remote sensing studies for the assessment of geohazards

128
Western Michigan University Western Michigan University ScholarWorks at WMU ScholarWorks at WMU Dissertations Graduate College 8-2008 Remote Sensing Studies for the Assessment of Geohazards: Remote Sensing Studies for the Assessment of Geohazards: Toxic Algal Blooms in the Lower Great Lakes, and the Land Toxic Algal Blooms in the Lower Great Lakes, and the Land Subsidence in the Nile Delta Subsidence in the Nile Delta Richard H. Becker Western Michigan University Follow this and additional works at: https://scholarworks.wmich.edu/dissertations Part of the Geology Commons Recommended Citation Recommended Citation Becker, Richard H., "Remote Sensing Studies for the Assessment of Geohazards: Toxic Algal Blooms in the Lower Great Lakes, and the Land Subsidence in the Nile Delta" (2008). Dissertations. 3382. https://scholarworks.wmich.edu/dissertations/3382 This Dissertation-Open Access is brought to you for free and open access by the Graduate College at ScholarWorks at WMU. It has been accepted for inclusion in Dissertations by an authorized administrator of ScholarWorks at WMU. For more information, please contact [email protected].

Upload: others

Post on 27-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Western Michigan University Western Michigan University

ScholarWorks at WMU ScholarWorks at WMU

Dissertations Graduate College

8-2008

Remote Sensing Studies for the Assessment of Geohazards: Remote Sensing Studies for the Assessment of Geohazards:

Toxic Algal Blooms in the Lower Great Lakes, and the Land Toxic Algal Blooms in the Lower Great Lakes, and the Land

Subsidence in the Nile Delta Subsidence in the Nile Delta

Richard H. Becker Western Michigan University

Follow this and additional works at: https://scholarworks.wmich.edu/dissertations

Part of the Geology Commons

Recommended Citation Recommended Citation Becker, Richard H., "Remote Sensing Studies for the Assessment of Geohazards: Toxic Algal Blooms in the Lower Great Lakes, and the Land Subsidence in the Nile Delta" (2008). Dissertations. 3382. https://scholarworks.wmich.edu/dissertations/3382

This Dissertation-Open Access is brought to you for free and open access by the Graduate College at ScholarWorks at WMU. It has been accepted for inclusion in Dissertations by an authorized administrator of ScholarWorks at WMU. For more information, please contact [email protected].

REMOTE SENSING STUDIES FOR THE ASSESSMENT OF GEOHAZARDS: TOXIC ALGAL BLOOMS IN THE LOWER

GREAT LAKES, AND THE LAND SUBSIDENCE IN THE NILE DELTA

by

Richard H. Becker

A Dissertation Submitted to the

Faculty of The Graduate College in partial fulfillment of the

requirements for the Degree of Doctor of Philosophy

Department of Geosciences Dr. Mohamed Sultan, Advisor

Western Michigan University Kalamazoo, Michigan

August 2008

Copyright by

Richard H. Becker

2008

ACKNOWLEDGMENTS

I would like to thank several people for helping and encouraging me

through the course of my doctoral work. I especially thank my advisor Dr.

Mohamed Sultan for his encouragement in pursuing this degree and for

mentoring me throughout the research process. Our discussions have been

very fruitful, and have helped me become a better researcher.

I would like to thank my dissertation committee, Dr. Joseph Atkinson,

Dr. Carla Koretsky, Dr. Barry Lesht, and Dr. William Sauck for their time

and effort in reviewