement of

seagrass across different species and geographic areas. Other studies stated that light requirements

of tropical seagrass genera such as Zostera and Cymodacea, which are both are present in the

Eastern Banks, were around 40% SI (Collier and Waycott 2009; Bach et al., 1998; Grice et al.,

1996). The upper limit of 45% found in this study was similar to Gatusso et al. (2006) using

SeaWiffs satellite data. While PAR gives the range value of seagrass light requirements in exact

irradiance units (e.g., mol photons · m-2 · day-1), biologists and conservationists often determine

light requirements using percentage of surface irradiance (% SI). This qualitative criteria is useful in

avoiding geographic differences in water properties or point-to-point variations at the same site. It is

also convenient for % SI criteria to be translated into seagrass management schemes. For example,

Page 63: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

43

seagrass management in Tampa Bay, Florida set the light minimum target as 20.5% while the target

for Indian River Lagoon, Florida was proposed as 24– 37% (Anastasiou, 2009).

PUR maps were generated by multiplying bottom PAR maps and seagrass leaf absorptance,

(ALPAR) using the range of PUR values as given in Table 3.2. It was originally planned to generate a

PUR map per species using the species maps based on WorldView-2 data analysis that were

available from the Biophysical Remote Sensing Group, at the University of Queensland. However,

the laboratory measurement using the integrating sphere revealed similarity in aborptance between

species (Cymodacea serrulata = 0.58, Halophila ovalis = 0.54, c = 0.52, average of 3 the three

species = 0.55, n spectra = 10; n leaf = 3). Absorptance for Zostera muelleri and Syringodium

isoetifolium were unsually high compared with literature values, reaching 0.71 and 1.03,

respectively (n spectra = 10; n leaf = 3). These high values were likely due to the small blades of the

two species that may have incompletely covered the integrating sphere’s measurement port, despite

well implemented stacking. Durako (2007) found that absorptances of nine seagrass species

collected from eastern (NSW) and western Australia showed low variability with an average value

of 0.57 ± 0.06, which is close to the average absorptance of the three species measured in this study

(0.55). Cummings and Zimmerman (2003) found PAR absorptance values of 0.48– 0.56 for Zostera

marina and Thalassia testudinum, and Enrique et al. (1994) obtained absorptance of 0.59 for 90 and

30 species of macroalga and marine angiospermae, respectively across wide variations in thickness,

chlorophyll a densities and chlorophyll-specific optical properties. Considering all these studies and

the result of the leaf absorptance coefficient (ALPAR) measurement in this study, the average

absorptance value of 0.55 was used to generate the PUR map.

Similar to other studies (e.g., Enrique et al., 1994; Cummings and Zimmerman, 2003;

Durako, 2007; Anastasiou, 2009), the seagrass leaf absorptance in this study showed high values at

blue and red wavelengths due to chlorophyll absorption (Kirk, 2010) and low values at green

wavelength (Figure 3.11.).

Page 64: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

44

Figure 3-11. Spectral absorptance of Si (Syringodium isoetifolium), Cs (Cymodacea serrulata), Zc (Zostera muelleri), Ho (Halophila ovalis ) and Hs (Halophila spinulosa ) collected from the Wanga Wallen Banks during the field campaign in February 2013 and measured using ASD® integrating sphere (Figure 3.4). Estimated carotenoids absorptance area in the green band (Kirk, 2010) is indicated.

It can also be seen that seagrass absorptance has the opposite spectral characteristic to the

available light spectral composition in the study site (see Figure 3.7.). This may be an indication of

the photo-adaptive capability of the chlorophyll pigments to low blue and red lights, or the presence

of carotenoid pigment which can extend its absorption range to green light spectrum (Kirk, 2010).

While PAR describes available light to seagrass expressed as equally weighted photons from

400–700 nm, PURλ is a more accurate parameter explaining selective spectral light absorption by

seagrass (Gallegos, 1994) (see Figure 3.11.). However, a PUR map at different wavelengths (e.g.,

blue, green and red) can be produced if only PAR maps at those wavelengths are available. To date

satellite-based PAR maps are only available from ocean colour satellite data with a broad band

range from 400–700 nm. The PUR maps generated in this study can be considered both as an

additional means to estimate potential seagrass areas in the Eastern Banks and as a quantitative

estimate of PAR absorbed by seagrass.

Finally, an analysis was conducted to reaffirm the relationship between light attenuation

(KdPAR) and seagrass distribution in the spatial context using the Eastern Banks case. Duarte (1991)

compiled literature data on seagrass depth distribution and light attenuation across different species

and geographic areas to establish the relationship between the maximum colonisation depth and

light attenuation. He found that the maximum colonisation depth was linearly related to the light

attenuation coefficient according to :

Page 65: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

45

Zc = 1.86/KdPAR (15)

where Zc = maximum depth of seagrass colonisation

Though similar equations have been produced for a particular species of seagrass (e.g., Dennison,

1987 and Nielsen et al., 1989 for Zostera marina in the northeast coast of the USA and Danish

estuaries, respectively) no attempt has been made to examine this sort of relationship using map

analysis. To do this the HICO KdPAR mean of each seasonal scene (Table 3.1.) was extracted and

then the resultant four KdPAR mean values of all four seasons were averaged. The

seasonally averaged KdPAR value of 0.56 was then used in equation 11, resulting in a maximum

colonisation depth of 3.31 m. Using the WorldView-2 satellite-based bathymetry, pixels with values

of 3.30–3.32 m were then identified and the result is presented in Figure 3.12.

Figure 3-12. Pixel values with the depth range of 3.30–3.32 m plotted on the seagrass map of the Eastern Banks (e.g., Figure 3.1c). The depth range value was determined from the relationship between KdPAR and the maximum depth of colonisation of Duarte Figure 3.12. Pixel values with the depth range of 3.30–3.32 m plotted on the seagrass map of the Eastern Banks (e.g., Figure 3.1c). The depth range value was determined from the relationship between KdPAR and the maximum depth of colonisation of Duarte (1991) (equation 12).

Using the published seagrass map (Figure 3.1b, 3.10e.) as the background map, it can be

seen that many pixels with depth values of 3.30–3.32 m reside at the outer boundary of the seagrass

area in the Eastern Banks. Longstaff (2003) found out using a survey staff and level method that the

maximum depth limit of Zostera capricorni in the study site, which is the dominant species in

Page 66: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

46

Eastern Banks, was close to 3 m. In addition to providing spatial analysis examination of Duarte’s

(1991) equation, this analysis can be considered as an alternative validation of the maximum extent

of seagrass in the Eastern Banks and a confirmation of the relationship between light and seagrass

spatial distribution.

3.3.3. Comparison with in situ data.

In general, the satellite-based bottom PAR and % SI were in good agreement with their

corresponding in situ data for the two sites (Table 3.3). For both parameters, point 1 showed a lower

value than point 2, primarily due to depth difference between the two points. In regard to KdPAR,

both satellite and in situ data showed increasing values from point 1 to point 2 indicating less

transparent water at point 2 than at point 1. Lee’s KdPAR (semi-analytical algorithm) predicted very

well the in situ KdPAR data for both points, while Morel’s KdPAR (chlorophyll-based algorithm)

underestimated in situ KdPAR at point 2. Previous studies confirm the robustness of the semi-

analytical algorithm for both case 1 (e.g., oceanic) and case 2 (e.g., coastal) waters and show that

chlorophyll-based algorithms tended to underestimate the in situ diffuse attenuation coefficient in

case 2 waters. Table 3-3. Comparison Between Satellite and In Situ Radiometer Data

Note. The in situ data were from daily averaged values of radiometer measurement, and the satellite KdPAR were extracted from HICO imagery acquired on 3 February 2013, two weeks earlier from the in situ data collection.

3.3.4. Comparison with Kd490–KdPAR empirical model.

The empirical model of Austin and Petzold (1986) was applied to in situ data to reconstruct

Kd (λ) in the range o f PAR at 5 nm intervals using the input value of Kd490 (diffuse attenuation

coefficient at wavelength 490 nm). The result shows that the model reconstructed Kd values at

wavelengths between 400– 600 nm very well but overestimated Kd beyond this wavelength range

for both sites.

SITESSatellite Satellite In situ

Satellite In situ Satellite In situ Kd PAR Morel Kd PAR LeePoint 1 (Depth ± 4.1 m) 25.542 24.277 14.805 13.5031 0.289 0.384 0.337Point 2 (Depth ± 1.9 m) 50.822 50.087 29.547 30.603 0.322 0.486 0.480

Parameters% SI Kd PAR (m-1)

(mol photons/m2/day)Bottom PAR

(Surface Irradiance)

Page 67: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

47

Figure 3-13. Calculation of Kd (λ) using empirical model of Austin and Petzold (1986) for (a) point 1 and (b) point 2.

The lower accuracies of the empirical model may have been related to the limited data used

in this work and may also have indicated that it has limited application in coastal environments, as it

was originally developed for case 1 water with Kd490 value limits of approximately 0.16 m-1 (Austin

and Petzold, 1986). The in situ Kd490 values of the study site were in the range of 0.1–0.5 m-1.

3.3.5. Comparison with local Kd490–KdPAR algorithm.

The local Kd490–KdPAR algorithm was developed by regressing in situ Kd490 against KdPAR

data from all radiometer profile measurements. The result showed that Kd490 potentially explained

90% of the pattern of KdPAR measured at the sampling sites (Figure 3.14.). As all the collected data

points were used to develop the KdPAR algorithm, the performance of the algorithm was examined

using independent data sets from Phinn and Dekker (2004) by comparing the result of KdPAR

calculated from the local algorithm and published KdPAR algorithms (equations 8–10b). The inputs

of this testing were in situ Kd490 at 10 points representing 10 locations in the Moreton Bay area, as

sampled by Phinn and Dekker (2004) (Figure 3.1a.).

Page 68: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

48

Figure 3-14. Local KdPAR algorithm developed by regressing in situ KdPAR against Kd490 data.

Figure 3-15. Performance of KdPAR algorithm of this study compared with published KdPAR algorithms (equation 8–11b). The input of the in situ Kd490 data are from Phinn and Dekker (2004) (solid green circles Figure 3.1 a).

The regression coefficients between KdPAR calculated using the local algorithm and those

calculated using published KdPAR algorithms showed values higher than 0.97 (Figure 3.15.),

meaning that if Kd490 data were fed into these algorithms (both local and published algorithms) the

resultant KdPAR values would be similar. The published algorithms used here cover a wide range of

Page 69: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

49

optical characteristics, for example, from clear oceanic waters (Morel et al., 2007) to very turbid

waters (Wang et al, 2009; Pierson et al., 2008). The result implies that although these algorithms

were developed regionally using empirical methods, they may be applicable to other regions.

Finally, the seasonal robustness of the local algorithm was examined using the HICO data

set for a selected transect across the Eastern Banks (red line Figure 3.1a.). For this purpose the

Kd490 values of the HICO data from the transect were extracted from four seasonal images

(summer, autumn, winter and spring; see table 3.1) and then converted into KdPAR using the local

algorithm. The Kd490 product used here was also processed using the semi-analytical algorithm of

Lee et.al. (2005a, 2005b). The calculated KdPAR was then regressed against the HICO KdPAR (using

Lee’s method) for each season (Figure 3.16.)

Figure 3-16. The comparison between calculated KdPAR using the local KdPAR algorithm and the original Lee's KdPAR for all seasons. The analysis was performed using HICO data from point transect traversing various area in the Eastern Banks (See Figure 3.1a).

Figure 3.16. shows a strong correlation between the calculated KdPAR and Lee’s KdPAR for

all seasons with coefficients of determination of 0.998, 0.998, 0.997, and 0.999 for summer,

autumn, winter and spring, respectively. The result confirmed the seasonal applicability of the local

KdPAR algorithm and also supported the results of previous studies on the robustness of the semi-

analytical algorithm of Lee et.al. (2005a, 2005b) for both case 1 (e.g., oceanic) and case 2 (e.g.,

coastal) waters (Saulquin et al., 2013; Zhao et al., 2013). The applicability, to a particular area, of

satellite-based parameters based on semi-analytical algorithms would bring convenience to studies

Page 70: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

50

requiring water optical parameters (e.g., KdPAR as in the present study), as they are directly

derivable from ocean colour satellite data using common image processing software (e.g. SeaDAS).

3.4. Conclusions, Perspectives and Future Work

This study has demonstrated that a combination of in situ and satellite data analysis can be

used to provide a local perspective (see Gattuso et al., 2006 for the global coverage) for the light

quality and the light limitations of the spatial distribution of seagrass in the Eastern Banks. In terms

of light quality, this study showed that the seagrass area in the study site was dominated by green

light and fell into the first coastal type category of Jerlov's optical classification (Jerlov, 1976).

Similar to the findings of Anastasiou (2009), the results here show that seagrass in the study site is

adapted to green light-dominated waters despite its common light photosynthetic absorption in the

blue and red wavelengths. Photosynthesis experiments using different light spectrum treatments on

seagrass, combined with the isolation of its light-harvesting pigments, would further explain this

photo-adaptive capability. In the context of the relevancy of water optical classification to remote

sensing applications, Mobley et al. (2003) questioned the relevance of the optical classification

based on the case of optically dynamic, marginal reef environments where the optical characteristics

could change from case 1 to case 2 water within a tidal course. However, the results here show that

optical classification is still relevant for areas with distinct optical characteristics, such as the

Eastern Banks (Phinn and Dekker, 2004), and also in the context of the ongoing development of

remote sensing algorithms that can be specifically applied to areas with certain types of optical

categories (e.g., case 1 or case 2 waters).

The result showed that available sunlight possibly determined the spatial distribution of

seagrass in the Eastern Banks, as certain ranges of values of PAR, % SI and PUR could be used to

cluster the majority of the seagrass surface area to match existing seagrass maps (e.g., Phinn et al.,

2008; Lyons et al., 2011, 2012; Roelfseme et al., 2014). More specifically, seagrass in the Eastern

Banks occupies an area with a minimum PAR value of 5 mol photons · m-2 · day-1, whereas the

maximum value varied seasonally. The value of 5.1 mol photons · m-2 · day-1 is the minimum light

requirement of seagrass across species and geographic areas found by Gatusso et al. (2006). The

maximum extent or outer boundary of the seagrass area in the Eastern Banks was well explained by

the parameter of diffuse light attenuation (KdPAR) through the relationship between KdPAR and the

maximum depth of seagrass colonisation of Duarte (1991).

In addition to providing explanatory parameters to assess seagrass spatial distribution,

results from this study demonstrate other possible uses of light as a monitoring and predictive tool

Page 71: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

51

for mapping and modelling seagrass habitat. For example, when PAR and or PUR ranges are

established for a particular seagrass area, any deviation from these values can be an indicator of

deterioration in the water quality in the seagrass area. Likewise, these established PAR and or PUR

ranges are useful for predicting potential areas for seagrass restoration.

This study showed good agreement between satellite-based PAR and % SI and their

corresponding in situ data. It also confirmed the applicability of the semi-analytical diffuse

attenuation (KdPAR) algorithm to the coastal waters of the Eastern Banks, as observed by previous

studies covering various coastal waters (Lee et al., 2005a, 2005b; Saulquin et al., 2013; Zhao et al.,

2013). The delopment of semi-analytical algorithms as part of SeaDAS software for deriving water

optical parameters (e.g. Kd, chlorophyll) from ocean colour sensors (e.g., MODIS, MERIS, HICO)

would facilitate the wider implementation of this rather complex algorithm for marine and coastal

applications.

There are some limitations of the present study and future work is expected to address

these issues:

• Only two point measurements of water optical profiling were made at the site of Wanga

Wallen Banks, which may not represent the optical characteristics of the whole Eastern

Banks. The measurement of water optical properties in other major bank areas of the Eastern

Banks is required to observe if the distinct seagrass environment in different individual

banks (Roelfsema et al., 2014) also reflects differences in the underwater light climate. In

addition to providing more comprehensive information on the overall underwater light

climate in the area, such measurements would also be useful for providing in situ validation

for satellite-based light parameters (such as bottom PAR, Kd490 and KdPAR in this study) and

parameters required for water column correction of satellite imagery (such as Kdλ).

• The light mapping part of this study (section 3.3.4) used satellite data with different pixel

sizes (MERIS PAR of 300 x 300 m, HICO KdPAR of 90 x 90 m and WorldView-2

Bathymetry of 2 x 2 m) for which the homogeneity assumption was given but the

quantitative assessment was unpredictable. Future refinement could be made by using PAR

parameters obtained from HICO satellite data that would enable the analysis of PAR and

KdPAR at the same spatial resolution (90 m of HICO spatial resolution). The key to this

possibility lies in extending the PAR algorithm of Frouin et al. (1989) to HICO data. It may

be tempting to acquire PAR, KdPAR and other satellite-based light parameters at the spatial

resolution of optical sensors (e.g., 2 m resolution of WorldView-2 data). However, satellite-

Page 72: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

52

based light parameters were acquired at a high spectral resolution (± 1 - 5 nm) as opposed to

the wide spectral resolution of optical sensors ( ± 50 - 60 nm). There is an inherent trade-off

between optimum spectral and spatial resolution that limits the acquisition of satellite-based

light parameters at very high spatial resolution (e.g., Jensen, 2005).

• The WorldView-2-based bathymetry used in this study still shows some anomalous deep

pixels in the shallow area. This may reflect the drawback of the Lyzenga method being used

to derive the bathymetry map, as the method was originally developed in oligotrophic waters

(e.g., reef environment) with bright substrate (Lyons et al., 2011). In addition to refining the

in situ coverage of bathymetry sounding for better interpolation purposes, the physics-based

method of Brando et al. (2009) or Dekker et al. (2011) can be used to improve the quality of

the bathymetry map.

Page 73: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

53

Chapter 4: Estimating Seagrass Leaf Area Index and Biomass Using WorldView-2

Satellite Data

Main Findings

• The accuracy of the satellite image-based LAI and biomass maps was in the ranges of 48–73%

and 41–71%, respectively.

• Green band reflectance was the best predictor for seagrass LAI, however no statistically

significant relationships were found between reflectance in any spectral band and biomass.

• Statistically significant relationships were found between measured LAI and biomass,

confirming allometric relationship between morphometric and biomass characteristics of

seagrass.

• The underwater light quality appeared to influence the nature and strength of the relationships

between reflectance and biophysical parameters of seagrass.

• The depth normalised green band (500–600 nm), corrected using the derivation of Beer's Law

for water column effect, was the best predictor for mapping seagrass LAI.

4.1. Introduction

Leaf area index (LAI) and above-ground biomass (hereafter referred to as biomass) are

important parameters for studying ecological and environmental aspects of vegetation, including

seagrass. LAI, defined as the single-sided leaf area per unit ground area (Green et al., 1997;

Zimmerman, 2013; Zhao et al., 2012), is one of the most important factors for characterising plant

canopy structure and process. It represents the total photosynthetic area of the plant canopy, which

in turn determines standing stock (biomass). As a consequence LAI correlates well with

photosynthetic processes and has been used as a diagnostic variable for crop growth rate, radiation

intensity and above-ground height (Solana-Arellano et al., 2003). LAI has also been used for

modelling seagrass productivity for Zostera noltii, in the French Mediterranean coast (Plus et al.,

2001) and Thalassia testudinum, Zostera marina and Syringodium isoetifolium, in Florida Bay

(Zimmerman et al., 2003, McPherson et al., 2011 and Stoughton, 2008). Biomass, often expressed

as gram dry weight/m2, is a variable required for carbon stock measurement that can be used in

carbon sequestration strategy–policy to reduce emissions from deforestation (Heiskanen, 2012;

Zhang et al., 2014).

Page 74: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

54

Although non-destructive photographic methods for estimating seagrass LAI (Zhao et al.,

2012) and visual assessment techniques for biomass (Mellors, 1991; Mumby et al., 1997) have been

developed, destructive methods by harvesting seagrass are still the most dominant. Time-series

seagrass parameters have been collected in this way through international seagrass initiatives, for

example, SeagrassNet (www.seagrassnet.org), SeagrassWatch (www.seagrasswatch.org). These

methods are site-specific and may not be suited to large area mapping and monitoring. Satellite and

airborne images can provide synoptic and repetitive measurements over the seascape and may be

used to estimate seagrass biophysical parameters. Using this principle a wide variety of seagrass

parameters such as cover, biomass and species have been mapped and monitored using remote

sensing data at various spatial and temporal scales (e.g., Dekker et al., 2005; Phinn et al., 2008;

Roelfsema et al., 2009; Lyons et al., 2011, 2012).

Considering the importance of seagrass LAI and biomass, it is perhaps surprising that

studies addressing the derivation of both parameters using remote sensing data are scarce. A review

of the literature found seagrass biophysical parameter mapping in Dierssen et al. (2003), Wicaksono

and Hafitz (2013) and Hill et al. (2014) who explicitly attempted to map seagrass LAI using remote

sensing data. Despite some differences in their detailed procedures, their studies used empirical

methods relating in situ LAI and remote sensing reflectance. Similar empirical methods were also

applied in remote sensing studies for seagrass biomass mapping (Armstrong, 1993; Mumby et al.,

1997; Phinn et al., 2008; Knudby and Nordlund, 2011). Still within this empirical category,

Roelfsema et al. (2014) developed an algorithm to estimate seagrass biomass from in situ data. For

terrestrial vegetation, the other alternative method for producing image-based LAI/biomass is the

analytical method employing the inversion of canopy reflectance models (e.g., Myneni et al., 1997;

Peddle et al., 2004). In this method, a radiative transfer model of the forest canopy is used to take

into account the viewing and illumination geometry and radiometry of each pixel and estimate

vegetation structural properties (Wulder et al., 1998). Canopy reflectance models for submerged

aquatic vegetation (SAV) exist (e.g., Plummer et al., 1997; Zimmerman, 2003; Hedley and Enrique,

2010) but their applications for deriving image-based seagrass LAI and biomass have not been

explored. Though rarely applied, previous researchers have explored the relationship between

seagrass LAI and biomass using linear relationships between the two to estimate seagrass biomass

from LAI (Solana-Arellano and Echavarria-Heras, 2003 and references therein). However, the

application of this LAI-based biomass approach to remote sensing data remains unexplored.

Page 75: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

55

The goal of this chapter was to develop a method for mapping the spatial and temporal

distribution of seagrass LAI and biomass in the Eastern Banks, Australia using WorldView-2

satellite data. Specifically, the objectives were to: (a) examine the relationships between spectral

reflectance and LAI and biomass using in situ reflectance and seagrass samples data, (b) examine

the relationships between LAI and biomass using in situ seagrass data, (c) develop, apply and assess

the accuracy of regression models to estimate LAI and biomass using World-View-2 satellite data,

and (d) compare mapped LAI and biomass estimates to determine changes in the study site between

summer and winter seasons. An empirical method was first used by examining the relationship

between in situ reflectance data and LAI using correlation analysis. This was followed by

regression models between image reflectance and in situ LAI. Development of an algorithm to

estimate biomass was then attempted by relating in situ biomass to either in situ image reflectance,

or in situ LAI. Maps of seagrass LAI and biomass for the study area were derived from two scenes

of WorldView-2 satellite data representing Austalian winter and summer seasons.

4.2. Data and Methodology

4.2.1. Field data collection and processing.

4.2.1.1. Seagrass samples collection and analysis.

Two data sets of seagrass core samples from June 2012 (n = 74) and February 2013 (n = 44)

were used for this analysis. Only the 2012 data set was used for algorithm development and this was

separated into data used for model development (n = 42) and validation (n = 37). The 2013 data set

consisted of samples, from 32 points collected from different parts of the Eastern Banks that were

all used for validation, and 12 points from Wanga Wallen Banks collected simultaneously with in

situ reflectance data for examining the relationships between reflectance and biomass and LAI (see

next section). Seagrass samples collected for this study were part of multi-year seagrass mapping

and monitoring projects in the Moreton Bay area (e.g., Phinn et al., 2008; Roelfsema 2009, Lyons et

al., 2011, 2012) (Figure 4.1.).

Page 76: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

56

Figure 4-1. Seagrass sampling points of June 2012 and February 2013 used in this study plotted on seagrass map of June 2012 (Roelfsema et al., 2014) and ALOS AVNIR-2 of 14 July, 2010, along with the names and boundaries of major individual banks in the Eastern Banks.

The seagrass samples were collected using 15 cm diameter x 20 depth of PVC core and were

kept frozen at -20°C until they were ready for analysis. For biomass and LAI analysis, the samples

were defrosted in fresh water and subsequently sorted into above-ground (pigmented leaves and

stems) and below-ground (root and rhizome) biomass per species per sample. Detritus was also

separated and processed, but for this study only the above-ground biomass component and LAI

were used in the calculation. Calcareous epiphytes from seagrass leaf blades were removed by a

quick immersion in 10% HCl before the samples were weighed. Samples were then oven-dried at

60°C for 48 hours and biomass was calculated as the difference between initial wet and final dry

weights of the samples weighed to the nearest 0.0001 g using an electronic balance. The final

biomass data were expressed as weight per area (g/m2). As the seagrass component separation was

conducted per species per sample, this method can produce both total biomass (e.g., all species) and

per species biomass per core sample.

For the purpose of LAI calculation, during biomass analysis, three random intact shoots

were sub-sampled per species and per core sample and photographed along with a ruler on the

Page 77: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

57

picture frame. The leaf areas were calculated photographically as the average of the three shoots

measured using open source imageJ software (http://imagej.nih.gov/ij/index.html) and expressed in

m2 of leaf area. LAI was then calculated either per species, or total species per core, by dividing

measured leaf area by core area. Similar to the biomass analysis product, this method resulted in

two types of LAI , total LAI for all species and species LAI per sample.

4.2.1.2. In situ reflectance data.

Simultaneous in situ reflectance data at the time of seagrass core collection were only

available from 12 points of the 2013 data set collected from the Wanga Wallen site (Figure 4.1.).

The core samples and their corresponding reflectance data were collected over different species and

the density was estimated by visual inspection. The underwater spectral device used consists of an

Analytic Spectral Device (ASD) FieldSpec HandHeld2 spectrometer (325–1075 nm, sampling + 1

nm, spectral resolution < 3 nm and 25° field of view) (ASD, 2011). The equipment is contained in

an acrylic underwater housing enabling neutral buoyancy and full external operation. In situ

reflectance for each point was taken before seagrass core sampling to avoid the stir-up of sediment.

For each in situ reflectance measurement, the first and second measurements were used to obtain

dark object (in this case was the black part of the operator's wet suit) and white reference panel

measurements respectively and then followed by eight reflectance measurement replicates of the

targeted seagrass core samples. The dark object measurements were used to monitor the consistency

of the spectrometer’s response with no light condition, while the white reference panel served as an

adjustment to ambient light, which could change from one point to another. The protocol developed

by the Biophysical Remote Sensing Group of the University of Queensland (BRG, 2013) or

Borrego-Acevedo et al. (2014) contains more detailed descriptions of the instrument and the

measurement procedure used.

4.2.2. Band selection using in situ reflectance data.

Band selection was conducted using Pearson correlation analysis applied to 12 points of the

in situ 2013 data set where in situ reflectance measurements and seagrass core samples were

collected simultaneously. Although it was possible to develop the regression model using in situ

data, for example, regressing in situ reflectance against in situ LAI/biomass and then applying the

model to satellite data, this was not implemented for two reasons. Firstly, the 12 data points have

low values of LAI and biomass compared with those of other points and cannot be used to represent

both parameters for the whole study site. Secondly, as the in situ reflectance data were taken within

close proximity to the seagrass core samples, the resultant model would not take into account water

Page 78: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

58

column effects present in the model implementation using satellite data. However, these limited

in situ reflectance data were still valuable for examining the nature of the relationship between

reflectance and LAI/biomass, as they were taken within close range of the seagrass samples (< 1.0

m) where the interfering atmospheric and water column effects were assumed to be minimal.

Prior to statistical analysis the in situ reflectance, data in the range of 400–700 nm were

resampled to the bandwidths of the WorldView-2 satellite data using ENVI® software, resulting in

five WorldView-2 image bands centred at 428.58 nm (coastal), 479.35 nm (blue), 548.07 nm

(green), 607.78 nm (yellow), and 658.92 nm (red). The bands and ratios examined in this study

(Table 1) were based on the maximum absorption of plant pigment at blue and red wavelengths and

the minimum absorption at green wavelength (Morel and Prieur, 1977). The blue–green ratio is the

basic algorithm for the extraction of ocean surface chlorophyll from ocean colour satellite data

(Morel and Prieur, 1977; O'Reilly et al., 1998). Additionally, correlation analysis was also applied

to the in situ LAI and biomass for both the years 2012 and 2013.

Table 4-1. Spectral Index Examined in This Study Calculated Using In Situ Reflectance Data

Note. WorldView-2 satellite bands (central bands) : coastal (428.58 nm), blue (479.35 nm), green (548.07 nm), yellow (607.78 nm) and red (658.92 nm).

4.2.3. Image processing and model development.

Two cloudless scenes of WorldView-2 satellite images (Digital Globe, Longmont, CO,

USA) acquired on 12 June 2012 and 3 February 2013 were available for analysis. The time

difference between each image acquisition and the fieldwork conducted in 2012 and 2013 was

3–10 days. Image pre-processing included atmospheric correction to at-surface reflectance using

FLAASH module (Fast Line-of-Sight Atmospheric Analysis of Spectral Hypercubes) in ENVI

(ITT, 2013) and geometric correction using co-registration of the WorldView-2 satellite images to a

Index Formula SourcesBand RatiosBand ratio 1 coastal/green This studyBand ratio 2 blue/green Morel and Prieur (1977)Band ratio 3 coastal/yellow This studyBand ratio 4 blue/yellow This studyBand ratio 5 red/green This studyNDVI-like RatiosVegetation Index 1 (green - coastal)/(green + coastal) This studyVegetation Index 2 (green - blue)/(green + blue) This studyVegetation Index 3 (yellow-log coastal)/(yellow + coastal) This studyVegetation Index 4 (yellow - blue)/(yellow + blue) This study

Page 79: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

59

pan-sharpened Quickbird image acquired on 4 July 2007. The Quickbird image was geometrically

corrected using ground control points positioned using a differential GPS receiver (Lyons et al.,

2011). Additionally, a bathymetry map produced from the 2012 WorldView-2 image using Lyzenga

method (Lyzenga, 1978; Lyons et al., 2011) was used to produce bottom reflectance image.

Model development was conducted by regressing in situ 2012 LAI or biomass to the 2012

WorldView-2 satellite imagery. The final parameters for regression analysis were based on the band

selection implemented in the previous step (section 4.2.2). In addition to linear regression, non-

linear regression analysis was also conducted as several studies reported non-linear linear

relationships between the forest variable including LAI (e.g., Myneni et al., 1997, 2002; Anderson

et al., 2004) or seagrass biomass (Bargain et al., 2012) and reflectance. The best model in terms of

coefficient of determination (R2) was selected for the subsequent mapping process.

To account for water column effect on the image data two water column correction

techniques were used: Lyzenga method (Lyzenga, 1978) and Beer's Law derivation implemented by

Dierssen et al. (2003). The Lyzenga method produced a depth-invariant index from two bands

where the influence of water column on benthic substrate was assumed to be minimal:

𝐷𝑒𝑝𝑡ℎ − 𝑖𝑛𝑣𝑎𝑟𝑖𝑎𝑛𝑡 𝑖𝑛𝑑𝑒𝑥𝑖𝑗 = ln(𝐿𝑖− 𝐿𝑠𝑖) − ��𝑘𝑖𝑘𝑗� . 𝑙𝑛�𝐿𝑗− 𝐿𝑠𝑗�� (1)

Here L, Ls and K are radiance, radiance over deep water and the diffuse attenuation

coefficient, respectively, and i and j refer to band i and j of the satellite data, respectively. While the

output of Lyzenga method is a band index from the combination of two bands, the Beer's Law

derivation of Dierssen et al. (2003) is applied on a single band producing a water column-corrected

band or bottom reflectance (Rb). The formula is:

𝑅𝑏 = 𝑅𝑟𝑠 𝑥 𝑄𝑏 𝑡

× 𝑒𝑥𝑝[−𝐾�����𝐿𝑢(𝑧) 𝑥 𝑧]𝑒𝑥𝑝(−𝐾𝑑 𝑥 𝑧) (2)

where Rrs is the remote sensing reflectance; t is light transmittance through air-water

interface, estimated as 0.54 (Mobley, 1994; Dierssen et al., 2003); Qb is the ratio of upwelling

irradiance (Eu) to upwelling radiance (Lu) at the bottom depth estimated as π (Dierssen et al., 2003);

−𝐾�����𝐿𝑢 is the depth-averaged upwelling diffuse attenuation coefficient; Kd is the downwelling

diffuse attenuation coefficient; and z is water depth. −𝐾�����𝐿𝑢 and Kd were parameterised from

Hydrolight® software with inputs of water quality parameters (chlorophyll and total suspended

Page 80: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

60

matter) from water sampling conducted in 2013 and of inherent optical properties (IOPs) from

either optical models embedded in Hydrolight or measured IOPs from Phinn and Dekker (2004).

Total suspended matter and spectrometer-based chlorophyll were measured using the method of

Parsons et al. (1984).

An LAI-based biomass model was developed by establishing relationships between in situ

LAI and biomass of the 2012 data set, either as a community (total species) or per species for each

sample using regression analysis. This analysis was conducted to examine the estimation of

seagrass biomass from LAI (McRoy, 1970; Hackney, 2003; Solana-Arellano et al., 2003).

4.2.4. Reliability statistics for seagrass LAI and biomass estimation.

Quantitative error estimation was done by matching the estimated LAI and biomass maps

from the 12 June 2012 and 3 February 2013 WorldView-2 satellite images to corresponding data

validation points collected on 7–10 June 2012 and 12–15 February 2013. The map accuracy

calculation of Green et al. (1997) for validating mangrove image-based LAI map was used in this

study. The accuracy was calculated as the proportion of validation points that lay within the 95%

confidence interval of the regression model between estimated (image) and in situ LAIs.

95% confidence level (Zahr, 2010) = 𝑌𝑜 ± (𝑆𝐸 𝑜𝑓 𝑌𝑜) × 𝑡(0.05)(2) 𝑑𝑓(𝑛−2) (3)

SE (Standard error) of Yo = �(𝑌−𝑌𝑜)2������������

𝑛 (4)

where Y is the estimated LAI, Yo is the in situ LAI, df is degree of freedom, n is number of

observations and t is critical value of the t distribution table.

Page 81: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

61

4.3. Results and Discussion

4.3.1. Seagrass biomass and LAI from field sampling

Figure 4-2. Seagrass Biomass and LAI spatial patterns from field sampling in the clockwise direction : biomass 2012, biomass 2013, LAI 2012 and LAI 2013, plotted on seagrass areal distribution of June 2012 (dark green area) (e.g Roelfsema et al., 2014).

Figure 4.2. depicts the spatial patterns of collected in situ biomass and LAI data, which were

subsequently analysed in the laboratory. The range of variability in the biomass 2012 data is slightly

higher than that of biomass 2013. For both years, the range of biomass values observed in this study

Page 82: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

62

are far greater than that of Phinn et al. (2008) who observed in situ biomass values of 0 - 34

g/dw/m2 in the same location. The highest biomass values of 2012 data was contributed by species

Cymodacea serulata, while the wide-blade type of Zostera muelleri contributed to the highest

biomass value for 2013 data. For general comparison, Knudby and Nordlund (2011) found seagrass

biomass of various species in the range values of 0 - 700 g/dw/m2 in Chumbe Island, Zanzibar,

Tanzania which has seagrass depth limit of 7 m. Three species of Eastern Banks (location of the

present study) are present in their study site (Cymodacea serulata, Halodule uninervis, Syringodium

isoetifolium) with biomass values of 10 - 350 g/dw/m2 (visual inspection of Figure 5 of Knudby and

Nordlund, 2011), which is comparable to the result of this study. The highest biomass value of their

result was contributed by species Thalassodendron ciliatum. Gonzalez-Liboy (1979) found biomass

values of Thalassia testudinum in La Parguera, Puerto Rico in the range of 400 - 2300 g/dw/m2,

whereas Armstrong (1993) observed biomass value range of < 5 g/dw/m2 in the shallow bank near

Lee Stocking island for the same species. These variability in seagrass biomass reflect variation in

species and local environmental condition of seagrass habitat (e.g. Lee et al., 2007a).

In terms of biomass spatial pattern among different banks, high seagrass biomass dominates

the Wanga Wallen Banks for both years (Figure 4.2). The banks is characterised by a more

sheltered area with oceanic flushing through tidal dynamics, allowing more light penetration to the

depth of seagrass and less disturbances from wave and current (Phinn et al., 2008; Lyons et al.,

2011; Roelfsema et al., 2014).

Similar to the biomass values variability, the maximum value of LAI of 2012 data is higher

than that of 2013 data, which was contributed by species Cymodacea serulata. If this species were

not present in the 2012 data, the value ranges of LAI for 2012 and 2013 data would have been very

similar. A similar pattern also emerges, where the Wanga Wallen Banks is dominated by high LAI

values. Unlike with biomass, previous studies in Eastern Banks for comparative purpose to the

present study were not available.

4.3.2. Correlation between LAI, biomass and in situ reflectance data.

All individual spectral bands of in situ reflectance data resampled to WorldView-2 image

bands were significantly correlated (p < 0.05) with in situ LAI (Table 4.2.). The form of the

correlations were all negative. With the exception of band ratios 2 and 3 (see Table 4.1. for the

complete description of the band ratios examined here), the standard logarithmic transformation

was found to enhance all correlations (Table 4.2.). For the single bands, the degree of correlation

progressively increased from the coastal band and reached a maximum at the green band, before

starting to decrease with minimum correlation at the red band.

Page 83: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

63

Table 4-2. Correlation Coefficients (r) of the LAI and Biomass Against the Resampled In Situ Reflectance Data

Note. Only r values with p < 0.05 significance were displayed (n = 12). See Table 4.1. for the complete description of the band ratios, vegetation index and the wavelengths of the bands examined here

It can be shown from the plots of LAI against the five WorldView-2 satellite bands

resampled from in situ reflectance data that the relationships are somewhat curvi-linear for the

coastal and blue bands (Figure 4.3 a and b).

Spectral Band/Index LAI BiomassCoastal -0.788 -0.579Blue -0.800 -0.579Green -0.838Yellow -0.821Red -0.787Log Coastal -0.860Log Blue -0.869Log Green -0.871Log Yellow -0.867Log Red -0.849Band ratio 1 (Coastal/Green) -0.658Band ratio 2 (Blue/Green) -0.599Band ratio 3 (Coastal/Yellow) -0.650Band ratio 4 (Blue/Yellow)Band ratio 5 (Red/Green)Vegetation index 1 (green - coastal)/(green + coastal) 0.654Vegetation index 2 (green - blue)/(green + blue) 0.591Vegetation index 3 (yellow-log coastal)/(yellow + coastal) 0.650Vegetation index 4 (yellow - blue)/(yellow + blue)Log Band ratio 1 -0.638 -0.629Log Band ratio 2 -0.648 -0.614Log Band ratio 3 -0.605Log Band ratio 4 -0.606Log Band ratio 5Log Vegetation index 1 0.742 0.594Log Vegetation index 2 0.695Log Vegetation index 3 0.698 0.591Log Vegetation index 4 0.654 0.589

Page 84: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

64

Figure 4-3. Plots showing the relationships between in situ LAI and WorldView-2 satellite band coastal (a), blue (b), green (c), yellow (d) and red (e).

The relationship between LAI and reflectance becomes more linear at the green and yellow

bands (Figure 4.3, c and d) and then the plotted points become more scattered at the red band

(Figure 4.3, e). The linear LAI–reflectance pattern explained the strong correlation between LAI

and green (r = 838) and yellow (r = 821) bands as reflectance values from these two bands are more

sensitive to changes in LAI than reflectance in the other bands. It is interesting to note that the log

vegetation index 1 also exhibited a good correlation with LAI. In the case of vegetation index 1,

vegetation reflectance at coastal band (blue wavelength) was lower than that of the green band.

Among all bands and band combinations, the highest correlation was found for the log green band.

Page 85: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

65

Like all higher plants, primary pigment in seagrass – the chlorophylls – absorb light at blue and red

wavelengths for the photosynthetic process (Kirk, 2010), which explains the negative correlation

between reflectance in the two bands and LAI. However, the presence of carotenoids can extend the

absorption range of seagrass into the blue-green region (Kirk, 2010), which may have caused the

negative correlation between LAI and the green band. In general, the negative correlation between

LAI and the visible bands found here was consistent with previous findings for terrestrial vegetation

using Aster satellite data (Heiskanen, 2006) and Landsat TM data (Lu et al., 2004) and for seagrass

using hyperspectral airborne data (Dierssen et al., 2003; Hill et al., 2014).

In addition to the physical nature of the interaction between sunlight and vegetation

pigments, it is likely that the underwater light conditions of seagrass environments also affect the

correlation between LAI and reflectance. Figure 4.4. shows the diffuse attenuation (Kd) measured

close to the seagrass location in Wanga Wallen Banks where simultaneous in situ reflectance data

and seagrass cores were collected.

Figure 4-4. Average diffuse attenuation coefficients (Kd) measured in situ using a radiometer device during a field trip in 2013 for two points close to the seagrass sample collection sites. The measurement was conducted by lowering a radiometer to a number of depths in a 30 min–1hr time interval. Kd was then calculated using standard light exponential principle of Beer's law. The relative positions of blue, green and red wavelengths were indicated.

Light with low Kd indicates that it penetrates deeper into the water column than light with a

high Kd. It can be seen from Figure 4.4. that the blue and red bands have much higher Kd values

than Kd at the green wavelengths. To account for the impact of this Kd pattern on light at the depth

of seagrass, light as measured using a radiometer at the bottom depth is presented in Figure 4.5.

Page 86: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

66

Figure 4-5. Light energy as measured using radiometer device at seagrass depth for two points close to seagrass sample collection sites, expressed as blue, green and red lights.

Figure 4.5. shows that green light was more abundant in this seagrass-dominated water than

the blue and red light that would allow more interaction between green light and seagrass. Using

similar techniques of simultaneous underwater spectroradiometer measurement, and collecting sand

samples in the blue dominated waters, Heron reef (Veal, 2010), Borrego-Acevedo et al. (2014)

found significant correlations between the chlorophyll pigment of microphytobenthos and

reflectance at blue wavelengths.

Page 87: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

67

Only two single bands and seven band combinations showed strong correlation with

biomass (r values with p < 0.05 significance) (Table 4.2). However, all correlations at these bands

were stronger with LAI than with biomass. The result indicated that reflectance was not a good

estimator for biomass. Despite the abundant literature on the application of remote sensing for

estimating terrestrial forest biomass, such studies were scarce for seagrass. According to their

approaches, these limited studies can be categorised into direct methods using relationships

between: (a) image reflectance and in situ biomass (Armstrong, 1993; Phinn et al., 2008); (b) in situ

reflectance and biomass conducted in an experimental setting (Bargain et al., 2012), and (3) indirect

methods using the relationship between image reflectance and visually-scaled in situ biomass

(Mumby et al., 1997; Knudby and Nordlund, 2011), or using a predictor algorithm developed

between in situ biomass and other in situ parameters (e.g., cover and species) (Roelfsema et al,

2014). There are similarities between the present study and approach (b) (Bargain et al., 2012).

However, the purpose of their study was to examine the effect of various levels of simulated

turbidity on biomass estimation and the setting of their study was experimental using controlled

tanks. The resultant low correlation between in situ reflectance and biomass found in this study was

likely because biomass is not a natural physical parameter from which its interaction with light can

be inferred. Biomass is a processed product from laboratory analysis after undergoing oven-drying

and weighting steps. While LAI is a calculated parameter it represents the leaf surface area, which

has a spectral interaction with light that is well-recognised and became the basis of remote sensing

study for vegetation application (e.g., Jensen, 2005).

4.3.3. Correlation between LAI and biomass.

Correlations between LAI and biomass for 2012 and 2013 data set were strong for

community (total species) or per species, except for species Syringodium isoetifolium in the 2013

data set (Table 4.3.).

Page 88: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

68

Table 4-3. Correlation Coefficients (r) of LAI Against Biomass for Community (Total Species) and Per Species of 2012 and 2013 data set

Note. Only r values with p < 0.05 significance were displayed.

Duarte (1991) compiled data on the architectural (also commonly termed as morphometric)

(e.g., leaf surface area) and dynamic (e.g., leaf production rate) parameters of seagrass across

different species and geographic areas to establish the relationships between the two parameters.

His analysis revealed strong and significant allometric relationships between seagrass architectural

and dynamic parameters, suggesting the importance of architectural characteristics of seagrass in

providing the basis of predicting seagrass productivity and ecological status. In line with this

allometric relationship concept, McRoy (1970) found high correlation (r = 0.88) between leaf length

and leaf biomass of Zostera marina, of Alaska. Statistically significant linear relationships between

seagrass LAI and biomass were observed by Gonzalez-Liboy (1979) for Thallasia testudinum, of

Puerto Rico. Similar to the research of Duarte (1991), Hackney (2003) attempted to establish

allometric relationships between morphometric and biomass characteristics of Thallasia testudinum,

of Florida Bay and found high correlation (r2 = 0.969, p < 0.001) between LAI and biomass.

Despite his findings on the significant correlation between LAI and biomass of Zostera marina,

McRoy (1970) mentioned the requirement to evaluate the general application of the relationship on

a seasonal basis. The correlation analysis result presented in Table 4.3. confirmed that high

correlations (except for Syringodium isoetifolium in the 2013 data set) were consistent seasonally as

the 2012 and 2013 data set were collected in winter (June) and summer (February), respectively.

4.3.4. Models for estimating biomass and LAI from image data.

4.3.4.1. Model for LAI.

Based on the correlation analysis between LAI and reflectance from the previous step,

model development was focussed on the use of band 3 (green) of WorldView-2 image data to

estimate LAI. The result of the regression analysis showed that the r2 values varied from 0.14–0.61

Seagrass category (community/species) 2012 2013Community (Total species) 0.863 (n=97) 0.836 (n=44)Zostera muelleri 0.888 (n=52) 0.859 (n=24)Halophila ovalis 0.825 (n=46) 0.802 (n=32)Halophila spinulosa 0.935 (n=21) 0.946 (n=6)Cymodacea serrulata 0.871 (n=20) 0.959 (n=7)Halodule uninervis 0.888 (n=17) 0.909 (n=25)Syringodium isoetifolium 0.858 (n=9)

Page 89: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

69

(Table 4.4.) with the highest value achieved by using the linear model of Rb 3 (r2 = 0.61)

(Figure 4.6.).

Table 4-4. Regression Models, Coefficients of Determination (R2) and Probability Corresponding to F Value (Pr > F, only available for linear model) for LAI

Figure 4-6. Regression plot between in situ LAI and log Rb 3. The data used for the regression were in situ LAI and WorldView-2 data of 2012. The complete results of the regression analysis are given in Table 4.4.

The next best predictor for the LAI was the exponential model of single band 3. The

logarithmic nonlinear model considerably improved the r2 for the single band 3, whereas all non-

linear models resulted in lower accuracy levels for data from the Lyzenga and the Rb 3 algorithms.

In terms of the type of the regression model used, previous studies found strong relationships

between image or aircraft reflectance and forest LAI using either linear (e.g., Heiskanen, 2006) or

Band or Transformed Band

Model R2 Pr > F Model R2

Green band (band 3) y = -19.33x + 2.57 0.35 0.0001 y = 2.66e-14.71x 0.38y = -1.68ln(x) - 3.45 0.59

Lyzenga band 1 and 3 y = 0.33x - 0.064 0.29 0.001 y = 0.39e0.23x 0.27y = 0.48ln(x) + 0.74 0.14

Rb 3 Y = -3.69x - 0.50 0.61 < 0.0001 Y = 0.315e-2.40x 0.50

Linear Regression Model Non Linear Regression ModelRegression analysis of Y = LAI, X = Band/Transformed Band

Page 90: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

70

nonlinear (e.g., Myneni et al., 1997, 2002; Anderson et al., 2004) regression analysis. Green et al.

(1997) also observed significant linear relationships between mangrove LAI and SPOT XS

reflectance in the Turks and Caicos Islands, in the British West Indies. Similar studies for seagrass

LAI were scarce and they were usually presented with less detailed information compared with

those of terrestrial studies (e.g., no accuracy assessment or statistical p significance value).

Wicaksono and Hafizt (2013) attempted to estimate seagrass LAI from several band transformations

of ALOS AVNIR-2 and ASTER VNIR reflectance and found R2 of linear regression analysis in the

range of 0.24–0.46 at the community level (total species) and of 0.11–0.559 for the species level.

The results obtained in the present study agree with those of Dierssen et al. (2003) using

both in situ and hyperspectral airborne image reflectance, and Hill et al. (2014) using airborne

image reflectance to estimate seagrass LAI from Rb at wavelength 555 nm (green spectrum). The

model obtained in this study (LAI = -3.69 Log Rb (548.07 nm) - 0.49; R2 = 0.61) was also similar

to those of Dierssen et al. (2003) (LAI = -3.05 Log Rb (555 nm) - 0.98; R2 = 0.88 - 0.98) and Hill et

al. (2014) (LAI = -3.14 Log Rb (555 nm) - 2.64; R2 = 0.81). The differences in the slope and

intercept of the models may reflect differences in the characteristics of the seagrass canopy and

sediment reflectance respectively (Hill et al., 2014). Myneni et al. (1995) provided a physical basis

for the relationship between LAI and spectral index or related bands and showed that spectral

index/band was indicative of the abundance and activity of the absorbers of leaves, and therefore

could be used to formalise relationships between vegetation reflectance spectra and leaf

biochemical constituents.

Water column correction using Rb algorithm proved useful for improving the regression

between image reflectance and LAI (Table 4.4). A water column correction technique using the

Lyzenga method was beneficial in establishing the relationship between image reflectance and

seagrass biomass (Armstrong, 1993; Mumby et al., 1997 and Knudby and Nordlund, 2011).

Wicaksono and Hafizt (2013) found improvement on their LAI–image reflectance correlation when

water column correction, based on the Lyzenga method, was applied. The superiority of the Rb

algorithm over the Lyzenga method observed in this study emphasised the necessity to use both the

depth-averaged upwelling diffuse attenuation (−𝐾�����𝐿𝑢) coefficient and the downwelling diffuse

attenuation coefficient (Kd) for water column correction, instead of only Kd as commonly used by

the Lyzenga method.

Page 91: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

71

4.3.4.2. Model for biomass.

Considering low correlation between in situ reflectance and LAI and high correlation

between in situ LAI and biomass, a model for biomass was developed from LAI. The result of the

regression analysis between LAI and biomass is shown in Table 4.5. for both community (total

species) and species levels.

Table 4-5. Regression Models, Coefficients of Determination (R2) and Probability Corresponding to F value (Pr > F) for Biomass

Note. Only the linear model is presented as all non-linear models showed less significant results

Regression analysis applied to total species and per species showed that LAI significantly

explains 64–92% of variation observed in biomass indicated by the range of R2 values from 0.64–

0.92 and p < 0.0001. Nonlinear regression analysis was found to reduce all R2 values, confirming

previous findings of the linear relationship between LAI and biomass (McRoy, 1970; Gonzalez-

Liboy, 1979; Hackney, 2003). The regression plots of community level of 2012 and 2013 data sets

and of Zostera muelleri (the most dominant species) and Cymodacea serulata (species with the

largest blade) of the 2012 data set were presented in Figure 4.7.

Seagrass category (community/species)

Model R2 Pr > F Model R2 Pr > FCommunity (Total species) Y = 24.04*X + 10.22 0.74 < 0.0001 Y = 22.66*X + 10.86 0.70 < 0.0001Zostera muelleri Y = 15.60*X + 9.44 0.79 < 0.0001 Y = 28.73*X + 0.13 0.74 < 0.0001Halophila ovalis Y = 34.18*X + 0.02 0.68 < 0.0001 Y = 26.21*X + 0.50 0.64 < 0.0001Halophila spinulosa Y = 28.06*X - 0.40 0.87 < 0.0001 Y = 25.91*X - 0.01 0.90 0.004Cymodacea serrulata Y = 22.74*X + 23.82 0.76 < 0.0001 Y = 19.65*X + 11.10 0.92 0.001Halodule uninervis Y = 18.01*X + 3.05 0.79 < 0.0001 Y = 18.55*X + 2.44 0.83 < 0.0001Syringodium isoetifolium Y = 48.66*X + 4.54 0.74 0.003 Y = 72.83*X - 6.95 0.76 0.129

Linear Regression Model Biomass (Y) and LAI (X)2012 2013

Page 92: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

72

Figure 4-7. Regression plots between LAI and biomass for community level of (a) 2012 and (b) 2013 data sets and (c) species Cymodocea serrulata of 2012 data set. The complete regression result are given in Table 4.5.

Despite the different number of data points used for the analysis at community and species

levels, the relationship between LAI and biomass showed a consistent, strong and significant linear

pattern overall, with points only deviating from linearity at higher values. This indicates an

optimum correlation between the two parameters at certain points before the complete deviation

takes place. In terms of seagrass physiology there appears to be an optimum LAI value that supports

optimum photosynthesis beyond which photosynthesis will decrease due to leaf self-shading

leading to a lower biomass at higher LAI values (Short, 1980). The relationship between LAI and

biomass observed in this study was in agreement with the concept of allometric relationships set out

by Nielsen and Sand-Jensen (1990) and Duarte (1991), and with similar studies by McRoy (1970),

Gonzalez-Liboy (1979) and Hackney (2003).

In regard to the option between community or species LAI algorithms, since the method

presented here produced models for both levels, the choice depends on the capability of the remote

sensor to discern seagrass species. When species levels are discernible, then species-based

algorithms should be applied, otherwise a community-based algorithm is used. However, since the

a) b)

c)

Page 93: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

73

LAI model was developed at the community level, the biomass model at community level was used

for the mapping process. Considering the seasonal consistency of the form of the biomass models at

community level (Table 4.5.) and the more spatially distributed points used to develop the 2012

model, the biomass map was then generated using biomass model of the 2012 data set.

4.3.5. Mapping seagrass LAI and biomass using WorldView-2 image data.

The LAI and biomass maps derived from WorldView-2 data of June 2012 and February

2013 were shown in Figure 4.8. Since biomass was estimated from LAI using a linear regression

model (section 4.3.3) the resultant biomass maps were very similar to the LAI maps and here, for

each year data set, both parameters were presented as a single map with two different legends,

representing the two parameters. The areal coverage of each LAI and biomass class for individual

banks were shown in Figure 4.9. and Figure 4.10.

Page 94: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

74

Figure 4-8. LAI and biomass maps of June 2012 (left) and February 2013 (right) calculated by applying equation in Figure 4.5 and equation in Table 4.5 for community level (the firs list) to WorlView-2 satellite data.

Land Deep

Shallow

(g dw m-2)

Page 95: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

75

Figure 4-9. Graphics of total surface area of each LAI class in each individual banks for June 2012 and February 2013 as calculated by applying equation in Figure 4.5 to WorlView-2 satellite data.

LAI

0–2 2.1–4 4.1–6 6.1– 8 8.1–10 10.1–12 12.1–16

Page 96: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

76

Figure 4-10. Graphics of total surface area of each biomass class in each individual banks for June 2012 and February 2013 as calculated by applying equation in Table 4.5 for community level (the firs list) to WorlView-2 satellite data.

From June 2013 to February 2013 the whole Eastern Banks showed consistent proportions

of the seven LAI and biomass classes used in this study (Figure 4.9. and Figure 4.10). The four LAI

Biomass Classes (g dw m-2)

0–50 51–100 101–150 151– 200 201–250 251–300 301– 400

Page 97: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

77

classes from largest-to-smallest area were LAI 2.1–4, LAI 0–2, LAI 4.1–6 and LAI 6.1–8. Biomass

class (in g · dry weight · m-2) of 50–100 dominated the Eastern Banks, followed by classes of 101–

150, 0– 50 and 151–200 (Figure 4.10). The three remaining LAI and biomass classes were present

in very small proportions and varied slightly among them. Larger variations in the proportion of

LAI and biomass classes were observed when considering individual banks.

One noticeable feature in the middle upper part of Figure 4.8. is the change from dominant

red to green colour. This area belongs to the Moreton Banks which shows a change from dominant

LAI 0–2 class to LAI 2.1–4 class (Figure 4.9.), corresponding to changes in the biomass class of 0–

50 to 51–100 (Figure 4.9.). This change may reflect the seasonal change of the shoot density of the

dominant species in Moreton Banks. According to Roelfsema et al. (2014) Moreton Banks is

dominated by species Z. muelleri, H. ovalis and H. Spinulosa. Preen (1992) studied temporal

variation in several seagrass parameters in the Eastern Banks and found that the shoot density and

above-ground biomass of H. ovalis, H. Spinulosa, H. Uninervis, C. Serrulata and S. isoetifolium

peaked in the summer–autumn period, while the peak time of the same parameters for Z.muelleri

was winter–spring. The study of Kocak et al. (2011) showed that the spatial and temporal changes

in LAI of P. oceanica during autumn, winter and spring was related to the corresponding change in

shoot density. Amity, Chain and Wanga Wallen Banks showed relatively stable or only slight

changes overtime in the proportion of LAI and biomass classes (Figure 4.9. and Figure 4.10.).

Similar to Moreton Banks, noticeable changes were observed at the Maroom Banks in the

sharp increase of LAI 0–2 and decrease of LAI 2.1–4 and LAI 4.1–6. Likewise, seasonal variation

in species-specific shoot density may contribute to these changes as seagrass abundance at Maroom

Banks was observed to change overtime along with an increase of H.spinulosa and decreases of Z.

muelleri and H. ovalis (Roelfsema et al., 2014). Wanga Wallen Banks is the only site that

comprised the highest LAI class (12.1–14) for both data sets though its areal class percentage is

small. The presence of high LAI classes in Wanga Wallen Banks (Figure 4.9.) is likely due to the

abundance of a species with larger blades, C. serrulata. Preen (1992) observed that the spatial

distribution of this particular species was very limited, occurring primarily in the protected lagoon

of Wanga Wallen Banks. This field-based observation by Preen (1992) was later confirmed using s

satellite-based species map by Phinn et al. (2008), Lyons et al. (2011) and Roelfsema et al. (2014).

Our laboratory analysis of seagrass samples showed that the highest LAI of 10.45 was contributed

by this species collected at the Wanga Wallen Banks.

Page 98: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

78

The average of biomass for individual banks is given in Table 4.6, along with the results of

previous studies for comparison purposes.

Table 4-6. Comparison of Average Biomass per Individual Banks

Note. Preen (1992) did not consider individual banks in his study and defined biomass as total above and below ground biomass but provided the value of above-to-below ground biomass ratio for the Eastern Banks. The above-ground biomass value given here is the average of his biomass points measurement in the Eastern Banks (Table 3.3. of his manuscript) converted to above-ground biomass using average above-to-below ground biomass ratio for the Eastern Banks (Table 3.5 of his manuscript).

The average biomass of the Eastern Banks are 79 and 76 g · dry weight · m-2 for June 2012

and February 2013, respectively. These values are very close to the value of 73 g · dry weight · m-2

observed by Roelfsema et al. (2014) but overestimate the value of Preen (1992) of 55 g · dry weight

· m-2. The average biomass of Maroom Banks of February 2013 is exactly the same as the one

observed by Roelfsema et al. (2014) with a value of 79 g · dry weight · m-2. However, the average

biomass of Amity, Chain and Wanga Wallen banks found in this study overestimate the values of

their study while underestimation occurs for the Moreton Banks. These discrepancies may reflect

differences in the methods used in this study (LAI-based biomass) and in Roelfsema et al. (2014)

(species-and-cover based biomass) and in Preen (1992) (point-based sampling), as well as the

difference in areal coverage (i.e., Moreton and Chain Banks coverage in this study was slightly

smaller than those in Roelfsema et al. (2014)).

4.3.6. Accuracy assessment of the satellite-based LAI and biomass.

Despite the common method of regressing observed and modelled parameters for assessing

the accuracy of satellite-based LAI or biomass, other options can be used. One approach is to use

the coefficient of determination (r2) as a determinant for overall accuracy of seagrass biomass or

LAI map (e.g., Armstrong, 1993; Mumby et al., 1997; Knudby and Nordlund, 2011; Wicaksono and

Hafizt, 2013; Roelfsema et al., 2014). A more rigorous method was modified from Green et al.

(1997) for mangrove LAI where accuracy was calculated as the proportion of validation points that

lay within the 95% confidence interval of the regression model between estimated (image) and in

Sources2012 2013 2012 2013 2012 2013 2012 2013 2012 2013 2012 2013

This Study 79 76 89 79 96 92 99 79 60 67 100 102Roelfsema et al.(2014)

Preen (1992)*

Average of Image data of June 2011, April 2012, June 2012, Februay 2013 and August 2013

Average Biomass (gr / m2)Eastern Banks Amity Chain Maroom Moreton Wanga Wallen

Average of point sampling of 428 sites in November - Dec 1989N/A55

5814879344873

Page 99: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

79

situ validation points. According to the first approach the accuracies of LAI maps for June 2012 and

February 2013 were 62% and 73%, respectively, whereas the accuracies of biomass maps for the

same time frame were 71% and 60% (Figure 4.11.). When the second approach was applied, the

accuracies decreased to 48% and 41% for both LAI and biomass maps of June 2012 and February

2013, respectively.

Figure 4-11. Plots between observed and estimated LAI of (c) 2012 and (d) 2013 and biomass of (c) 2012 and (d) 2013 data sets. The observed values are in situ data that were not used for model development.

Consideration needs to be made when comparing the accuracy of the biomass map in this

study to those of previous studies as they may have been developed using different methods (see

section 4.3.1 for different biomass mapping approaches). For comparison purposes, only studies

delivering accuracy assessments from validation points are presented here as some studies presented

Page 100: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

80

coefficients of determination from the model development points (e.g., Phin et al., 2008 for seagrass

biomass; and Dierssen et al., 2003 and Hill et al., 2014 for seagrass LAI). Using direct empirical

method of relating the Lyzenga-transformed Landsat TM and in situ biomass Armstrong (1993)

achieved r2 = 0.80 in the Bahamas. Knudby and Nordlund (2011) found the accuracy with r2 = 0.83

using the indirect method of relating the Lyzenga-transformed IKONOS reflectance to visually-

scaled in situ biomass in Chumbe Island, Zanzibar, Tanzania, but the accuracy decreased to r2 =

0.57 when species T.ciliatum was present. The same indirect method was applied by Mumby et al.

(1997) in Turks and Caicos and achieved coefficients of determination of r2 = 0.79 and r2 = 0.74

using SPOT XS and Landsat TM, respectively. Using a different variation of indirect methods

Roelfsema et al. (2014) found r2 = 0.81 for seagrass biomass maps generated using algorithms

developed from in situ seagrass cover and species in the Eastern Banks. In regard to LAI, only the

study of Wicaksono and Hafitz (2013) in Karimunjawa Islands, Indonesia provided coefficients of

determination from regression between ALOS AVNIR-2 and ASTER VNIR, and in situ LAI

validation points, with r2 = 0.26 and r2 = 20 for PCA-transformed ALOS AVNIR-2 and PCA-

transformed ASTER VNIR, respectively.

From figure 4.11 it is realised that the image-based model for LAI underestimates the

observed LAI values, whereas overestimation occurs with the biomass model. At first it was

speculated that depth factor contributed to these errors.

Figure 4-12. Transect line traversing various depth, biomass and values used for correlation analysis. The depth variation is shown on the right-side panel. Seagrass area is delimited by white line. Solid gray areas show land masses.

A

Meter

B

Page 101: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

81

Pearson correlation analysis was then implemented between depth, LAI and biomass values from a

transect traversing various depth values as depicted in figure 4.12. Table 4.7 shows that though

depth contributes effect to the derived biomass and LAI values, the degree of correlations are very

small, which may justify the depth correction procedure applied for this study.

Table 4-7. The result of Pearson correlation analysis between depth, and LAI and biomas. All values are significant at p < 0.05.

Two other factors may contribute to the underestimation of LAI : internal reflection inside

seagrass canopy causing reduced reflection (e.g. Zimmerman, 2003) and the absorption of epiphyte

of incoming PAR radiation (Drake et al., 2003). Using a two-flow model of plane irradiance

distribution through a seagrass canopy Zimmerman (2003) observed the modelled downwelling

irradiance were within 15 % of field measurement. It is still subject to further research to implement

the method for remote sensing application. Drake et al. (2003) found that turtlegrass epiphyte

absorbed 36 - 60 % of incident light, depending on the epiphyte loads. However, the absorption

occurred in the peak chlorophyll absorption bands (blue and red wavelengths), whereas the present

study used green band for LAI estimation. Unfortunately, the available previous studies on the

estimation of LAI from remote sensing (Dierssen et al., 2003; Wicaksono and Hafizt., 2013) did not

address the underestimation / overestimation of their models in relation to the observed data as they

only presented the correlation coefficient of their regression models.

Despite the availability of more published studies on the estimation of biomass from satellite

data than that of LAI estimation (Phin et al., 2008; Armstrong, 1993; Mumby et al., 1997; Knudby

and Nordlund, 2011; Roelfsema et al., 2014), it was only Roelfsema et al. (2014) who plotted their

model results against the observed values. In contrast to their study that found biomass

underestimation (figure 5 of Roelfsema et al., 2014), biomass overestimation occurred in the present

study (Figure 4.11). It is speculated that this overestimation was due to the reflectance contribution

by the background substrate as seagrass meadows in Eastern Banks are characterised by shallow

and clear environment (Hyland et al., 1989; Preen, 1992; Phinn et al., 2008; Lyons et al., 2011;

Roelfsema et al., 2014). In an experimental setting, Bargain et al. (2012) addressed the effect of

background substrate on the biomass of Zostera noltii using various vegetation index. Their results

showed that the blue band corrected vegetation index was promising for further application to

remote sensing data. Future work will address this issue.

Pearson Corr. R2 Pearson Corr. R2 Pearson Corr. R2 Pearson Corr. R2

0.384 0.148 0.481 0.232 0.384 0.148 0.384 0.232

Biomass 2012 Biomass 2013 LAI 2012 LAI 2013

Page 102: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

82

In addition to the above mentioned factors, the patchiness and heterogeneous nature of

seagrass in the study area (Phinn et al, 2008) could also be contributing to the overall accuracy of

the satellite-based maps in this study. The values of LAI > 10.45 and biomass > 231 g · dry weight ·

m-2 should be treated with caution because these values were the highest calibration points used in

the regression models. However, the result showed that the values greater than these calibration

points were present in small proportions (Figure 4.9 and Figure 4.10). One important factor that

could affect the accuracy of this type of study would be the suitable sampling scheme in relation to

the spatial resolution of the satellite imagery. The in situ biomass and LAI used for model

development in this study were calculated from seagrass samples collected using cores with the size

of 176.6 cm2 (0.08 m2). The use of the regression model thus implied that the seagrass biomass and

LAI sampled at the size of the core could represent the same parameters for a 2 m x 2 m area (the

pixel size of WorldView-2) whose validity assumption was unknown at that stage. However, the

expectation of a suitable sampling scheme at the satellite's pixel size should consider the trade-off

between the cost-effectiveness of using remote sensing methods and the logistical challenge of

collecting large in situ data.

Phinn et al. (2008) and Knudby and Nordlund (2011) found that species specific canopy

structure influenced spatial variations in biomass, suggesting the inclusion of species information

for deriving better image-based biomass maps, which was later implemented by Roelfsema et al.

(2014). This study showed that the correlation and regression analysis between LAI and biomass

was consistently high, for either species or communities (total species) (Table 4.3., Table 4.5. and

Figure 4.7.). This alternative indirect method, which has an underlying scientific background in

allometry, can be used in the future with necessary refinement to avoid the effect of this species-

specific canopy structure. In addition to quantitative accuracy assessment using regression analysis,

an alternative qualitative method can be proposed for validating LAI and biomass maps. Both LAI

and biomass maps revealed high value patterns in the offshore part of the Wanga Wallen Banks

(Figure 4.8, 4.9, 4.10.). According to the species maps of Phinn et al. (2008), Lyons et al. (2011)

and Roelfsema et al. (2014) this particular part of Wanga Wallen Banks was dominated by

Cymodocea serrulata, which is the largest species among the other five species in the area.

4.4. Conclusions, Perspectives and Future Work

In one aspect, this study demonstrated how in situ spectral measurement can be used to

examine the nature of the relationships between reflectance, and seagrass LAI and biomass. This

type of examination is difficult to achieve by only using image-based approaches due to

Page 103: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

83

atmospheric, water column and image spatial resolution issues. In this regard, LAI was strongly

correlated with green reflectance, whereas low correlations were found between biomass and

reflectance. In addition, the nature of the relationship between reflectance and vegetation where

green light is reflected the most compared with blue and red wavelengths, the more abundant green

light than blue and red lights at the study site may influence the strong relationship between LAI

and the green band. This is supported by the study of Borrego Acevedo (2014) who found a

significant relationship between microphytobenthos – whose spectral absorption characteristics are

similar to seagrass (see e.g. Figure 3.8. and Figure 3.10. of Chapter 3 ) – and the blue reflectance

blue-dominated waters of Heron Reef (Veal, 2010). This implies that the knowledge on underwater

light climate or water optical classification when establishing relationships between spectral

reflectance and the biophysical parameters of benthic habitats may be required. The laboratory

analysis in this study confirmed the significant allometric relationship between LAI and biomass,

which was reflected in the high correlation between LAI and biomass regardless of species,

community type and temporal variations. Overall, the analysis conducted at in situ and laboratory

stages in this study provided the basis for the subsequent stages conducted at image analysis level.

Our findings also showed that the image corrected using the bottom reflectance technique

(Rb) of Dierssen et al. (2003) produced more accurate estimates of seagrass LAI compared with the

single green band and the Lyzenga-transformed band. This implied the use of both down-welling

and up-welling diffuse attenuation coefficients to correct for water column effect on satellite

imagery as opposed to the use of the single diffuse attenuation coefficient commonly applied using

the Lyzenga method (Lyzenga, 1978). The importance of partitioning the upwelling and

downwelling diffuse attenuations for mapping benthic habitats, using remote sensing, was indicated

by Maritorena et al. (1994) though they suggested the single diffuse attenuation coefficient for

simplicity reasons. Another advantage of using the Rb technique is the preservation of the original

band(s), instead of outputting transformed band(s) or band index (e.g., Lyzenga, 1978, 1981). The

availability of original water-column corrected bands could accommodate the identification of

benthic habitats using increasingly available spectral libraries (e.g., Hochberg et al., 2003;

Fyfe., 2003).

Using regression technique between validation points and the resultant satellite-based maps,

the accuracies of LAI maps for June 2012 and February 2013 were 62% and 73%, respectively,

whereas the accuracies of biomass maps for the same timeframe were 71% and 60%. For the LAI

maps, the accuracies were higher than those from the only similar published study conducted by

Page 104: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

84

Wicaksono and Hafitz (2013) in the Karimunjawa Islands, Indonesia with accuracy of 20% and

26%, while for the biomass maps the achieved accuracies were in agreement with similar published

studies with the range of 57–83% (Armstrong, 1993; Mumby et al., 1997; Knudby and Nordlund,

2011 and Roelfsema et al., 2014).

The resultant LAI and biomass maps revealed consistent proportions of biomass and LAI

classes overtime in the Eastern Banks but variations occurred among individual banks, which were

likely due to the species-specific seasonal variation in shoot density (Preen, 1992; Kocak et al.,

2011). This inter-bank variation is in accordance with Roelfsema et al. (2012) suggesting that

seagrass in the Eastern Banks should be treated individually as four distinct seagrass environments

rather than a single continuous meadow.

The limitations of the present study and future work to address them are described

as follows:

• This study used in situ seagrass samples collected in an irregular sampling scheme (e.g.,

sporadic points in different individual banks) that may overlook the true spatial

characteristic of seagrass (e.g., LAI, biomass, cover, diversity, and density) in the study site.

While increasing the use of a seagrass core area larger than 0.08 m2, as used in this study,

would be technically challenging, a more systematic sampling scheme to represent

parameter(s) of interest at the satellite's pixel resolution by increasing the spatial

distributions of the cores or density of them may be recommended. However, this should

also take into account the trade-off between the cost-effectiveness of using remote sensing

methods and the logistical challenges of collecting large amounts of in situ data. Another

approach would be to use proxy parameter(s) such as shoot density or cover to estimate

biomass / LAI. Close to the completion of this thesis a publication from Roelfsema et al.

(2014) appeared that used multiple regression of cover and species to estimate biomass.

Their result is compared to this study's in Table 4.6 (p. 74) and shows very close similarity

in biomass for the whole Eastern Banks though varies for the other individual banks.

However, our in situ data analysis shows that LAI - biomass relationship is similar and

consistent at both community and species levels (Table 4.3, p.64). In the future it may be

useful to include cover and LAI to estimate biomass.

• The laboratory-based measurement of LAI is laborious, requiring the partition of seagrass

leaves/shoot per species and then measuring their surface area in the ImageJ photographic

software. Reducing the laboratory processing effort for LAI by using the in situ

Page 105: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

85

photographic method of Zhao et al. (2012) is thus recommended. At the time of its

publication, the method was suitable only for vertical upward-oriented photographs, which

would be rather challenging for seagrass underwater sampling techniques. The other

alternative technologies are using underwater video operated by a diver or snorkeler for

obtaining cover (McDonald et al., 2006), or the latest technology, AUV (autonomous

underwater vehicle) equipped with two underwater cameras for obtaining seagrass species

information (Roelfsema et al., 2015). The seagrass species composition information

obtained from the AUV method was used as validation data for image-based seagrass

species map (Roelfsema et atl., 2015). Once the correlation between cover and biomass /

LAI can be established and both the AUV and underwater video methods can be used to

derive cover, then the logistical burden of collecting seagrass cores for obtaining biomass /

LAI can be reduced.

• The generation of the water-column corrected bottom reflectance (Rb3) used in this study

required additional processing time using the additional commercial software of Hydrolight.

The use of open source software such as PlanarRad (www.planarrad.com) would be

beneficial for studies requiring in-water radiative transfer calculation. Alternatively,

applying physics-based method to remote sensing data (e.g., Brando, 2009; Dekker et al.,

2011) to retrieve water-column corrected bottom reflectance would be an option. The

application of this method would allow the direct inversion of remote sensing reflectance to

bathymetry, substrate reflectance/type and water quality parameters with minimum

additional inputs from other software(s) or ancillary data (i.e., diffuse attenuation

coefficients from radiative transfer software).

Page 106: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

86

Chapter 5: Application of Remote Sensing Data for Estimating Seagrass Productivity

Main Findings

• Process- and biomass-based methods were successfully applied to produce image-based

productivity maps with values that were still within the ranges of the compiled seagrass

literature values.

• Analysis using both process- and biomass-based methods showed that Wanga Wallen bank had

the highest productivity compared with the other banks, confirming the favourable

environmental conditions for seagrass growth in this bank reported by previous studies.

• Seasonal analysis revealed no significant differences in the average productivities in all banks

for winter and summer, probably due to the contrasting growth patterns between the dominant

seagrass species Zostera muelleri and the other species that cancelled out any seasonal pattern.

5.1. Introduction

In addition to its function as a nursery ground, for ecologically and economically important

fish and as a coastal filter for pollutants and suspended sediment (Short et al., 2006), the

contribution of seagrass habitats to carbon sequestration is significant due to the high rates of both

net primary production (NPP) and carbon burial (Duarte et al., 2010, 2011). With global coverage

of 0.6 x 106 km2, seagrass NPP can reach 0.49 Pg C · m-2 · year-1 (Mateo et al., 2006). As a

comparison, if the surface areas of marine phytoplankton and terrestrial forest are converted to

seagrass surface area, their NPP corresponds to 0.07 and 0.24 Pg C · m-2 · year-1, respectively

(recalculated from Mateo et al., 2006). Primary production data are commonly obtained using

in situ or in vivo measurements that vary from relatively simple methods such as biomass

accumulation, leaf marking, plastochron interval techniques and lepidochronological analysis, to

more complex metabolic methods such as oxygen evolution in open or closed systems and

carbon 14 tracer techniques (Erftemeijer et al., 1993; Short et al., 2001). Using these methods

seagrass production data have been extensively collected, encompassing different geographic areas,

environmental settings and species (review by Lee et al., 2007a). Field measurements are usually

challenging in term of time and cost and they cannot be used to conduct measurements over a

continuum of scales (Mumby et al., 2004). Remote sensing emerges as a promising tool to scale-up

field observations, providing synoptic, continuous and repetitive measurement over an area.

Page 107: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

87

Despite progress in coastal management and information gathering, there is insufficient

spatial assessment of seagrass ecosystems in terms of their physiological function such as

productivity. Production or productivity is likely to be an important indicator for the health of

photosynthetic organisms as the basic components in the process of metabolism, – respiration and

photosynthesis – both respond to environmental variables such as light, temperature and nutrient or

water chemistry (Duarte et al., 2008). With the global decline of seagrass due to anthropogenic

factors (Duarte et al., 2008) and the potential impact of climate change (Bjork et al., 2008) it is of

interest to establish a synoptic observation system of seagrass productivity with which the health of

the seagrass ecosystem can be monitored overtime. Such a system has been in operational use in the

form of satellite-based primary productivity maps for both terrestrial ecosystems

(http://edcdaac.usgs.gov/dataproducts.asp, http://modis-land.gsfc.nasa.gov/) and ocean surface,

(http://oceancolor.gsfc.nasa.gov/) which is mainly based on the application of bio-optical models to

satellite data.

To date, the application of similar systems to benthic habitat is still under development,

including for coral reefs (Atkinson and Grigg, 1984; Ahmad and Neil, 1994; Andrefouet and Payri,

2000; Brock et al., 2006; Hochberg and Atkinson, 2008; Moses et al., 2009) and seagrass (Dierssen

et al., 2010). The effort to scale-up coral reef gross and net productivity was pioneered by Atkinson

and Grigg (1984) using analogic photographs for French Frigate shoals, in Hawaii. Improvement,

using digital imagery, was later made by Ahmad and Neil (1994) for Heron Reef by producing

semi-quantitative productivity maps classified as low, medium and high. Andrefouet and Payri

(2000) then refined the previous methods by producing quantitative maps of reef metabolisms

(production, respiration and calcification) from SPOT images using published community (biotope)

metabolism values assigned to the satellite-based community map. Significant advancement was

later made by Brock et al. (2006) who compared the performance of three digital imageries (AISA,

ASTER and IKONOS) and used in situ metabolism values from a large portable incubation system

(SHARQ - The Submersible Habitat for Assessing Reef Quality) to scale-up coral reefs metabolism

in Northern Florida. Finally, Moses et al. (2009) then applied the method of Brock et al. (2006) to

the same area using Landsat ETM+.

These studies have been successful in mapping coral reef metabolism but as their mapping

units are communities (e.g., sand, seagrass, dense live substrate, etc.), instead of pixels, they cannot

accommodate metabolism variation within a particular community. This variation within the same

community is relevant for seagrass as the same seagrass community inhabiting different

Page 108: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

88

environments (e.g., bay site, river mouth or reef flat) showed different productivity rates (Moriarty

et al., 1990; Rasheed et al., 2008). Hochberg and Atkinson (2008) addressed this issue by applying

a light-use efficiency (LUE) model on a pixel basis to IKONOS imagery using the parameters of

incident light (Edλ), absorbed light by community (Aλ) and ɛ, photosynthetic efficiency converting

absorbed light to oxygen evolution. The use of LUE model to estimate productivity using remote

sensing has been established for terrestrial vegetation (review by Hilker et al., 2008). In this regard,

the absorbed light parameter (commonly termed as FPAR in the terrestrial remote sensing texts)

was modelled either empirically, from spectral vegetation indices, or analytically using complex

physical models (Hilker et al., 2008). Likewise, rigorous methods were implemented to vary

assumed maximum ɛ according to environmental stress (e.g., using temperature parameters) or

direct empirical relationships between ɛ and spectral indices that quantify photo-protective

mechanisms (e.g., Photochemical Reflectance Index or PRI) (Hilker et al., 2008).

For aquatic organisms, however, ɛ (also often symbolised as α or ϕ in aquatic biology or

ecology texts) is commonly obtained from a photosynthesis–irradiance (P-I) experiment setting

where there is a light level when photosynthesis saturates (termed as Pmax, maximum

photosynthesis) and then decreases beyond this light level due to photo inhibition (Kirk, 2010). The

use of aquatic ɛ in the LUE model, without the inclusion of its accompanying Pmax parameter,

implies that productivity can increase without bound with light. It was also observed that ɛ would

vary with decreasing light in coral (Chalker et al., 1983) and in seagrass (Dawes, 1998; Campbell et

al., 2007) as a photo-adaptive mechanism. Additionally, the absorbed light parameter (Aλ) in

Hochberg and Atkinson (2008) calculated using the formula of 1 - Rλ (where Rλ = reflectance) may

have overestimated the true absorbed light. While it is understood the main light component process

in coral is the absorption by symbiotic algae, and reflectance and scattering by coral skeletons

(Stambler and Dubinsky, 2005; Enriquez et al., 2005), at the scale of remote sensing there may be

light transmission components from sand (Kuhl and Jorgensen, 1994), seagrass (Durako et al.,

2007) or algae (Haxo and Blinks, 1950) that also influence the total light absorption.

In regard to seagrass, Dierssen et al. (2010) have attempted to estimate seagrass NPP in the

Great Bahama Bank using SeaWIFS imagery (1 km resolution) by empirically relating bottom

reflectance at 555 nm to in situ NPP. Dekker (1993) stated that spurious results might occur with

the empirical method as a causal relationship may not exist between reflectance and the parameter

of interest. Additionally, extending the empirical method to other areas should be conducted with

caution, as the applicability of this method may be site-specific.

Page 109: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

89

The present study aimed to demonstrate the estimation of seagrass above-ground gross

primary productivity (hereafter referred to as productivity) using remote sensing by considering

biophysical parameters and physiological processes that could affect seagrass productivity. Mainly,

the objectives of the present chapter were to (a) estimate seagrass productivity using a productivity

model applied to remote sensing data that incorporated seagrass physiological parameters and

environmental controlling factors, (b) estimate seagrass productivity using a biomass-based

productivity model applied to remote sensing data, and (c) perform seasonal seagrass productivity

analysis using the resultant image-based productivity maps.

5.2. Data and Methodology

5.2.1. Outline.

The present chapter uses the outputs obtained in the previous chapters in the form of satellite

image-based PAR, LAI and biomass maps to produce seagrass productivity maps. The productivity

maps were generated using two methods: (a) a method that takes into account incident light, depth,

maximum productivity, canopy light attenuation, light-saturation threshold for photosynthesis and

photosynthetic efficiency (hereafter referred to as the process-based method), and (b) a method that

converts seagrass biomass to a productivity rate (hereafter referred to as the biomass-based

method).

5.2.2. Image data and processing.

Satellite-based maps of PAR, LAI and biomass produced in the previous chapters were used

in the present chapter to calculate seagrass productivity. In short, the bottom PAR maps (PAR maps

of the depth of seagrass) were processed using the implementation of the exponential light

attenuation equation of Beer’s law applied to surface MERIS PAR, HICO KdPAR and WorldView-

2-based bathymetry data. LAI and biomass maps were produced using a series of processes that

combined field-, laboratory- and remote sensing-based approaches. Briefly, the LAI map was

generated using a model developed from the regression analysis between the depth-corrected green

band and in situ LAI. The use of the green band was based on the correlation analysis between in

situ reflectance and LAI data. The biomass map was then generated from the LAI map using the

allometric relationship between LAI and biomass developed from in situ data.

Page 110: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

90

5.2.3. Process-based productivity.

Two process-based productivity algorithms were used in the present chapter: the algorithm

that includes several seagrass physiological parameters and environmental controlling factors and

the algorithm that mainly relies on the light parameter. The algorithms were selected based on the

relevancy of their parameters for estimating seagrass productivity and of their suitability for map

analysis using remote sensing data. The first algorithm was modified from Charles-Edwards et al.

(1986) to give the following form:

GPP = (α + (tanh d)) I Pmax[1 − exp (−kc LAI]/((α + (tanh d)) kc I + Pmax)

(in µmol O2 g−1 · dry wt · hr−1) (1)

where α = photosynthetic efficiency, tanh = hyperbolic tangent function, d = depth, I = incident

irradiance (PAR), Pmax = maximum productivity, kc = canopy light attenuation and LAI = leaf area

index.

The inputs of d, I and LAI were from the corresponding satellite-based maps produced in the

previous chapters, whereas α, Pmax and kc were derived from literature values. The main differences

between equation 1 and the original formula of Charles-Edwards et al. (1986) were the omission of

parameter Rc (canopy dark respiration) as the present chapter aimed to estimate gross productivity

instead of net productivity, and the addition of the term (α + (tanh d)) to vary the photosynthetic

efficiency parameter (α) according to decreasing irradiance (light). Previous researchers have

observed that seagrass (Dawes, 1998; Campbell et al., 2007) and coral (Chalker et al., 1983)

showed their photo-adaptive capabilities to low irradiance via increased photosynthetic efficiencies.

Here, depth was used as a proxy for irradiance (light) as light decreases through depth. The

hyperbolic tangent function was employed to vary α according to depth following the commonly

used equation of Jassby and Platt (1976) that described the basic photosynthesis vs irradiance

relationship using hyperbolic tangent function. The use of the tanh function will set a maximum α

value at the depth of 8 m and beyond this depth, the value of the maximum α will be used. This

condition was assumed acceptable for the study site as analysis in the previous chapter (section

3.3.4 of Chapter 3) showed that the maximum seagrass colonisation depth in the Eastern Banks was

around 3.31 m. This value was calculated using the formula of Duarte (1991) that established the

relationship between the maximum seagrass colonisation depth and light attenuation coefficient in

the water column. Phinn et al., (2008) stated that 82 % of Eastern Banks area was located at 3 m or

Page 111: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

91

shallower depth, while Longstaff (2003) observed that the maximum depth limit (MDL) of Zostera

capricorni in the eastern region of Moreton Bay, which is Eastern Banks, was close to 3 m.

In regard to the parameter of canopy light attenuation (kc), Enriquez et al. (2005) and Perez-

Llorens and Niell (1993) in Plus et al. (2001) observed that the kc of Thallasia testudinum was

remarkably higher (2.1–11.5) than that of Zostera noltii (0.97). An Australian study by Carruthers

and Walker (1997) observed kc values of 0.93 (summer) and 0.44 (winter) for Amphibolis griffithii.

This kc variation was likely due to the leaf morphological differences among species and local

environmental condition (e.g. turbidity). Considering that the leaf morphology of the dominant

species in the Eastern Banks (Zostera muelleri) bears a greater resemblance to Zostera noltii than to

Thallassia testudinum and Amphibolis griffithii, the kc value of Zostera noltii (0.97) (Perez-Llorens

and Niell, 1993 in Plus et al., 2001) was used.

The second process-based algorithm used was the formula describing the relationship

between photosynthesis and irradiance from Jassby and Platt (1976):

GPP = 𝑃𝑚𝑎𝑥 (𝐼

𝐼𝑘−tanh(𝑑)) (in µmol O2 g−1 · dry wt · hr−1) (2)

where Ik = Light-saturation threshold for photosynthesis. Again, Ik was set to vary with depth as the

photo-adaptive capability of seagrass to low irradiance took the form of decreasing Ik (Dawes,

1998).

In regard to temperature, a comprehensive review by Lee et al. (2007) on the effects of

irradiance (light), temperature and nutrients on seagrass growth showed that although the seasonal

growth patterns of seagrass were regulated by irradiance and temperature, the separation of these

effects is difficult due to correlation between the two factors. For the case of the study site (Moreton

Bay) Preen (1992) attempted to correlate temperature and light to seagrass abundance (biomass and

shoot density) and observed similar result to that of Lee et al. (2007) as irradiance largely

determines water temperature (Preen, 1992).

5.2.4. Biomass-based productivity.

The seagrass biomass-based productivity map was generated using the formula of Duarte

and Chiscano (1999) that related seagrass biomass to productivity. The form of the formula is:

𝐺𝑃𝑃 = 0.1 𝑥 𝐵0.64 (𝑖𝑛 𝑔 · 𝑑𝑟𝑦 𝑤𝑡 · 𝑚−2 · 𝑑𝑎𝑦−1) (3)

Page 112: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

92

Here GPP and B are productivity and biomass (above-ground), respectively. The input for B

was from the satellite-based biomass map produced in Chapter 4 using WorldView-2 images of 12

June 2012 and 13 February 2013. The formula was developed from a cross-species global review of

published seagrass literature where five of the six seagrass species present in the Eastern Banks

(Cymodocea serrulata, Halodule uninervis, Syringodium isoetifolium, Halophila ovalis, Zostera

muelleri, except Halophila spinulosa) were also represented in the analysis (Duarte and Chiscano,

1999). Similar associations between seagrass productivity and biomass have been confirmed for

species Thallasia testudinum (Pantoja Rejes, 2003 in Hedley and Enriquez, 2010).

5.2.5. Seasonal analysis.

Seasonal analysis was conducted only using the biomass-based productivity maps of winter

2012 and summer 2013, as the satellite-based PAR map required to produce the process-based

productivity map for the winter season was compromised by high cloud cover. The analysis was

applied to the whole Eastern Banks and the five individual banks (Amity, Chain, Maroom, Moreton

and Wanga Wallen banks) (Figure 5.1.) by comparing the surface areas of productivity classes used

(e.g., low, medium, high) between winter (June 2012) and summer (February 2013).

Figure 5-1. ALOS AVNIR-2 (14 July 2010) coverage of the study area in the Eastern Banks, Australia, with the names and boundaries of the individual banks indicated.

Page 113: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

93

5.3. Results and Discussion

5.3.1. Process-based seagrass productivity.

5.3.1.1. Literature estimates of seagrass maximum productivity (Pmax), light-saturation

threshold for photosynthesis (Ik) and photosynthetic efficiency (α).

For the purpose of implementing the process-based productivity, the photosynthetic

parameters (Pmax, Ik, and α) of seagrass were compiled from published seagrass literature. The three

required parameters can only be obtained from metabolic methods such as oxygen evolution in open

or closed systems, as other methods (e.g., leaf marking techniques or rhizome tagging methods)

commonly produced only the productivity or production rate (in gram dry weight or carbon per area

per specific time) without Pmax, Ik, and α parameters. However, even using the same or similar

metabolic methods researchers often used different units to express the photosynthetic parameters

(Lee et al., 2007a). The literature estimation of Pmax and α in this chapter only considered studies

that expressed both parameters in μmol O2 g-1 · dry weight · hr-1, extracted and modified mostly

from Lee et al. (2007a). The abundant available data on metabolic parameters from PAM -

fluorometry measurements (e.g. Ralph and Burchett, 1995; Ralph and Gademann, 2005) were not

used here as "... there is currently no clear quantitative link between relative fluorescence and rates

of photosynthesis..." (Hochberg and Atkinson, 2008). It was found that most of metabolic

parameters from PAM-fluorometry measurements were unitless and therefore have limitation to be

used as a link for irradiance and production (e.g. Hochberg and Atkinson, 2008).

According to Preen (1992), there was no significant difference in the seagrass productivity

rate between temperate and subtropical/tropical regions, which was also reflected in the seagrass

productivity reviews by Hillman et al. (1989) and Lee et al. (2007a). In their reviews, some of the

highest and lowest production rates occurred both in temperate and subtropical or tropical regions.

Additionally, Lee et al. (2007a) also observed that seagrass productivity was species-specific and

influenced by light, temperature and nutrients, implying the influence of geographic factors on

seagrass productivity, as species and environmental factors may vary geographically.

Unfortunately, no published works were available that specifically measured in situ or in vivo

seagrass productivity for species from Moreton Bay or the Eastern Banks that resulted in the

parameters of Pmax, Ik, and α expressed in the required unit (μmol O2 g-1 · dry weight · hr-1). In

considering the absence of the required seagrass photosynthetic parameters from the study site, the

influence of species-specific and environmental factors on productivity (Lee et al., 2007a) and the

latitudinal similarities of seagrass productivity (Hillman et al., 1989; Lee et al., 2007a), the

Page 114: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

94

compilation of the required photosynthetic parameters was subsequently based on the seagrass

bioregional model of Short et al. (2007). In their work Short et al. (2007), developed bioregional

models to categorise global seagrass distribution and diversity into six bioregions based on major

oceans, climate and species assemblages: temperate North Atlantic, tropical Atlantic,

Mediterranean, temperate North Pacific, tropical Indo-Pacific and temperate Southern Oceans. The

lists of seagrass species present in these six bioregions were also provided in Short et al. (2007).

According to this categorisation the Eastern Banks belongs to the bioregion of the tropical Indo-

Pacific and the photosynthetic parameters (Pmax, Ik, and α) of seagrass species listed to this bioregion

were then compiled (Table 5.1.).

Table 5-1. Photosynthetic Parameters (Pmax, Ik, and α) Extracted and Modified from Lee et al. (2007a)

Note: Values in parentheses are mid-points (median) values for measurements given in a range of values. a = calculated from the relation 𝐼𝑘 = 𝑃𝑚𝑎𝑥

𝛼 (Dunton and Tomasko, 1994; Pollard, 1999). b = visually inspected from graphic

Table 5-2. Descriptive Statistics for Pmax, Ik, and α Given in Table 5.1.

Note. The mean values were used for producing the satellite-based seagrass productivity using the process-based methods

Species Tissue Pmax Ik α Original Sources(μmol O2 g

-1 dw h-1) (μmol m2 s-1) (μmol O2 g-1 dw h-1)

Ruppia maritima Above-ground 426.3 1.922 Enríquez et al. (1995)Zostera carpriconi Above-ground 182 Flanigan and Critchley (1996)

Above-ground 179.4 195 0.92a Schwarz (2004)Above-ground 167.5 242 0.69a Schwarz (2004)

Cymodocea rotundata Above-ground 169.7 – 261.3 (215.5) 115a 0.625 – 3.125 (1.875) Agawin et al. (2001)Cymodocea serrulata Whole plant 200b Hena et al. (2001)

Whole plant 200b Hena et al. (2001)Enhalus acoroides Above-ground 40.9 – 196.3 (118.6) 0.313–6.250 (3.28) Agawin et al. (2001)Halodule wrightii Whole plant 456 – 651 (553.5) 245–429 (337) 1.5–2.3 (1.9) Dunton (1996)

Whole plant 203 – 652 (427.5) 147–652 (399) 0.6–2.2 (1.4) Dunton (1996)Whole plant 140 - 1104 (622) 189–453 (321) 0.5–2.4 (1.5) Dunton and Tomasko (1994)Whole plant 441 349 1.3 Dunton and Tomasko (1994)

Above-ground 421 101 4.2 Dunton and Tomasko (1994)Thalassia hemprichii Above-ground 141.3–330.3 (235.8) 0.02–0.17 (0.095) Agawin et al. (2001)

Thalassodendron ciliatum Above-ground 165.6 Parnik et al. (1992)

Descriptive Statistics Pmax (n = 12) Ik (n = 11) α (n = 11)

(μmol O2 g-1 dw h-1) (μmol m2 s-1) (μmol O2 g

-1 dw h-1)Min 118.600 101.000 0.095Max 622.000 399.000 4.200

Median 328.400 200.000 1.500Mean 331.142 240.091 1.735Stdev 169.858 98.321 1.155

Page 115: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

95

Table 5.1. and Table 5.2. show a relatively large variability in Pmax, Ik, and α values that

reflect differences in the methods used, time of measurement, environmental setting (e.g. river

mouth or reef flat) and the tissues measured. Productivity and its photosynthetic parameters often

show seasonal variability, with typical peaks in summer and lows in winter including in tropical

areas, although the seasonality is less-pronounced than in temperate areas (Preen, 1992; Lee et al.,

2007a). Moriarty et al. (1990) observed that the same seagrass community (a mixture of different

seagrass species) inhabiting different habitats (bay site, river mouth and reef flat) in the Gulf of

Carpentaria, Australia also showed different productivity rates. A similar result was found by

Rasheed et al. (2008), in Torres Strait, Australia, who observed that the intertidal community had

higher productivity than that of the sub-tidal community.

In applying the LUE model to coral reef communities using IKONOS satellite imagery,

Hochberg and Atkinson (2008) assumed that photosynthesis did not saturate (e.g., there were no

Pmax or Ik values) at community level, as the complex 3-D structure of a branching coral-dominated

reef community may increase light-capturing capacity and thus prevent photosynthesis saturation

(Smith, 1981). However, measurements of seagrass communities in the field using benthic chamber

methods showed consistent photosynthesis saturation at high light levels for Zostera noltii (Clavier

et al., 2011, 2014; Ouisse et al., 2011), Zostera marina (Ouisse et al., 2011) and Cymodocea nodosa

(Clavier et al., 2014). This may be due to the high light requirements of seagrass, restricting

seagrass habitat to a light-abundant environment that is also prone to occasional or regular

excessive light exposure. In regard to seagrass modelling, this condition necessitates the use of Pmax

and Ik parameters to model seagrass productivity, which is attempted in the present chapter.

5.3.1.2. Output of process-based seagrass productivity analysis.

The average value of seagrass productivity rates extracted from the satellite-based seagrass

productivity maps for the whole of the Eastern Banks and individual banks are given in Table 5.3.

In general, the average productivity values calculated using equation 1 were higher than those

calculated using equation 2, which reflected the different parameterisation of the productivity

models used. Assessing the quantitative accuracy of these average values, including which

algorithm is more accurate, is difficult as productivity measurements in the study site using

metabolic methods expressed in the same units are not available.

Page 116: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

96

Table 5-3. Average Seagrass Productivity Values of Individual Banks

Note. Calculated using process-based methods of equation 1 (modified from Charles-Edwards et al., 1986) and equation 2 (Jassby and Platt, 1976) applied to WorldView-2 satellite of February 2013

However, qualitatively speaking, the average values in Table 5.3. are close to the mean Pmax

value in Table 5.2 (331.142 μmol O2 g-1 · dry wt · hr-1) obtained from literature values and are still

within the range of the compiled Pmax values (Table 5.1.). The maximum Pmax value in Table 5.1.

(1104 O2 g-1 · dry wt · hr-1) is also close to the maximum values of the satellite-based productivity

maps calculated using equation 1 (1033 O2 g-1 · dry wt · hr-1) and equation 2 (1084 O2 g-1 · dry wt ·

hr-1). One notable feature is that the highest average productivity rates calculated using both

algorithms were observed for the Wanga Wallen Bank (Table 5.3.).

Roelfsema et al. (2014) conducted multi-temporal mapping of seagrass cover in the Eastern

Banks from 2004–2013 using semi-automated object-based image analysis applied to high

resolution imageries (Quickbird, IKONOS and WorldView-2) and observed that Wanga Wallen

Banks had a higher seagrass percentage cover than the other banks during the period studied. Our

result of biomass mapping using WorldView-2 imageries of June 2012 (winter) and February 2013

(Table 4.6 Chapter 4) also shows that Wanga Wallen bank has the highest average biomass (winter

= 100 g · dry wt · m-2, summer = 102 g · dry wt · m-2) compared with the other banks. It is likely

that the high seagrass productivity in the Wanga Wallen banks drives the high biomass in the area,

as seagrass primary production is partly biomass dependent (Preen, 1992) and shows similar, but

more exaggerated, seasonal patterns than biomass (Hillman et al., 1989). Hillman et al. (1989) used

similar linear relationships between seagrass cover and biomass to estimate seagrass primary

production in the Swan–Canning Estuary in Western Australia.

Although the analysis using average values is useful, the nature of its point-based

measurement cannot reveal spatial information for seagrass productivity in the study site, for which

Process-Based 1 Process-Based 2Location (Equation 1) (Equation 2)

(μmol O2 g-1 dw h-1) (μmol O2 g

-1 dw h-1)Eastern Banks 289 205

Amity 284 201Chains 139 98

Maroom 174 118Moreton 336 249

Wanga Wallen 493 341

Average Values

Page 117: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

97

the remote sensing approach provides the solution. Figure 5.2. shows satellite-based seagrass

productivity maps calculated using the process-based methods (equation 1 and 2). To allow for

consistent and easy comparison between the two maps produced using different algorithms, five

productivity classes were used (in μmol O2 g-1 · dry wt · hr-1): class 1 (0–216.80), class 2 (216.81–

433.61), class 3 (433.62–650.41), class 4 (650.42–867.22), and class 5 (867.23–1084.02). The

surface area (in km2) of each class for the whole of the Eastern Banks and individual banks was

then extracted from the maps and presented graphically in Figure 5.3. Figure 5-2. Maps of classified seagrass process-based productivity calculated using equation 1 (modified from Charles-Edwards et al., 1986) (left) and equation 2 (Jassby and Platt, 1976) (right). The maps were produced from the summer image data set (e.g., the LAI input map was processed from WorldView-2 image of 13 February, 2013 and the PAR map was from the average of summer months of 2003–2012).

(O2 g-1 dw h-1) Land Deep waters Shallow waters Cloud/turbidity

Page 118: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

98

Figure 5-3. Graphics of total surface area of productivity classes in each individual banks calculated using the process-based methods. Process-based 1 was calculated using equation 1 (modified from Charles-Edwards et al., 1986) and process-based 2 was based on eq uation 2 (Jassby and Platt, 1976).

.

In general, the two process-based methods revealed the presence of all productivity classes

in all the banks. In Figure 5.3., in some banks class 4 and 5 may not be clearly visible due to the

small size of the areas present but the surface areas of those classes are different between the two

methods. Productivity class 1 (0–216.80 O2 g-1 · dry wt · hr-1 ), class 2 (216.80–433.61 O2 g-1 · dry

wt · hr-1), and class 3 (433.61–650.41 O2 g-1 · dry wt · hr-1) were dominant in all banks, with

0–216.80 216.81–433.61 433.62–650.41 650.42–867.22 867.23–1084.02 Productivity Classes (O2 g-1 dw h-1)

Page 119: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

99

slightly higher surface areas in class 4 (650.41–867.22 O2 g-1 · dry wt · hr-1) and class 5 (867.22–

1084.02 O2 g-1 · dry wt · hr-1) observed in the Moreton and Wanga Wallen banks. Based on process-

based 1 method (equation 1) Wallen banks had the largest area of the highest productivity class

(class 5: 867.22–1084.02 O2 g-1 · dry wt · hr-1) and it had the second largest area of class 5

productivity after Moreton Banks, based on the process-based 2 method (equation 2).

These results were in agreement with the previous analysis using the average productivity

values of individual banks (Table 5.3.) showing that although covering a small area, Wanga Wallen

bank had higher productivity than the other larger banks. The analysis using classified satellite-

based productivity maps revealed the spatial heterogeneity of seagrass productivity, even within the

same bank, which may not be resolvable using field measurements. In this regard, if the clustered

spatial pattern of seagrass productivity was consistent through time, the result of averaging field

measurements of seagrass productivity at one point may not represent the average seagrass

productivity for the whole, or a larger area. Further discussion on the possible factors influencing

the spatial variation in seagrass productivity in the study site will be provided in the end of the next

section as these factors may also affect productivity estimated using the biomass-based method.

5.3.2. Biomass-based seagrass productivity and seasonal analysis

Table 5.4. shows that the average seagrass productivity values of the whole of the Eastern

Banks were relatively stable for June 2012 (winter) (1.578 g · dry wt · m-2 · day-1) and February

2013 (summer) (1.547 g · dry wt · m-2 · day-1). Temporal variations in productivity were observed

in the individual banks (Table 5.4.) but their differences were negligible.

Page 120: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

100

Table 5-4. Average Seagrass Productivity Values of Individual Banks

Note. Calculated using biomass-based method (equation 3; Duarte and Chiscano, 1999) applied to WorldView-2 images of 12 June, 2012 and 13 February, 2013

Table 5-5. Literature Estimates of the Productivity of the Five Seagrass Species Present in The Eastern Banks

Note. Species Halophila spinulosa not included. Values in parentheses are mid-points (median) values for measurements given in a range of values

ProductivitySpecies (g dw m-2 d-1) Condition / location Methods Sources

0.5 New South Wales (NSW), Australia Leaf marking technique West and Larkum (1983)0.69 Moreton Bay, Australia, 10 % Incident light Leaf marking technique Hansen et al. (2000)1.7 Moreton Bay, Australia, 50 % Incident light Leaf marking technique Hansen et al. (2000)

Zostera Capricorni 3.5 Moreton Bay, Australia Plastochrone interval technique Moriarty et al. (1985)3.6 NSW, Australia Leaf marking technique McComb et al. (1981) in West and Larkum (1983) 3.7 Wanga Wallen, Australia (Control) Rhizome tagging technique Udy and Dennison (1997)4.1 Botany Bay, NSW, Australia Leaf marking technique Larkum et al. (1984) in Preen (1992)1.2 Wanga Wallen, Australia (Control) Rhizome tagging technique Udy and Dennison (1997)

Cymodocea serrulata 0.4-2.8 (1.6) Gulf of Carpentaria, Australia (Bay site), Seasonal Diel curve procedure Moriarty et al. (1990)5 Papua New Guinea Leaf marking technique Brouns (1987)

8.3 Gulf of Carpentaria, Australia (open bay) Diel curve procedure Pollard and Moriarty (1991)1.2 Wanga Wallen, Australia (Control) Rhizome tagging technique Udy and Dennison (1997)

Halodule uninervis 2 Green Island, Australia Rhizome tagging technique Udy et al. (1999)3.8 Papua New Guinea Leaf marking technique Brouns (1987)

2.5-9.7 (6.1) Gulf of Carpentaria,Australia (river mouth), Seasonal Diel curve procedure Moriarty et al. (1990)Halophila ovalis 1.9 Swan/Canning estuary, Western Australia Leaf marking technique Hillman and McComb (1988) in Preen (1992)

Syringodium isoetifolium 3.6 Gulf of Carpentaria, Australia (open bay) Diel curve procedure Pollard and Moriarty (1991)0.8-22.5 (11.7) Gulf of Carpentaria, Australia (bay site), Seasonal Diel curve procedure Moriarty et al. (1990)

Location June 2012 (Winter) February 2013 (Summer)

Eastern Banks 1.578 1.547Amity 1.649 1.551Chains 1.731 1.496

Maroom 1.830 1.598Moreton 1.316 1.465

Wanga Wallen 1.865 1.840

Average Biomass-based Productivity(g dw m-2 d-1)

Page 121: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

101

Table 5-6. Literature Estimates of the Productivity of Seagrass Communities Containing One or More Seagrass Species Present in The Eastern Banks

Note. Values in parentheses are mid-points (median) values for measurements given in a range of values

The range of biomass-based seagrass productivity values of June 2012 (0.026–6.425 g · dry

wt · m-2 · day-1) and February 2013 (0.016– 4.024 g · dry wt · m-2 · day-1) were still within the

ranges published in most of the literature for values of seagrass productivity at species (Table 5.5.)

and community (Table 5.6.) levels. Similar to the compiled literature estimates of the

photosynthetic parameters (Table 5.1.), variations in the literature estimates of seagrass productivity

were also subject to differences in the methods used, environmental settings and the time of

measurement. Here, productivity measurements using the diel curve procedure and lacunal gas

technique (Moriarty et al., 1990; Pollard and Moriarty, 1991) tended to result in higher values than

the other techniques (Table 5.5. and Table 5.6.). Similar to the previous result of using the process-

based methods, the result of the biomass-based method also placed Wanga Wallen as the bank with

the highest seagrass productivity for both June 2012 and February 2013 (Table 5.4.). This is likely

due to the good water quality of Wanga Wallen banks allowing sufficient light penetration for

seagrass survival (e.g., Abal and Dennison, 1996) and the sheltering by North Stradbroke Island

from tidal flushing (Lyons et al., 2011). Lyons et al. (2011) observed that this sheltering effect

enhanced seagrass cover and macroalgae in Wanga Wallen banks and in Moreton banks, sheltered

Community Major Species Productivity Condition / location Methods Sources(g dw m-2 d-1)

1 Thalassia hemprichiiThalassodendron ciliatumHalophila ovalisSyringodium isoetifolium 1.882 Torres strait, Australia, Intertidal Conversion from biomass Rasheed et al. (2008)Halophila spinulosa 0.161 Torres strait, Australia, subtidal > density > productivity Rasheed et al. (2008)Cymodocea serrulataHalodule uninervisEnhalus acoroides

2 Thalassia hemprichii 7.2Cymodocea rotundata Makassar, Indonesia Bell jar method Erftemeijer et al. (1993)Halodule uninervisEnhalus acoroides

3 Cymodocea rotundata 4.2-6.9 (5.6) Puttalam Lagoon, Sri Lanka Benthic chamber Johnson and Johnstone (1995)Enhalus acoroides

4 Enhalus acoroides 11.1-25 (18.1)Syringodium isoetifolium Gulf of CarpentariaCymodocea serrulata Australia Lacunal gas technique Pollard and Moriarty (1991)Thalassia hemprichiiCymodocea rotundata

5 Syringodium isoetifolium 9.2-25.8 (17.5) Gulf of CarpentariaCymodocea serrulata Australia (Bay site)Halodule uninervisThalassia hemprichii 7.5-11.1 (9.3) Gulf of Carpentaria Lacunal gas technique Moriarty et al 1990Cymodocea rotundata Australia (river mouth)

9.2-23.3 (16.3) Gulf of CarpentariaAustralia (reef flat)

Page 122: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

102

by the southern tip of Moreton Island. In terms of the parameters used to model seagrass

productivity either using the process- or biomass- based method, these favourable environmental

conditions would allow sufficient light availability to seagrass and enhance leaf area index (LAI)

and biomass.

Figure 5.4. shows satellite-based seagrass productivity maps of June 2012 and February

2013 calculated using the biomass-based method (equation 3) and classified into five classes with

equal ranges (in g · dry wt · m-2 · day-1): class 1 (0–0.90), class 2 (0.90–1.80), class 3 (1.80– 2.70),

class 4 (2.70–3.60) and class 5 (3.60–6.42). The surface areas (in km2) of each class for the whole

of the Eastern Banks and the five individual banks are presented graphically in Figure 5.5. Except

for Moreton banks, the temporal change of the productivity classes between June 2012 and

February 2013 were minor, suggesting overall small seasonality in seagrass productivity between

winter and summer. Similar to the case with the process-based method, the analysis using classified

image-based productivity provided new insight on the “spatial condition” of seagrass productivity,

which may not be observable using single average productivity values (e.g., Table 5.4.). This is well

exemplified with the case of Moreton Banks where the winter–summer average productivities were

only slightly different (winter = 1.316 g · dry wt · m-2 · day-1, summer = 1.465 g · dry wt · m-2 ·

day-1), but more notable difference was observable in the winter and summer classified productivity

maps (Figure 5.4. and Figure 5.5.).

The non-seasonality of seagrass dynamics in Moreton Bay were observed by Young and

Kirkman, (1975), Young (1978) and Boon (1986) and were later updated by Preen (1992) using

seasonal seagrass data to confirm significant seasonal variation in several seagrass parameters (e.g.,

shoot density, biomass, shoot length). The winter–summer condition of the seagrass productivity

observed using satellite analysis in this study is likely due to the combined effect of the different

growth patterns between species Zostera muelleri and the other species inhabiting the Eastern

Banks (Halophila spinulosa, Halophila ovalis, Halodule uninervis, Cymodocea serrulata, and

Syringodium isoetifolium). Preen (1992) observed that the dominant species Zostera muelleri had a

winter–spring growth period, deviating from the typical summer–autumn growth period of the other

species. He also discovered that when Zostera muelleri dominated a particular site, the pooling of

different species in that site cancelled out any seasonal pattern in seagrass parameters. In this

regard, seagrass productivity may not decrease in June (winter) as this is the growing season for the

dominant Zostera muelleri and it may not reach its growth peak in February (summer) as the

maximum growth of the other species may still occur later in autumn. It can be observed from

Page 123: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

103

Roelfsema et al. (2014), who conducted multi-temporal analysis of seagrass species, biomass and

cover using high resolution images from 2004–20013 that in regard to the species composition, the

proportions of species Zostera muelleri and the other species were nearly equal for the Eastern

Banks between June 2012 and February 2013 (Figure 7 of Roelfsema et al., 2014). They also

concluded that the species composition in the Wanga Wallen bank (Zostera muelleri, Cymodocea

serrulata, Halophila ovalis, Syringodium isoetifolium) was the most consistent over time, which –

assuming the contrasting growth between Zostera and non Zostera species – was reflected in the

consistent winter–summer average productivity values (Table 5.4.) and productivity classes

(Figure 5.4. and Figure 5.5.). Further research is required to investigate the seasonality of seagrass

dynamics in Moreton Bay and the Eastern Banks, for example, by extending the study period from

the winter–summer period to the spring–autumn period or variations of both periods.

Figure 5-4. Maps of classified seagrass productivity of June 2012 (left) and February 2013 (right) estimated using biomass-based method (equation 3).

Land Deep

Shallow

(g d wt m-2 d-1)

Page 124: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

104

Figure 5-5. Graphics of total surface area of productivity classes in each individual banks calculated using the biomass-based methods for June 2012 (winter) and February 2013 (summer).

To give an insight on the effects of depth and PAR to the calculated productivity, regression

analysis between depth and PAR, and productivity (Table 5.7) were conducted on the selected

transect traversing seagrass area with various depths (Figure 5.6).

0–0.90 0.91–1.80 1.81–2.70 2.71–3.60 3.61–6.43

Productivity Classes (g d wt m-2 d-1)

Page 125: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

105

Figure 5-6. Transect line traversing seagrass area in the study site plotted on the image-based depth map of WorldView-2 (left) with its corresponding depth shown in the right side panel. Seagrass area of June 2012 from Roelfsema et al. (2014) is bounded by white line. Solid gray areas show land masses.

Table 5-7. The result of Pearson correlation analysis between productivity, PAR and depth. All values are significant at p < 0.05. The productivity, PAR and depth values included in the analysis were from transect line depicted in figure 5.6.

Table 5.7. shows that although depth significantly effects the process-based productivity, the

contribution of bottom PAR is much higher. The result also emphasises the importance of Kd in

driving the process-based productivity model as the bottom PAR parameter was derived using depth

and Kd parameters. If there was no contribution from Kd at all, the correlation coefficients of PAR

and depth would have been the same. The result of the correlation analysis supports the use of the

previously published light-based productivity models (e.g. Fourqurean and Zieman., 1991;

Zimmerman et al., 1994; Plus et al., 2001; Burd and Dunton., 2001) which may be suitable for

seagrass meadows characterised by shallow and clear environment such as Eastern Banks. On the

contrary, biomass-based productivity does not significantly correlates with PAR and depth (Table

5.7), which is reasonable due to the empirical nature of the model, directly converting biomass to

productivity. According to Preen (1992) measures of seagrass primary production are partly

biomass dependent, implying that other factor(s) could have more influences on productivity. The

A

Meter

B

Pearson Corr. R2 Pearson Corr. R2

Process-based prod 0.936 0.876 -0.785 0.617Biomass-based prod -0.236 0.056 0.464 0.215

PAR Depth

Page 126: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

106

type of substrate and sediment was also suspected to be one of the environmental factors affecting

seagrass dynamics in the Eastern Banks (Roelfsema et al., 2014). However, an in-depth study of the

sedimentary framework of Moreton Bay, using dredging samples from the sea floor by Maxwell

(1970), revealed that the substrate of the major part of the Eastern Banks, including its individual

banks, consisted of homogeneous non-carbonate clean sand facies with < 1% mud (Figure 6 and 9

in Maxwell, 1970). Lee et al. (2007a) stated that seagrass productivity was influenced by three main

factors: nutrition, light, and temperature, with the effects of the last two factors often difficult to

separate due to their high correlation (see Preen, 1992 for the high temperature–light correlation in

the Moreton Bay area).

Longstaff (2003) conducted study on the light requirement of seagrass at several locations in

Moreton Bay, which included the investigation of factor(s) controlling seagrass distribution in the

area. His study concluded that sediment resuspension is the most important physical factor

controlling water clarity in Moreton Bay and long-term status of the water clarity affected seagrass

distribution in the bay. As sediment resuspension is triggered by wave and current actions, the

understanding of fluid dynamics in seagrass environment at various scales (e.g. from molecules to

ecosystems) (Koch et al, 2006) would be important to understand the spatial distribution of seagrass

biophysical parameters, including productivity, in Moreton Bay.

5.3.3. Conclusions, Perspectives and Future Work

This study has demonstrated that the combination of high resolution remote sensing, field

measurement and bio-optical models enabled the estimation of seagrass productivity at

local/regional scale. The approach conducted in this study is different from the empirical method

relating in situ productivity and image reflectance (e.g., Dierrsen et al., 2010) or the scaling-up

method assigning satellite-based community maps with literature or in situ values of community

metabolism (e.g., Andrefouet and Payri., 2000; Brock et al., 2006; Moses et al., 2009). The

empirical method can provide robust and useful results as long as there is a causal relationship

between the parameters of interest and image reflectance. Additionally, the application of the

empirical method to other areas should be conducted with caution, as the empirical method may be

site-specific. The scaling-up method using the combination of remote sensing and in situ

metabolism measurement determined by a large benthic chamber was proven useful to estimate

production and calcification of reef systems at the reef zone scale (Brock et al., 2006). This method,

however, would assign the same value of metabolism rate (i.e., production) to the same community

(i.e., seagrass) inhabiting different environments (i.e., intertidal and sub-tidal). Variations in the

Page 127: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

107

productivity of the same seagrass community living in different environments have been previously

reported (e.g., Moriarty et al., 1990; Rasheed et al., 2008) and need to be considered in the satellite-

based methods for estimating productivity. Our approach addressed this issue by using two models

(process-based and biomass-based methods) that incorporate key parameters affecting seagrass

productivity, and then applying the models by pixel basis, allowing productivity rates to vary

according to pixel-by-pixel changes of the parameters used in the model.

Two models (process-based and biomass-based methods) have been successfully applied to

satellite data to estimate seagrass productivity in the Eastern Banks with results that were still

within the range of literature estimates of seagrass productivity. Overall, average seagrass

productivity in the Eastern Banks calculated using process-based 1 (equation 1), process-based 2

(equation 2), biomass-based 1 (equation 3; June 2012) and biomass-based 2 (equation 3; February

2013) methods were 288.85 O2 g-1 · dry wt · hr-1, 205.49 O2 g-1 · dry wt · hr-1, 1.58 g · dry wt · m-2 ·

day-1 and 1.55 g · dry wt · m-2 · day-1 respectively. Regardless of the method used, the productivity

of the Wanga Wallen banks stood out among other individual banks, supporting previous studies

that observed highest seagrass cover and the most consistent species composition overtime in this

bank (Lyons et al., 2011, 2012; Roelfsema et al., 2014). Analysis using the biomass-based method

applied to WorldView-2 imageries of June 2012 (winter) and February 2013 (summer) revealed no

significant seasonal change in average seagrass productivities, for either the whole Eastern Banks or

the five individual banks. This condition may be explained by the contrasting temporal growth

patterns of the dominant species Zostera muelleri and the other species inhabiting the Eastern

Banks, cancelling out seasonal productivity between summer and winter. Analysis using classified

satellite-based productivity maps revealed spatial variations in seagrass productivity among and

within individual banks, which may be explained by local variations in nutrient or water quality

parameters driven by local hydrodynamics regimes.

In regard to the consideration of implementing between the process-based and biomass-

based method, the choice would depend on the available input parameters to implement either

method. When light and LAI maps are available the process-based method can be applied,

otherwise biomass-based productivity maps can be generated from the available image-based

biomass map.

The first obvious limitation of this study is the absence of representative concurrent or near-

concurrent in situ data to validate the satellite-based seagrass productivity maps and the model

Page 128: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

108

parameters obtained from literature values (α, Pmax, Ik and kc). At remote sensing scale the ideal

validation technique of satellite-based productivity map would require the use of benthic chamber

covering both single and multiple species to measure in situ community metabolism (Nakamura and

Nakamori., 2009; Brock et al., 2006). This has never been done in the study site. However, despite

the limitation, it would not limit the direct applicability of seagrass productivity formula used as it

is a analytical/semi-analytical form and for the thesis it was modified to accommodate the variation

in seagrass physiology parameters (e.g. the photosynthetic efficiency parameter- α was decreased

according to decreasing irradiance light). Alternatively, the cheaper leaf marking or rhizome

tagging methods could be used to validate the satellite-based productivity map process using the

biomass-based method.

The second limitation relates to the parameterisation of the process-based model that has not

incorporated nutrient and temperature parameters, although acknowledging the correlation between

light and temperature. Addressing this issue would require close examination of the form of the

relationship between productivity and temperature and nutrients, and then formalising the

relationships using suitable bio-optical models. From a remote sensing perspective, the use of

nutrient and temperature parameters for habitat biophysical modelling is promising as both

parameters (nutrient data could be in the form of chlorophyll a, particulate organic or inorganic

carbon – POC/PIC) are the products of ocean colour satellites (e.g., MODIS, MERIS).

The third limitation is the lack of local hydrodynamics information for the study site. Waves

and currents can have direct effect on seagrass growth as they can displace seagrass seedlings or

erode sediments causing the smothering of the emerging seedlings (Koch et al., 2006; Marion and

Orth., 2012). In regard to the environmental factors limiting seagrass growth, waves and currents

can also distribute salt, heat and nutrients (Wilhelmus and Dabiri., 2014). Although seagrass in

Moreton Bay and the Eastern Banks is well studied, the possible role of local hydrodynamic factor

in seagrass dynamics is largely overlooked. The inclusion of this factor, from either field

measurements or modelling, would complete the understanding of the spatial and temporal

dynamics of seagrass in Moreton Bay and the Eastern Banks.

Page 129: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

109

Chapter 6: Conclusions, Research Limitations and Future Work

6.1. Summary

Seagrasses are green marine flowering plants that widely inhabit coastal areas and have

important ecological functions in aquatic ecosystems. Light availability is the primary factor that

determines seagrass density, distribution and productivity and, through that, its integrity as a marine

primary producer. A review on seagrass in the 1970s described its primary production as

"comparable to the world's best agricultural crop" (McRoy and McMillan, 1977 in McRoy and

Phillips, 1980, p. ix) which was later confirmed by a more recent study stating that seagrass

productivity stood out compared with terrestrial and other aquatic primary producers (Mateo et al.,

2006). In addition to light availability and productivity, leaf area index (LAI) and biomass are

important parameters that can be used to characterise and monitor the condition of seagrass habitat.

These four parameters are regularly measured in situ, that is, by voluntary-based international

seagrass initiatives monitoring seagrass condition around the world (e.g., www.seagraswatch.org;

www.seagrassnet.org).

Scientists and managers, monitoring and managing seagrass habitat, require information on

seagrass parameters in spatially continuous and multi-temporal representations. Field measurements

are usually costly to produce and time-consuming to analyse over a continuum of scales (Mumby et

al., 2014). Remote sensing is a prospective and cost-effective means for mapping and monitoring

seagrass light climate, productivity, LAI and biomass on a regular basis. However, remote sensing

studies addressing these parameters have been limited and lacking in both appropriate scale of

coverage and suitable bio-optical model parameterisation (Dekker et al., 2013).

The aim of this PhD thesis was to characterise the spatial and temporal distribution of light

climate, LAI, biomass and productivity of an area of sub-tropical seagrass in the Eastern Banks of

Moreton Bay, Australia by integrating in situ measurements, laboratory analyses, bio-optical

models and remote sensing methods. To achieve the research aim, the thesis addressed the

following objectives: (a) to investigate light quality and quantity in seagrass environments and their

influence on seagrass surface area, by using field and remote sensing data, (b) to develop a method

for the estimation of seagrass LAI and biomass using WorldView-2 image data and, (c) to estimate

seagrass gross primary productivity using remote sensing and bio-optical models.

Page 130: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

110

In addition to providing the understanding of the in situ condition of seagrass habitat in the

study site in terms of light climate, LAI and biomass parameters, the result of this thesis provided

insight into how to map these parameters using remote sensing and how to incorporate them into

suitable bio-optical models for producing satellite-based seagrass productivity maps. Improvements

in the comprehension of the biophysical parameters of seagrass achieved in this study will assist in

the understanding and monitoring of seagrass habitats and contribute to continuing advancements in

the methods of mapping and modelling seagrass, which is a habitat of considerable ecological and

economic importance in coastal ecosystems.

6.2. Contributions and Outcomes

In this section, individual research objectives to achieve the thesis aim are restated and their

main outcomes are presented:

Objective 1.

Characterise light quality and quantity in seagrass environments using in situ optical

measurements and remote sensing.

The major outcome of this objective was the development of a method that combined in situ

light quality and satellite-based light quantity information to assess the effect of underwater light

climate on seagasss spatial distribution in the study site. This method resulted in the identification

of green-light dominance of the seagrass site and indications that the seagrass in the Wanga Wallen

site was blue-light limited. Additionally, the seagrass surface area was successfully mapped using

the satellite-based light parameters of PAR, PUR and % SI, and the seagrass maximum colonisation

depth was estimated from the satellite-based water column diffuse attenuation coefficient (Kd). The

comparison between satellite-based PAR and % SI values and the corresponding in situ data was

very close, although the rigorous standard accuracy assessment could not be implemented due to the

limited number of validation points. The outcome was supported by the comparable performance

between the standard semi-analytical satellite algorithm for Kd and the local algorithm developed

empirically from in situ data. This would allow the cost-effective use, for the study site, of satellite-

based light parameters produced using the semi-analytical algorithm, as the algorithm had already

been regularly applied in free-based ocean colour satellite data. In a broader context, the outcome of

research objective one enables scientists and managers to set the light thresholds sustainable for

seagrass using satellite-based light parameters and implement the spatial and temporal monitoring

of seagrass light climates using the threshold for research or management purposes.

Page 131: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

111

Objective 2.

Develop a method to estimate seagrass LAI and biomass using WorldView-2 image data

The main outcome of this objective was a set of methods integrating field-, laboratory- and

remote sensing-based approaches to achieve LAI and biomass mapping using WorldView-2

satellite data. At the in situ level, the nature of the correlation between LAI and biomass and

reflectance data collected using a field spectrometer was examined, resulting in a high correlation

between green band light and LAI and a non-significant correlation between reflectance and

biomass. At the laboratory analysis level, significant allometric relationships between LAI and

biomass were confirmed, allowing the estimation of biomass from LAI. At the remote sensing level,

the depth-corrected green band was used to generate an LAI map, from which the biomass map was

later generated. The accuracy of the satellite-based LAI and biomass maps was found to be at an

acceptable level (LAI = 62% for June 2012 and 73% for February 2013; Biomass = 71% for June

2012 and 60% for February 2013). In addition to augmenting field measurements and facilitating

large-scale monitoring, the satellite-based LAI and biomass maps can be used as inputs

for productivity models, as LAI has been related to growth rate and productivity is partly

biomass dependent.

Objective 3.

Estimate seagrass gross primary productivity using remote sensing data

The major outcome of this objective was the application of productivity models using

remote sensing that allowed the realistic incorporation of seagrass productivity parameters and the

adjustments of these parameters based on physiological and environmental parameters. The chapter

explored the use of light availability, leaf area index and biomass to produce seagrass productivity

maps that could vary on a pixel basis according to the input parameters used. With this method, for

example, the productivity of intertidal seagrass would be different from that of sub-tidal seagrass.

This variation would be difficult to achieve using the scaling-up method that assigns the same

productivity values to generally classified satellite-based communities (e.g., seagrass, sand, and

reef). The models proved to be useful for producing seagrass productivity values that were still

within the range of published literature values, and both the process- and biomass-based models

showed consistently high productivity values from the Wanga Wallen banks. Overall, the outcome

provided methodological insight into satellite-based estimations of seagrass productivity and

facilitated additional progress towards an operational satellite-based method for estimating primary

Page 132: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

112

production similar to that achieved for terrestrial forest and ocean surface systems. Such systems

would provide essential tools for stakeholders in coastal management to monitor the health of

seagrass habitats.

6.3. Limitations and Future Work

The individual chapters of this thesis covered chapter-specific limitations and future

directions. Any limitations that applied to all chapters are discussed in this section and combined

with a few additional key limitations and future directions for the thesis.

The first limitation is the limited availability and spatial coverage of in situ data. Remote

sensing is by no means to replace field survey methods, and therefore sufficient in situ data to

establish remote sensing algorithms and models and to validate the satellite-based maps are

required. In this regard, more in situ optical measurements will give a more complete understanding

of light quality in the Eastern Banks. As previous researchers suggested, treating individual banks in

the area as separate seagrass environments implies possible different light climate conditions among

the banks. This will also facilitate a rigorous accuracy assessment for the satellite-based light

parameters using the collected in situ data.

About this first limitation, better accuracies will also be expected for the resultant satellite-

based LAI and biomass maps if more spatially and evenly distributed in situ data (instead of the

rather sporadic points used in this study) are available for model development and validation. In situ

data were also required for more accurate parameterisation of productivity models and for

validating the resultant satellite-based productivity. According to Green et al. (2000), there are four

main basic components when undertaking a remote sensing project: (a) set-up costs (hardware and

software), (b) field survey costs, (c) image acquisition costs, and (d) the time spent on field data

analysis and image processing. When a remote sensing facility is already available, field survey

costs are estimated to be about 25% of the total cost and its time consumption is about 70% of total

time spent on the project.

The second aspect that deserves further research is the limitation of the image-based

bathymetry data used in this study. Reliable bathymetry data are essential for shallow water remote

sensing applications for benthic habitats as the water column effect interferes with signals detected

by the remote sensors and correction using depth information is required (Lyzenga, 1978, 1981;

Maritorena et al., 1994; Stumpf and Holderied, 2003; Lyzenga et al., 2006). In this thesis, the effect

Page 133: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

113

of water column (depth) on image reflectance was taken into account in the regression analysis

between LAI and the image green band, with the result that the correlation was much stronger with

the depth-corrected green band than with the original green band (Chapter 4).

The majority of anomalous pixel values observed in the resultant satellite-based light (PAR,

PUR, % SI), LAI, biomass, and productivity maps coincided with shallow dense seagrass areas that

were mistakenly classified as deep areas in the WorldView-2 image-based bathymetry. Lyons et al.

(2011) suggested that this might reflect the unsuitability of the Lyzenga method (1978, 1981) to

derive image-based bathymetry over dark seagrass substrate as the original method was developed

over the bright substrate of coral reef environments. Similar to the recommendation of Lyons et al.

(2011), if a better image-based bathymetry method is sought then the alternative physics-based

methods of Brando et al. (2009) or 4SM (the Self-calibrated Supervised Spectral Shallow-sea

Modeler) (http://www.watercolumncorrection.com/) are some of the available options. Another

option is to complete the areal coverage of the shipboard echosounder for the Eastern Banks to

facilitate representative depth extrapolation. Compiling, then extrapolating the available

echosounder data was attempted but the resultant bathymetry map was unreliable as many areas

were not covered by the echosounder’s measurements.

The third limitation is the lack of the use of hydrodynamics data, which together with

bathymetry data, are indicators of important determinants that drive water motions/circulation in an

area. Seagrass growth can be disrupted if seedlings are dislodged by wave motion and currents.

Also, since seagrass in the Eastern Banks is limited to shallow areas (e.g., Hyland et al., 1989,

Roelfsema et al., 2014) it is subject to the impact of sediment re-suspension that can affect light

penetration or smother the newly emerging seedlings. Waves and currents also play important roles

in distributing environmentally limiting factors in seagrass habitats, such as heat (temperature) and

nutrients. As the various satellite-based parameters derived in this thesis are influenced by

hydrodynamic factors, having data available for these factors, either through direct measurement or

modelling, would be useful in explaining seagrass dynamics or environmental factors among or

within individual banks in the Eastern Banks. This would make sure the resultant satellite-

based seagrass parameters were meaningful, either from remote sensing or seagrass

ecology/biology perspectives.

Finally, one aspect that deserves further consideration is the transferability of the methods

used in this thesis in term of the site locality. The Eastern Banks area is characterised by clear and

Page 134: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

114

shallow waters, with relatively continuous seagrass meadows. Further adjustments and

advancements in the satellite band and water-column correction algorithm used may be required to

test the applicability of the methods used here to less favourable environments, such as in turbid and

deep waters with patchy seagrass habitat (see Knudby and Nordlund, 2011 for patchy seagrass

habitat). Additionally, regression analysis between reflectance and the parameters of interest may

need to be re-defined to suit local conditions. Remote sensing applications for benthic habitat

mapping and analysis were mostly conducted in shallow and clear waters (e.g. oligothrophic or

near-oligothropic waters) employing a water column correction technique (review by Green et al.,

1996 and Green et al., 2000). Lyzenga method (Lyzenga, 1978, 1981), probably one of the most

widely used water column correction techniques, requires clear waters and homogeneous

assumption of water optical properties (e.g. Kd). Despite available publications addressing the

theoretical aspects of remote sensing application for benthic habitat in turbid waters (Spitzer and

Dirks., 1987; Tassan, 1996), "at present it is not possible to give a threshold turbidity at which

satellite imagery will be ineffective" (Mumby et al., 1997c). In a similar context for coral reefs,

Hedley et al. (2012) found that the main limiting factor for benthic habitat mapping was the spectral

variation of benthic types and sub-pixel mixing, whereas instrument noise only played a minor role.

When turbidity becomes a limiting factor for seagrass mapping objectives using remote sensing,

underwater video technique (McDonald et al., 2006) or AUV (autonomous underwater vehicle)

(Roelfsema et al., 2015) combined with interpolation technique (e.g. Holmes et al., 2007) can be

used as alternative means. Moving beyond simple mapping objectives (e.g. general classification of

benthic habitat), it may be required to conduct seagrass analysis or derive biophysical parameters at

species level using remote sensing. While seagrass species mapping from remote sensing has

gradually become operational for Eastern Banks area (Phinn et al., 2008; Lyons et al., 2011.,

Roelfsema et al., 2014), the derivation of, for example, seagrass photosynthetic or canopy

parameters cannot be implemented using image data. In this case a three-dimensional radiative

transfer model for shallow water environments (Hedley, 2008) can be used to derive seagrass

parameters at species or canopy levels even when their collection methods are challenging using in

situ method, e.g. within-canopy diffused attenuation coefficient, PAR absorption within canopy,

canopy bidirectional reflectance distribution functions (BRDF) (Hedley and Enriquez., 2010). The

various outputs of the model then can used as inputs for biophysical models involving the use of

remote sensing data. The inputs for the model can be seagrass morphometry, water optical

characteristics and or hydrographic information (tide, current), allowing realistic representation of

seagrass and its complex interacting factors. Still, it is subject to future studies to integrate this

model with remote sensing method for operational use of understanding seagrass habitat.

Page 135: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

115

List of References

Abal, E. G., & Dennison, W. C. (1996). Seagrass depth range and water quality in southern

Moreton Bay, Queensland, Australia. Marine and Freshwater Research, 47, 763-771.

Ahmad, W., & Neil, D. T. (1994). An evaluation of landsat thematic mapper (TM) digital data for

discriminating coral reef zonation: Heron Reef (GBR). International Journal of Remote

Sensing, 15, 2583-2597.

Agawin, N. S. R., Duarte, C. M., Fortes, M. D., Uri, J. S., & Vermaat, J. E. Temporal changes in the

abundance, leaf growth, and photosynthesis of three co-occurring Philippine seagrass. (2001).

Journal of Experimental Marine Biology and Ecology, 260, 217-239.

Al-Najjar, M. A. A. (2010). Flow of light energy in benthic photosynthetic microbial mats. PhD

Thesis. University of Bremen.

Anastasiou, C. J. (2009). Characterization of the underwater light environment and its relevance to

seagrass recovery and sustainability in Tampa Bay, Florida. PhD Thesis. University of South

Florida.

Anderson, M. C., Neale, C. M. U., Li, F., Norman, J. M., Kustas, W. P., Jayanthi, H., & Chavez, J.

Upscaling ground observations of vegetation water content, canopy height, and leaf area

index during SMEX02 using aircraft and Landsat imagery. Remote Sensing of Environment,

92(4), 447–464.

Andrefouet, S., & Payri, C. (2000). Scaling-up carbon and carbonate metabolism of coral reefs

using in situ data and remote sensing. Coral Reefs, 19, 259–269.

Armstrong, R. A. (1993). Remote-sensing of submerged vegetation canopies for biomass

estimation. International Journal of Remote Sensing, 14, 621–627.

Atkinson, M.J., & Grigg, R. W. (1984). Model of coral reef ecosystem II : Gross and net benthic

primary production at French Frigate Shoals, Hawaii. Coral Reefs, 3, 13–22.

Austin, R. W., & Petzold, T. J. (1986). Spectral dependence of the diffuse attenuation coefficient of

light in ocean water. Optical Engineering, 25 (30), 471-479.

Bach, S. S., Borum, J., Fortes, M. D., & Duarte, C. M. (1998). Species composition and plant

performance of mixed seagrass beds along a siltation gradient at Cape Bolinao, The

Philippines. Marine Ecology-Progress Series, 174, 247-256.

Badosa, J., Martial, H., & Chepfer, H. (2013). Scales of spatial and temporal variation of solar

irradiance on Reunion tropical island. Solar Energy, 88, 42-56.

Page 136: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

116

Bargain, A., Robin, M., Le Men, E., Huete, A., & Barille, L. (2012). Spectral response of the

seagrass Zosteranoltii with different sediment backgrounds. Aquatic Botany, 98, 45-56.

Behrenfeld, M. J., & Falkowski, P. G. (1997). Photosynthetic rates derived from satellite-based

chlorophyll concentration. Limnology and oceanography, 42, 1–20,

doi:10.4319/lo.1997.42.1.0001.

Bishop, J. K. B., & Rossow, W. B. (1991). Spatial and temporal variability of global surface solar

irradiance. Journal of Geophysical Research, 96 (C9), 16839-16858.

Borrego-Avecedo, R., Roelfsema, C., Phinn, S. R., & Grinham, A. (2014). Predicting distribution of

microphytobenthos abundance on a reef platform by combining in-situ underwater

spectrometry and pigment analysis. Remote Sensing Letters, 5 (5), 461-470.

Björk M., Short, F. T., Mcleod, E., & Beer, S. (2008). Managing seagrasses for resilience to

climate change. Gland, Switzerland: IUCN.

Boon, P. I. (1986). Nitrogen pools in seagrass beds of Cymodocea serrulata and Zostera capricorni

of Moreton Bay, Australia. Aquatic Botany, 25, 1-19.

Bracchini, L., Loiselle, S. A., Tognazzi, A., Datill, A. M., Forcadi, S., Cozar, A., & Rossi, C.

(2007). The optical qualities of shallow wetland lined bays in Lake Victoria. Wetlands

Ecology and Management, 15, 509 – 519.

Brando, V. E., Anstee, J. M., Wettle, M., Dekker, A. G., Phinn, S. R., & Roelfsema, C. (2009). A

physics based retrieval and quality assessment of bathymetry from suboptimal hyperspectral

data. Remote Sensing of Environment, 113, 755-770.

Bricaud, A., Morel, A., & Prieur, L. (1983). Optical efficiency factors of some phytolankters.

Limnology and Oceanography, 28 (5), 816-832.

Brock, J. C., Yates, K .K., Halley, R. B., Kuffner, I.B., Wright, C. W., & Hatcher, B. G. (2006).

Northern Florida reef tract benthic metabolism scaled by remote sensing. Marine Ecology

Progress Series, 312, 123–139.

Brouns, J. J. W. M. (1987). Aspects of production and biomass of four seagrass species

(Cymodoceoideae) from Papua New Guinea. Aquatic Botany, 27, 333-362.

Burd, A. B. & Dunton, K. H. (2001). Field verification of a light-driven model of biomass changes

in the seagrass Halodule wrightii. Marine Ecology Progress Series, 209, 85-98.

Page 137: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

117

Campbell, J., et al. (2002). Comparison of algorithms for estimating ocean primary production from

surface chlorophyll, temperature, and irradiance. Global Biogeochemical Cycles, 16 (3),

1029-2001. doi:10.1029/2001GB001444.

Campbell, S. J., McKenzie, L. J., Kerville, S. P., & Bite, J. S. (2007). Patterns in tropical seagrass

photosynthesis in relation to light, depth and habitat. Estuarine, Coastal and Shelf Science,

73, 551-562.

Carr, M. E., et al. (2006). A comparison of global estimates of marine primary production from

ocean color. Deep Sea Research, Part II, 53, 741–770, doi:10.1016/j.dsr2.2006.01.028.

Carruthers, T. J. B & Walker, D. I. (1997). Light climate and energy flow in the seagrass canopy of

Amphibolis griffithii (J.M. Black) den Hartog. Oecologia, 109, 335-341.

Chalker, B. E., & Dunlop, W. C. (1983). Bathymetric adaptation of reef-building corals at Davies

Reef, Great Barrier Reef, Australia. II: Light saturation curves for photosynthesis and

respiration. Journal of Experimental Marine Biology and Ecology, 73, 37-56.

Charles-Edwards, D., Doley, D., & Rimmington, G. M. (1986). Modelling Plant Growth and

Development. New York: Academic Press.

Chartrand, K. M., Rasheed, M., Petrou, K., & Ralph, P. (2012). Establishing tropical seagrass light

requirements in a dynamic port environmental. Proceedings of the 12th International Coral

Reef Symposium, Cairns, Australia.

Clavier, J., Chauvaud, L., Carlier, A., Amice, E., Van der Geest, M., Labrosse, P., Diagne, A., &

Hily, C. (2011). Aerial and underwater carbon metabolism of a Zostera noltii seagrass bed in

the Banc d'Arguin, Mauritania. Aquatic Botany, 95, 24-30.

Clavier, J., Chauvaud, L., Amice, E., Lazure, P., Van der Geest, M., Labrosse, P., Diagne, A.,

Carlier, A., & Chauvaud, S. (2014). Benthic metabolism in shallow coastal ecosystems on the

Banc d'Arguin, Mauritania. Marine Ecology Progress Series, 501, 11-23.

Collier, C. J., & Waycott, M. (2009). Drivers of change to seagrass distributions and communities

on the Great Barrier Reef: Literature review and gaps analysis. Reef and Rainforest Research

Centre Limited, Report to the Marine and Tropical Sciences Research Facility.

Cummings, M. E., & Zimmerman, R. C. (2003). Light harvesting and the package effect in the

seagrasses Thalassia testudinum Banks ex Konig and Zostera marine L.: Optical constraints

on photoacclimation. Aquatic Botany, 75, 261-274.

Dawes. C. J. (1998). Marine Botany. New York: John Wiley and Sons.

Page 138: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

118

Dekker, A. G. (1999). Detection of optical water quality parameters for eutrophic waters by high

resolution remote sensing. PhD Thesis. VrijeUniversiteit, The Netherlands.

Dekker, A. G., Brando, V. E., & Anstee, J. M. (2005). Retrospective seagrass change detection in a

shallow coastal tidal Australian lake. Remote Sensing of Environment, 97 (4), 415-433.

Dekker, A. G., Brando, V. E., Anstee, J. M., Fyfe, S., Malthus, T., & Karpouzli, E. (2006). Remote

sensing of seagrass ecosystems : Use of spaceborne and airborne sensors. In A. W. Larkum,.

D., R. J. Orth & C. M. Duarte (Eds.), Seagrasses : Biology, Ecology and Conservation. The

Netherlands: Springer.

Dekker, A. G., Phinn, S. R., Anstee, J., Bisset, P., Brando, V. E., Casey, B., Fearns, P., Hedley, J.,

Klonowski, W., Lee, Z. P., Lynch, M., Lyons, M., Mobley, C., & Roelfsema, C. (2011).

Intercomparison of shallow water bathymetry, hydro-optics, and benthos mapping techniques

in Australia and Carribean coastal environments. Limnology and Oceanography: Methods, 9,

394-423.

Dennison, W. C. (1987). Effects of light on seagrass photosynthesis, growth and depth distribution.

Aquatic Botany, 27, 15-26.

Dennison, W.C., Orth, K. A., Moore, R. J., Stevenson, J. C., Carter, V., Kollar, S., Bergstrom, P.

W., & Batiuk, R. A. (1993). Assessing water quality with submersed aquatic vegetation.

BioScience, 43, 86-94.

Dierssen, H. M., Zimmerman, R. C., Leathers, R. A., Downnes, T. V., & Davis, C. O. (2003).

Ocean color remote sensing of seagrass and bathymetry in the Bahamas Banks by high-

resolution airborne imagery. Limnology and Oceanography, 48 (1, part 2), 444-455.

Dierssen, H. M., Zimmerman, R. C., Leathers, R. A., Downnes, T. V., & Davis, C. O. (2010).

Benthic ecology from space: Optics and net primary production in seagrass and benthic algae

across the Great Bahama Bank. Marine Ecology Progress Series, 48 (1, part 2), 444-455.

Drake, A. L., Dobbs, F. C. & Zimmerman, R. C. (2003). Effects of epiphyte loads on optical

properties and photosynthetic potential of the seagrass Thalassia testudinum Banks ex Konig

and Zostera marina L. Limnology and Oceanography, 48 (1, part 2), 456-463.

Duarte, C. M. (1991). Seagrass depth limits. Aquatic Botany, 40, 363-377.

Duarte, C. M., & Chiscano, C. L. (1999). Seagrass biomass and production : A reassessment.

Aquatic Botany, 65, 159-174.

Page 139: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

119

Duarte, C. M., Borum, J., Short, F. T., & Walker, D. I. (2008). Seagrass ecosystem : Their global

status and prospects. In N. Polunin (Ed.), Aquatic Ecosystems. London: Cambridge University

Press.

Duarte, C. M., Marbà, N., Gacia, E., Fourqurean, J. W., Beggins, J., Barrón, C., & Apostolaki, E. T.

(2010). Seagrass community metabolism: Assessing the carbon sink capacity of seagrass

meadows. Global Biogeochemical Cycles, 24, 1-8.

Duarte C. M., Kennedy, H., Marba, N., Hendriks, I. (2011). Assessing the capacity of seagrass

meadows for carbon burial: Current limitations and future strategies. Ocean & Coastal

Management, 10, 1−11.

Dubinsky, Z., Falkowski, P. G., Porter, J. W., & Muscatine, L. (1984). Absorption and utilization of

radiant energy by light- and shade-adapted colonies of the hermatypic coral Stylophora

pistillata. Proceeding of the Royal Society of London B, 222, 203-214.

Dunton, K. H., & Tomasko, D. A. (1994). In situ photosynthesis in the seagrass Halodule wrightii

in a hypersaline subtropical lagoon. Marine Ecology Progress Series, 107, 281-293.

Durako, M. J. (2007). Leaf optical properties and photosynthetic leaf absorptances in several

Australian Ssagrasses. Aquatic Botany, 87.83-89.

Dustan, P., Dobson, E., & Nelson, G. (2001). Landsat Thematic Mapper: Detection of shifts in

community composition of coral reefs. Conservation Biology, 15(4), 892–902.

Eakin, C. M., Nim, C. J., Brainard, R. E., Aubrecht, C., Elvidge, C., Gledhill, D. K., Muller-Karger,

F., Mumby, P. J., Skirving, W. J., Strong, A. E., Wang, M., Weeks, S., Wentz, F., & Ziskin,

D. (2010). Monitoring coral reefs from space. Oceanography, 23 (4), 118-133.

Enriquez, S., Agusti, S., & Duarte, C. M. (1994). Light absorption by marine macrophytes.

Oecologia, 98, 121-129.

Enriquez, S. (2005). Light absorption efficiency and the package effect in the leaves of the seagrass

Thalassia testudinum. Marine Ecology Progress Series, 289, 141-150.

Enriquez, S., & Pantoja-Reyes, N. I. (2005). Form-function analysis of the effect of canopy

morphology on leaf self-shading in the seagrass Thalassia testudinum. Oecologia,14, 235–

243.

Enriquez, S., Mendez, E. R., & Iglesias-Prieto., R. (2005). Multiple scattering on coral skeletons

enhances light absorption by symbiotic algae. Limnology and Oceanography, 50(4), 1025-

1032.

Page 140: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

120

Erftemeijer, P. L. A., & Lewis III, R. R. R. (2006). Environmental impacts of dredging on

seagrasses: A review. Marine Pollution Bulletin, 52, 1553-1572.

Erftemeijer, P. L. A., Osinga, R., & Mars, A. E. (1993). Primary production of seagrass beds in

South Sulawesi (Indonesia): A comparison of habitats, methods and species. Aquatic Botany,

46, 67-90.

Flanigan, Y. S., & Critchley, C. (1996). Light response of D1 turnover and photosystem II

efficiency in the seagrass Zostera capricorni. Planta, 198, 319–323.

Fourqurean, J. W., Willsie, A., Rose, C. D., & Rutten. L. M. (2001). Spatial and temporal pattern in

seagrass community composition and productivity in South Florida. Marine Biology, 138,

341-354.

Falkowski, P. G., & Raven, J. A. (1997). Aquatic Photosynthesis. Washington, DC: Blackwell

Science, Capital City Press.

Fourqurean, J. W. & Zieman, J. C. (1991). Photosynthesis, respiration and whole plant carbon

budget of the seagrass Thalassia testudinum. Marine Ecology Progress Series, 69, 161-170.

Frouin, R., Lingner, D. W., & Gautier, C. (1989). A simple analytical formula to compute clear sky

total and photosynthetically available solar irradiance at the ocean surface. Journal of

Geophysical Research, 94 (7), 9731–9742.

Fyfe, S. K. (2003). Spatial and temporal variation in spectral reflectance: are seagrass species

spectrally distinct ? Limnology and Oceanography, 48, 464–479.

Gallegos, C. L. (1994). Refining habitat requirements of submersed aquatic vegetation: role of

optical models. Estuaries, 17, 187-199.

Gallegos, C. L., & Kenworthy, W. J. (1996). Seagrass depth limits in the Indian River Lagoon

(Florida, U.S.A.): Application of an optical water quality model. Estuarine, Coastal and Shelf

Science, 42, 267-288.

Gamon, J. A., Penuelas, J., & Field. C. B. (1992). A narrow-waveband spectral index that tracks

diurnal changes in photosynthetic efficiency. Remote Sensing of Environment, 41, 35– 44.

Gamon, J. A., Serrano, L., & Surfus. J. S. (1997). The photochemical reflectance index: an optical

indicator of photosynthetic radiation use efficiency across species, functional types and

nutrient levels. Oecologia, 112, 492–501.

Page 141: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

121

Gattuso, J. P., Gentili, B., Duarte, C. M., Kleypas, J. A., Middleburg, J. J., & Antoine. D. (2006).

Light availability in the coastal ocean: Impact on the distribution of benthic photosynthetic

organism and their contribution to primary production. Biogeosciences, 3, 489-513.

Gray, J., & Song, C. (2012). Mapping leaf area index using spatial, spectral and temporal

information from multiple sensors. Remote Sensing of Environment, 119, 173-183.

Green, E. P., Mumby, P. J, Edwards, A. J., & Clark, C. D. (1996). A review of remote sensing for

the assessment and management of tropical coastal resources. Coastal Management, 24 (1), 1-

40.

Green, E. P., Mumby, P. J., Edwards, A. J., & Clark, C. D., & Ellis, A.C. (1997). Estimating leaf

area index of mangroves from satellite data. Aquatic Botany, 58, 11-19.

Green, E. P., Mumby, P .J., Edwards, A. J., & Clark, C. D. (2000). Remote Sensing Handbook for

Tropical Coastal Management. Paris: UNESCO.

Grice, A. M., Loneragan, N. R., & Dennison, W. C. (1996). Light intensity and the interactions

between physiology,morphology and stable isotope ratios in five species of seagrass. Journal

of Experimental Marine Biology and Ecology, 195, 91-110.

Gonzalez-Liboy, J. (1979). An examination of the present condition of seagrass meadows in La

Parguera, Puerto Rico. Final Report. Department of Natural Resources, Commonwealth,

Puerto Rico.

Hackney, J. W. (2003). Morphometric variability and allometric relationships in the seagrass

Thalassia testudinum in Florida Bay. Master Thesis. University of North Carolina at

Wilmington.

Hansen, J. W., Udy, J. W., Perry, C. J., Dennison, W. C., & Lomstein, B. A. 2000. Effect of the

seagrass Zostera capricorni on sediment microbial process. Marine Ecology Progress Series,

199, 83-96.

Haxo, F. T., & Blinks, L.R. (1950). Photosynthetic action spectra of marine algae. Journal of

General Physiology, 33(4), 389-422.

Hedley, J. (2008). A three-dimensional radiative transfer model for shallow water environments.

Optics Express, 16 (26), 21887-21902.

Hedley, J., & Enriquez, S. (2010). Optical properties of canopies of the tropical seagrass Thalassia

testudinum estimated by a three-dimensional radiative transfer model. Limonology and

Oceanography, 55 (4), 1537-1550.

Page 142: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

122

Hedley, J. D., Roelfsema, C. M., Phinn, S. R., & Mumby, P. J. Environmental and sensor

limitations in optical remote sensing of coral reefs: Implications for monitoring and sensor

design. Remote Sensing, 4, 271-302.

Heiskanen, J. (2006). Estimating aboveground tree biomass and leaf area index in a mountain birch

forest using ASTER satellite data. International Journal of Remote Sensing, 27 (6), 1135-

1158.

Hena, M. K. A., Misri, K., Sidik, B. J., Hishamuddin, O., & Hidir, H. (2001). Photosynthesis of

seagrass Cymodocea serrulata (Magnoliophyta/Potamogetonales/Cymodoceaceae) in field

and laboratory. Indian Journal of Marine Sciences, 30 (4), 253-256.

Hill, V., Zimmerman, R. C., Bissett, P., & Kohler, D. (2011). Exploring the Utility of the

Broadband Multispectral Sensor WorldView 2 for Investigation of Nearshore Coastal

Enviroment. 34th International Symposium on Remote Sensing of Environment, Sydney.

Hill, V. J., Zimmerman, R. C., Bisset, W. P., Dierssen, H., & Kohler, D. D. R. (2014). Evaluating

light availability, seagrass biomass, and productivity using hyperspectral airborne remote

sensing in Saint Joseph's Bay, Florida. Estuaries and Coasts, doi:10.1007/s12237-013-9764-

3.

Hilker, T., Coops, N. C., Wulder, M. A., Black, T. A., & Guy, R. D. (2008). The use of remote

sensing in light use efficiency based models of gross primary production: A review of current

status and future requirements. Science of The Total Environment, 404, 411-423.

Hillman, K., Walker, D.I., Larkum, A.W.D., & McComb, A.J. (1989). Productivity and nutrient

limitation. In A.W.D. Larkum, A. J. McComb, S. A. Shepherd (Eds.), Biology of seagrasses

(pp. 635–685). Amsterdam: Elsevier.

Hillman, K., McComb, A. J., & Walker, D. I. (1995). The distribution, biomass and primary

production the seagrass Halophilaovalis in the Swan/Canning estuary, Western Australia.

Aquatic Botany, 51, 1-54.

Hochberg, E. J., Atkinson, M. J., & Andrefouet, S. (2003). Spectral reflectance of coral reef bottom-

types worldwide and implications for coral reef remote sensing. Remote Sensing of

Environment, 85, 159– 173.

Hochberg, E. J., & Atkinson, M. J. (2008). Coral reef benthic productivity based on optical

absorptance and light-use efficiency. Coral Reefs, 27, 49–59.

Page 143: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

123

Holden, H. (2002). Characterisation of optical water quality in Bunaken National Marine Park,

North Sulawesi, Indonesia. Singapore Journal of Tropical Geography, 23 (1), 23-36.

Holmes, K. W., Van Niel, K. P., Kendrick, G. A., & Radford, B. (2007). Probabilistic large-area

mapping of seagrass species distributions. Aquatic Conservation: Marine and Freshwater

Ecosystems, 17, 385-407.

Hyland, S. J., Courtney A. J., & Butler, C. T. (1989). Distribution of seagrass in the Moreton

Region from Coolangatta to Noosa. Queensland Government, Brisbane: Department of

Primary Industries.

Isaksson, T., & Kowalski, B. (1993). Piece-wise multiplicative scatter correction applied to near-

infrared diffuse transmittance data from meat products. Applied Spectroscopy, 47, 702–709.

IOCCG. 2000. Remote Sensing of Ocean Colour in Coastal, and Other Optically-Complex, Waters.

In S. Sathyendranath (ed.), Reports of the International Ocean-Colour Coordinating Group,

No. 3, IOCCG, Dartmouth, Canada.

Jassby, A. T., & Platt, T. (1976). Mathematical formulation of the relationship between

photosynthesis and light for phytoplankton. Limnology and Oceanography, 21, 540–7.

Jerlov, N. G. (1976). Marine optics. Amsterdam: Elsevier.

Jensen, J. R. (2005). Introductory digital image processing: A remote sensing perspective. Upper

Saddle River, NJ: Prentice Hall.

Johnson, P., & Johnstone, R. (1995). Productivity and nutrient dynamics of tropical sea-grass

communities in Phuttalam Lagoon, Sri Lanka. Ambio, 24, (7/8), 411-417.

Kinsey, D. W. (1985). Metabolism, calcification and carbon production I: Systems level studies.

Proceeding of the 5th International Coral Reef Symposium, 4, 05– 526.

Kirk, J. T. O. (2010). Light and Photosynthesis in Aquatic Ecosystems. Cambridge: Cambridge

University Press.

Knudby, A., & Nordlund, L. Remote sensing of seagrasses in a patchy multi-species environment.

2011. International Journal of Remote Sensing, 32 (8), 2227-2244.

Kocak, F., Uluturham, E., Giel, G. Y., & Onen, S. A. (2011). Impact of environmental conditions

on Posidonia oceanica meadows in the Eastern Mediterranean Sea. Indian Journal of Geo-

Marine Sciences, 40 (6), 770-778.

Page 144: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

124

Koch, E. W., Ackerman, J. D., Verduin, J., & Van Keulen, M. (2006). Fluid dynamics in seagrass

ecology - from molecules to ecosystems. In A. W. D. Larkum, R. J. Orth and C. M. Duarte

(Eds.), Seagrasses : Biology, Ecology and Conservation. Berlin: Springer.

Kuhl, M., & Jorgensen, B. B. (1994). The light field of microbenthic communities: Radiance

distribution and microscale optics of sandy coastal sediments. Limnology and Oceanography,

39 (6), 1368-1398.

Kutser, T., Vahtmae, E., Roelfsema, C. M., & Metsamaa, L. (2007). Photo-library method for

mapping seagrass biomass. Estuarine, Coastal and Shelf Sciences, 75, 559-563.

Kutser, T., Miller I., & Jupp, D. L. B. (2006). Mapping coral reef benthic substrates using

hyperspectral space-borne images and spectral libraries. Estuarine, Coastal and Shelf Science,

70, 449-460.

Larkum, A. W. D., Collet, L. C., & Williams, R.J. (1984). The standing stock, growth and shoot

production of Zostera capricorni Aschers in Botany Bay, New South Wales, Australia.

Aquatic Botany, 19, 307-327.

Lee, Z. P., Carder, K. L., Mobley, C. D., Steward, R. G., & Patch, J. S. (1998). Hyperspectral

remote sensing for shallow waters 1: A semianalytical model. Applied Optics, 37, 6328–6338,

Lee, Z. P., Carder, K. L., Mobley, C. D., Steward, R. G., & Patch, J. S. (1999). Hyperspectral

remote sensing for shallow waters 2: Deriving bottom depths and water properties by

optimization. Applied Optics, 38, 3831–3843.

Lee, Z. P., Carder, K. L., & Arnone, R. A. (2002). Deriving inherent optical properties from water

color: A multi-band quasi analytical algorithm for optically deep waters. Applied Optics, 41,

5755-5772.

Lee, Z. P., Du, K., & Arnone, R. (2005a). A model for the diffuse attenuation coefficient of

downwelling irradiance. Journal of Geophysical Research, 110, C02016.

doi:10.1029/2004JC002275.

Lee, Z. P, Darecki, M., Carder, K. L., Davis, C. O., Stramszki, D., & Rhea. W. J. (2005b). Diffuse

attenuation coefficient of downwelling irradiance: An evaluation of remote sensing methods.

Journal of Geophysical Research, 110, C02017, doi:10.1029/2004JC002573

Lee, Z. P., Weidemann, A., Kindle, J., Arnone, R., Carder, K. L., & Davis, C. (2007). Euphotic

zone depth: Its derivation and implication to ocean-color remote sensing. Journal of

Geophysical Research, 112, C03009, doi:10.1029/2006JC003802.

Page 145: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

125

Lee, K., Park, S. R., & Kim, Y. K. (2007a). Effects of irradiance, temperature and nutrients on

growth dynamics of seagrasses : A review. Journal of Experimental Marine Biology and

Ecology, 350, 144-175.

Lee, Z. P. (2009). KPAR : An optical property associated with ambiguous values. Journal of Lake

Science, 21 (2), 159-164.

Liang, S., Zheng, T., Liu, R., Fang, H., Tsay, S. C., & Running, S. (2006). Estimation of incident

photosynthetically active radiation from moderate resolution imaging spectrometer data.

Journal of Geophysical Research-Atmospheres, 111, D15208. doi: 10.1029/2005JD006730.

Longstaff, B. (2003). Investigations into the light requirements of seagrasses in Northeast

Australia. PhD Thesis. The University of Queensland.

Lu, D., Mausel, P., Brondizio, E., & Moran, E. (2004). Relationships between forest stand

parameters and Landsat Thematic Mapper spectral response in the Brazilian Amazon Basin.

Forest Ecology and Management, 198, 149-167.

Luning, K., & Dring, M.J. (1985). Action spectra and spectral quantum yield of photosynthesis in

marine macroalgae with thin and thick thalli. Marine Biology, 87, 119-129.

Lyons, M. B., Phinn, S. R., Roelfsema, C. M. (2011). Integrating quickbird multi-spectral satellite

and field data: mapping bathymetry, seagrass cover, seagrass species and change in Moreton

Bay, Australia in 2004 and 2007. Remote Sensing , 3, 42–64.

Lyons, M. B., Phinn, S. R., & Roelfsema, C. M. (2012). Long term land cover and seagrass

mapping using landsat and object-based image analysis from 1972 to 2010 in the coastal

environment of South East Queensland, Australia. ISPRS Journal of Photogrammetry and

Remote Sensing, 71, 34-46.

Lyzenga, D. R. (1978). Passive remote sensing techniques for mapping water depth and bottom

features. Applied Optics, 17, 379-383.

Lyzenga, D. R. (1981). Remote sensing of bottom reflectance and water attenuation parameters in

shallow water using aircraft and Landsat data. International Journal of Remote Sensing, 2 (1),

71-82.

Lyzenga, D. R., Malinas, N.P., & Tanis, F.J. Multispectral bathymetry using a simple physically

based algorithm. IEEE Transaction on Geoscience and Remote Sensing, 44, 2251-2259.

Maritorena, S., Morel, A., & Gentili, B. (1994). Diffuse reflectance of oceanic shallow waters:

Influence of water depth and bottom albedo. Limnology and Oceanography, 39, 1689–1703.

Page 146: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

126

Marion, S. R., & Orth, R. J. (2012). Seedling establishment in eelgrass: Seed burial effects on

winter losses of developing seedlings. Marine Ecology Progress Series, 448, 197-207.

Markager, S., & Sand-Jensen, K. (1994). The physiology and ecology of light-growth relationship

in macroalgae, in F. E. Round & D. J. Chapman (Eds), Progress in Phycological Research

(pp. 210–266). Biopress Ltd.

Mateo, M. A., Cebrian, J., Dunton, K., & Mutchler,. T. (2006). Carbon flux in seagrass ecosystems.

In A. W. D. Larkum, R. J. Orth and C. M. Duarte (Eds.), Seagrasses : Biology, Ecology and

Conservation. Berlin: Springer.

Maxwell, W. G. H. (1970). The sedimentary framework of Moreton Bay, Queensland. Australian

Journal of Marine and Freshwater Research, 21, 71-88.

McDonald, J. I., Coupland, G. T., & Kendrick, G. A. (2006). Underwater video as monitoring tool

to detect change in seagrass cover. Journal of Environmental Management, 80, 148-155.

McPherson, M. L., Hill, V. J. Zimmerman, R. C., & Dierssen, H. M. (2011). The optical properties

of Greater Florida Bay: Implications for seagrass abundance. Estuaries and Coasts. doi:

10.1007/s12237-011-9411-9.

McRoy, C. P. (1970). Standing stocks and other features of Eelgrass (Zostera marina) populations

on the coast of Alaska. Journal of Fisheries Research Board of Canada, 27 (10), 811-1821.

Mellors, J. E. (1991). An evaluation of a rapid visual technique for estimating seagrass biomass,

Aquatic Botany, 42, 67-73.

Mellors, J. E., Marsh, H.,& Coles, R. G. (1993). Intra-annnual changes in seagrass standing crop,

Green Island, Northern Queensland. Australian Journal of Marine and Freshwaters

Resources, 44, 33-41.

Milford, S. N., & Church, J. A. (1977). Simplified circulation and mixing models of Moreton Bay,

Queensland. Australian Journal of Marine and Freshwater Research, 28, 23-34.

Mishra, D. R., Narumalani, S., Rundquist, D., & Lawson, M. (2005). Characterizing the vertical

diffusive attenuation coefficient for downwelling irradiance in coastal waters: Implications for

water penetration by high resolution satellite data. Photogrammetry and Remote Sensing, 60,

48-64.

Mishra, D. R., Narumalani, S., Rundquist, D., & Lawson, M. (2006). Benthic habitat mapping in

tropical marine environments using Quickbird multispectral data. Photogrammetric

Engineering and Remote Sensing, 72 (9), 1037-1048.

Page 147: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

127

Mobley, C. D. (1994). Light and Water: Radiative transfer in natural waters. Academic.

Mobley, C.D., & Sundman, L.K. (2001). Hydrolight 4.1 users’ guide. Sequois Scientific, Inc.,

Redmond, Wash.

Mobley, C. D., Stramski, D., Bissett, W. P., & Boss, E. (2004). Is the case 1 - case 2 classification

still useful ? Oceanography, June, pp61-67.

Monteith, J. (1972). Solar radiation and productivity in tropical ecosystems. Journal of Applied

Ecology, 9, 747–766.

Monteith, J. L. 1977.Climate and the Efficiency of Crop Production in Britain. Philosophical

Transaction of The Royal Society, London B, 281, 277–294.

Morel, A., & Smith, R. C. (1974). Relation between total quanta and total energy for aquatic

photosynthesis. Limnology and Oceanography, 19 (4), 591-600.

Morel, A., & Prieur, L. (1977). Analysis of variations in ocean color. Limnology and

Oceanography, 22 (4), 709-722.

Morel, A. (1978). Available, useable and stored radiant energy in relation to marine photosynthesis.

Deep-Sea Research, 25, 637–688.

Morel, A., Huot, Y., Gentili, B., Werdell, P. J., Hooker, S. B., & Franz, B. A. (2007). Examining

the consistency of products derived from various ocean color sensors in open ocean (case 1)

waters in the perspective of a multi-sensor approach. Remote Sensing of Environment, 111,

69–88.

Moriarty, D. J. W., Boon, P. I., Hansen, J. A., Hunt, W. G., Poiner, I. R., Pollard, P. C., Skyring G.

W., & White, D. C. (1985). Microbial mass and productivity in seagrass beds.

Geomicrobiology Journal, 4(1), 21-51.

Moriarty, D. J. W., Roberts, D. G., & Pollard, P. C. (1990). Primary and bacterial productivity of

tropical seagrass communities in the Gulf of Carpentaria, Australia. Marine Ecology Progress

Series, 61, 145-157.

Moses, C. S., Andrefouet, S., Kranenburg, C. J., & Muller-Karger, F. E. (2009). Regional estimates

of reef carbonate dynamics and productivity using Landsat 7 ETM+, and potential impacts

from ocean acidification. Marine Ecology Progress Series, 380, 103-115.

Mueller, J. L. (2000). Seawifs algorithm for the diffuse attenuation coefficient, kd490, using water-

leaving radiances at 490 and 555 nm. SeaWiFS Postlaunch Technical Report Series, 24–27.

Greenbelt, Maryland: NASA Goddard Space Flight Center.

Page 148: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

128

Mumby, P. J., Green, E. P., Edwards, A. J., Clark, C. D. (1997a). Measurement of seagrass standing

crop using satellite and digital airborne remote sensing. Marine Ecology Progress Series, 159,

51-60.

Mumby, P. J., Edwards, A. J., Green, E.P., Anderson, C. W., Ellis, A. C., & Clark, C. D. (1997b). A

visual assessment technique for estimating seagrass standing crop. Aquatic Conservation:

Marine and Freshwater Ecosystems, 7, 239-251

Mumby, P. J., Green, E. P., Edward, A. J. & Clark, C. D. 1997c. Coral reef habitat mapping: how

much detail can remote sensing provide ? Marine Biology, 130, 193-202.

Mumby, P. J., Skirving, W. Strong, A. E., Hardy, J. T., LeDrew, E. F. Hochberg, E. J., Stumpf, R

P., & David, L. T. (2004). Remote sensing of coral reefs and their physical environment.

Marine Pollution Bulletin, 48(34), 219.

Myneni, R. B., Hall, F. G., Sellers, P. J., & Marshak, A. L. (1995). The interpretation of spectral

vegetation index. IEEE Transaction On Geoscience and Remote Sensing, 33 (2), 481-486.

Myneni, R. B., Nemani, R. R., & Running, S. W. (1997). Estimation of global leaf area index and

absorbed par using radiative transfer models. IEEE Transactions on Geoscience and Remote

Sensing, 35, 1380–1393.

Myneni, R. B, Hoffman, S., Knyazikhin, Y., Privette, J. L., Glassy, J., Tian, Y., Wang, Y., Song,

X., Zhang, Y., Smith, G. R., Lotsch, A., Friedl, M., Morisette, J. T., Votava, P., Nemani, R. R,

Running, S. W. (2002). Global products of vegetation leaf area and fraction absorbed PAR

from year one of MODIS data. Remote Sensing of Environment, 83(1–2), 214–231.

Nakamura, T., & Nakamori, T. (2009). Estimation of photosynthesis and calcification rates at a

fringing reef by accounting for diurnal variations and the zonation of coral reef communities

on reef flat and slope: a case study for the Shiraho reef, Ishigaki island, southwest Japan.

Coral Reefs, 28, 229-250.

Nellemann, C., Corcoran, E., Duarte, C. M., Valdés, L., De Young, C., Fonseca, L., & Grimsditch,

G. (Eds). 2009. Blue carbon - the role of healthy oceans in binding carbon: A rapid response

assessment. United Nations Environment Programme, GRID-Arendal, www.grida.no.

Nemani, R., Pierce, L., Running, S., & Band, L. (1993). Forest ecosystem processes at the

watershed scale — Sensitivity to remotely-sensed leaf-area index estimates. International

Journal of Remote Sensing, 14, 2519–2534.

Page 149: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

129

Nielsen S. L., & Sand-Jensen, K. (1990) Allometric scaling of maximal photosynthetic growth rate

to surface/volume ratio. Limnology and Oceanography, 35, 177-181

O’Brien, K. R., Grinham, A., Roelfsema, C. M., Saunders, M. I., & Dennison, W. C. (2011).

Viability criteria for the presence of the seagrass Zostera muelleri in Moreton Bay, based on

benthic light dose. 19th International Congress on Modelling and Simulation, Perth, Australia.

Olofsson, P., & Eklundh., L. (2007). Estimation of absorbed par across Scandinavia from satellite

measurements, part II: Modeling and evaluating the fractional absorption. Remote Sensing of

Environment, 110 (2), 240–251.

Olofsson, P., Lagergren, F., Lindroth, A., Lindstrom, J., Klemedtsson, L., Kutsch, W., & Eklundh,

L. (2008). Towards operational remote sensing of forest carbon balance across Northern

Europe. Biogeosciences, 5, 817-832.

O’Reilly, J., Maritorena, S., Mitchell, B. G., Siegel, D. A., Carder, K. L., Kahru, M., Garver, S. A.,

& McClain, C. R. (1998). Ocean color algorithms for SeaWiFS. Journal of Geophysical

Research, 103, 24937–24953.

Orth, J. R. et al. (2006). A global crisis for seagrass ecosystem. BioScience, 56 (12), 987-996.

Ouisse, V., Migne, A., & Davoult, D. (2011). Community-level carbon flux variability over a tidal

cycle in Zostera marina and Z. noltii beds. Marine Ecology Progress Series, 437, 79-87.

Palandro, D. (2006). Coral reef habitat change and water clarity assessment (1984-2002) for the

Florida Keys National Marine Sanctuary using landsat satellite data. PhD Thesis. University

of South Florida.

Papenbrock, J. (2012). Highlights in seagrasses' phylogeny, physiology, and metabolism: what

makes them special? International Scholarly Research Network Botany, 2012.

doi:10.5402/2012/103892.

Parnik, T., Bil, K., Kolmakov, P., & Titlyanov, E. (1992). Photosynthesis of the seagrass

Thalassodendron ciliatum: Leaf anatomy and carbon metabolism. Photosynthetica, 26, 213–

223.

Parsons, R. T., Maita, Y., & Lalli, C. M. (1984). Determination of chlorophylls and total

carotenoids: spectrophotometric method. In T. R. Parsons, Y. Maita & C. M. Lalli [eds.], A

manual of chemical and biological methods for seawater analysis. Pergamon Press.

Page 150: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

130

Peddle, D. R., Johnson, R. L., Cihlar, J., & Latifovic, R. (2004). Large area forest classication and

biophysical parameter estimation using the 5-scale canopy reflectance model in multiple-

forward-mode. Remote Sensing of Environment, 89, 252–263.

Phillips, R. C., & McRoy, C. P. (Eds.). (1980). Handbook of seagrass biology: An ecosystem

perspective. New York: Garland STPM Press.

Phinn, S., & Dekker, A. G. (Eds.), (2004). An Integrated Approach for Adaptive Management of

Complex Coastal Waters – The Moreton Bay Case Study. Final Report – Moreton Bay

Remote Sensing Tasks (MR2).

Phinn, S., Roelfsema, C., Dekker, A., Brando, V., & Anstee, J. (2008). Mapping seagrass species,

cover and biomass in shallow waters: An assessment of satellite multi-spectral and airborne

hyper-spectral imaging systems in Moreton Bay (Australia). Remote Sensing of Environment,

112, 3413-3425.

Pierson, D. C., Kratzer, S., Strombeck, N., & Hakansson, B. (2008). Relationship between the

attenuation of downwelling irradiance at 490 nm with the attenuation of PAR (400 nm - 700

nm) in the Baltic Sea. Remote Sensing of Environment, 112, 668-680.

Platt, T., & Gallegos, C. L. (1980). Modeling primary production. In P.G. Falkowski (Ed.), Primary

Productivity in the Sea, 339-362. Plenum Press.

Plummer, S. E., Malthus, T. J., & Clark, C. D. (1997). Adaptation of a canopy reflectance model for

sub-aqueous vegetation definition and sensitivity analysis. Proceeding of the 4th International

Conference on Remote Sensing for Marine and Coastal Environments, 149-157. Ann Arbor:

Environmental Research Institute of Michigan.

Plus, M., Deslous-Paoli, J. M., Auby, I., & Dagault, Francoise. (2001). Factors influencing primary

production of seagrass beds (Zostera noltii Hornem.) in the Thou lagoon (French

Mediterranean coast). Journal of Experimental Marine Biology and Ecology, 259, 63-84.

Pollard, P. C., & Moriarty, D. J. W. (1991). Organic carbon decomposition, primary and bacterial

productivity, and sulphate reduction in tropical seagrass beds of the Gulf of Carpentaria,

Australia. Marine Ecology Progress Series, 69, 149-159.

Pollard, P. C. (1999). Measuring photosynthetic characteristics of the seagrass Syringodium

isoetifolium: Implications for in situ productivity estimates. New Zealand Journal of Marine

and Freshwater Research, 33(2), 173-180.

Page 151: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

131

Preen, A. R. (1992). Interaction Between Dugongs and Seagrasses in A Subtropical Environment.

Ph.D Thesis, James Cook University of North Queensland, Townsville.

Preisendorfer, R. W. (1976). Hydrologic Optics. Vol. 1, Introduction. Natl.Tech. Inf. Serv.,

Springfield, Va.

Rasheed, M. A., Dew, K .R., McKenzie, L. J., Coles, R. G., Kerville, S. P., & Campbell, S. J.

(2008). Productivity, carbon assimilation and intra-annual change in tropical reef platform

seagrass communities of the Torres Strait, North-Eastern Australia. Continental Shelf

Research, 28, 2292-2303.

Riihimaki, L. D., & Long, C. N. (2014). Spatial variability surface irradiance measurement at the

Manus ARM site. Journal of Geophysical Research:Atmosphere, 119, 5475-5491.

doi:10.1002/2013JD021187.

Roelfsema, C., Marshall, J., Hochberg, E., Phinn, S., Goldzien, A., & Joyce, K. (2006). Underwater

Spectrometer System 2006 (UWSS04). Biophysical Remote Sensing Group, School of

Geography, Planning and Environmental Management, The University of Queensland.

Roelfsema, C. M., Phinn, S. R., Udy, N., & Maxwell., P. (2009). An integrated field and remote

sensing approach for mapping seagrass cover, Moreton Bay, Australia. Journal of Spatial

Science, 54 (1), 45-62.

Roelfsema, C. M., Lyons, M., Kovacs, E. M., Maxwell, P., Saunders, M. I., Samper-Villarreal, J. &

Phinn, S. R. (2014). Multi-temporal mapping of seagrass cover, species and biomass: a semi-

automated object based image analysis approach. Remote Sensing of Environment, 150, 172-

187.

Roelfsema, C., Lyons, M., Dunbabin, M., Kovacs, E. M., & Phinn, S. (2015). Integrating field

survey data with satellite image data to improve shallow water seagrass maps: the role of

AUV and snorkeller surveys ? Remote Sensing Letters, 6 (2), 135-144.

Sathyendranath, S., Prieur, L., & Morel, A. (1989). A three-component model of ocean colour and

its application to remote sensing of phytoplankton pigments in coastal waters. International

Journal of Remote Sensing, 10, 1373-1394.

Saulquin, B, Hamdi, A., Gohin, F., Populus, J., Mangin, A., & d'Andon. O. F. (2013).Estimation of

the diffuse attenuation coefficient KdPAR using MERIS and application to seabed habitat

mapping. Remote Sensing of Environment, 128, 224-233.

Page 152: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

132

Saba, V. S., et al. (2010). Challenges of modeling depth-integrated marine primary productivity

over multiple decades: A case study at BATS and HOT. Global Biogeochemical Cycles, 24,

GB3020, doi:10.1029/2009GB003655.

Schwarz, A. M. (2004). Contribution of photosynthetic gains during tidal emersion to production of

Zostera capricorni in a North Island, New Zealand estuary. New Zealand Journal of Marine

and Freshwater Research, 38, 809–818.

Short, F. T. (1980). A Simulation Model of the Seagrass Production System. In R. C. Phillips and

M.C. Roy (Eds), handbook of seagrass biology: An ecosystem perspective (pp. 277–295).

New York: Garland.

Short, F. T., & Coles, R. G. (Eds). (2001). Global Seagrass Research Methods. Amsterdam:

Elsevier.

Short, F. T., Koch, E. W., Creed, J. C., Magalhaes, K. M., Fernandez, E., & Gaeckle, J. L. (2006).

SeagrassNet monitoring across the Americas: Case studies of seagrass decline. Marine

Ecology, 27, 277–289.

Short, F. T., Carruthers, T., Dennison,W., & Waycott, M. (2007). Global seagrass distribution and

diversity: A bioregional model. Journal of Experimental Marine Biology and Ecology, 350, 3-

20.

Sims, D. A., Rahman, A. F., Cordova, V. D., El-Masri, B. Z., Baldocchi, D. D., Flanagan, L. B.,

Goldstein, A. H., Hollinger, D. Y., Misson, L., Monso, R. K., Oechel, W. C. ,Schmid, H .P.,

Wofsy, S. C., & Xu, L. K. (2006.) On the use of MODIS EVI to Assess Gross Primary

Productivity of North American ecosystems. Journal of Geophysical Research, 111, G04015,

doi:10.1029/2006JG000162.

Solana-Arellano, E., Echavarria-Heras, H., & Martinez, M. G. (2003). Improved leaf area index

based biomass estimations for Zostera Marina L. Mathematical Medicine and Biology, 20,

367-375.

Spitzer, D., & Dirks, R. W. J. (1987). Bottom influence on the reflectance of the sea. International

Journal of Remote Sensing, 8 (3), 279-308.

Stambler, N., & Z. Dubinsky. (2005). Corals as light collectors: An integrating sphere approach.

Coral Reefs, 24, 1-9.

Stoughton, M. A. (2009). A bio-optical model for SyringodiumFiliforme canopies. Master Thesis.

Old Dominion University.

Page 153: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

133

Stumpf, R. P., Holderied, K., & Sinclair, M. (2003). Determination of water depth with high

resolution satellite imagery over variable bottom types. Limnology and Oceanography, 48,

547-556.

Tassan, S. (1996). Modified Lyzenga's method for macroalgae detection in water with non-uniform

composition. International Journal of Remote Sensing, 17 (8), 1601-1607.

Terrados, J., & Borum, J. (2004). Why are seagrass important ? - Goods and services provided by

seagrass meadows. In J. Borum, C. M. Duarte, D. Krause-Jensen and T. M. Greve (Eds),

European Seagrass: an Introduction to Monitoring and Management. The M&MS Project.

Thorhaug, A., Richardson, A. D., & Berlyn, G. P. (2007). Spectral reflectance of the seagrasses:

Thalassia testudinum, Halodule wrightii, Syringodium filiforne and five marine algae.

International Journal of Remote Sensing, 28 (7), 1487-1501.

Udy, J. W., & Dennison, W. C. (1997). Growth and physiological responses of three seagrass

species to elevated sediment nutrients in Moreton Bay, Australia. Journal of Experimental

Marine Biology and Ecology, 217, 253-277.

Udy, J. W., Dennison, W. C., Long, W. J. L., & McKenzie, L. J. (1999). Responses of seagrass to

nutrients in the Great Barrier Reef, Australia. Marine Ecology Progress Series, 185, 257-271.

Veal, C. J. (2010). Meso and Micro-scale Light Fields Within Coral Environments. PhD Thesis.

The University of Queensland.

Veal, C. J., Michael, K. J., & Nunez, M. (2009). Partitioning of underwater direct and diffuse

ultraviolet irradiance in a shallow water coral reef. Marine and Freshwater Research, 60,

1244-1253.

Wang, M., Son, S., & Harding, L. (2009). Retrieval of diffuse attenuation coefficient in the

Chesapeake Bay and turbid ocean regions for satellite ocean color applications. Journal of

Geophysical Research, c10011, 1-15.

Wang, D., Liang, S., Liu, R., & Tao, Z. 2010. Estimation of daily-integrated PAR from sparse

satellite observations: Comparison of temporal scaling methods. International Journal of

Remote Sensing, 31 (6), 1661-1677.

Waycott, M. et al. 2009. Accelerating loss of seagrass across the globe threatens coastal

ecosystems. Proceedings of the National Academy of Sciences, 106 (30), 12377-12381.

Page 154: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

134

Weeks, S., Werdell, P. J., Schaffelke, B., Canto, M., Lee, Z., Wilding, J. G., & Feldman. G. C.

(2010). Satellite-derived photic depth on the Great Barrier Reef : Spatio-temporal patterns of

water clarity. Remote Sensing, 4, 3781-3795.

West, R. J., & Larkum, A. W. D. (1983). Seagrass primary production - a review. Proceedings of

the Linnean Society of New South Wales, 106, 213-223.

Wicaksono, P., & Hafizt, M. (2013). Mapping seagrass from space : Addressing the complexity of

seagrass lai mapping. European Journal of Remote Sensing, 46, 18-39.

Wilhelmus, M. M., & Dabiri, J. O. (2014). Observations of large-scale fluid transport by laser-

guided plankton aggregations. Physics of Fluids, 26.

Wulder, M. A., LeDrew, E. F., Franklin, S. E., & Lavigne, M. B. (1998). Aerial image texture

information in the estimation of northern decidous and mixed wood forest leaf area index

(LAI). Remote Sensing of Environment, 64, 64-76.

Wyman, K., Dubinsky, Z., Porter, J., & Falkowski, P. (1987). Light absorption and utilization

among hermatypic corals: A study in Jamaica, West Indies. Marine Biology, 96, 283–292.

Young, P. C., & Kirkman, H. (1975). The seagrass communities of Moreton Bay, Queensland.

Aquatic Botany, 1, 191-202.

Young, P. C. 1978. Moreton Bay, Queensland: a nursery area for juvenile penaeid prawns.

Australian Journal for Marine and Freshwater Research, 29, 55-75.

Zainal, A., Dalby, D. H., & Robinson, I. S. (1993). Monitoring marine ecological changes in the

east coast of Bahrain with Landsat TM. Photogrammetric Engineering and Remote Sensing,

59, 415-421.

Zaneveld, J. R. V., Boss, E., & Barnard, A. (2001). Influence of surface waves on measured and

modeled irradiance profiles. Applied Optics, 40 (9), 1442-1449.

Zhang, G., Ganguly, S., Nemani, R. R., White, M. A., Milesi, C., Hashimoto, H., Wang, W.,

Saatchi, S., Yu, Y., & Myneni, R. B. (2014). Estimation of forest aboveground biomass in

California using canopy height and leaf area index estimated from satellite data. Remote

Sensing of Environment, 151, 44-56.

Zhao, D., Xie, D., Zhou, H., Jiang, H., & An, S. (2012). Estimation of leaf area index and plant area

index of a submerged canopy using digital photography. Plos One, 7, 12.

Page 155: Characterisation of the Light Environment and Biophysical Parameters of Seagrass Using Remote

135

Zhao, J., Barnes, B., Melo, N., English, D., Lapointe, B., Karger, F. M., Schaeffer, B., & Hu, C.

(2013). Assessment of satellite-derived diffuse attenuation coefficients and euphotic depths in

South Florida coastal waters. Remote Sensing of Environment, 131, 38-50.

Ziegler, S., & Benner, R. (1998). Ecosystem metabolism in a subtropical, seagrass-dominated

lagoon. Marine Ecology Progress Series, 173, 1-12.

Zimmerman, R. C., Cabello-Pasini, A., & Alberte, R. S. (1994). Modeling daily production of

aquatic macrophytes from irradiance measurements : A comparative analysis. Marine Ecology

Progress Series, 114, 185-196.

Zimmerman, R. C. (2003). A bio-optical model of irradiance distribution and photosynthesis in

seagrass canopies. Limnology and Oceanography, 48, 568–585.

Zimmerman, R. C. (2006). Light and photosynthesis in seagrass meadows. In A. W. D. Larkum, R.

J. Orth and C. M. Duarte (Eds.), Seagrasses : Biology, Ecology and Conservation. Berlin:

Springer.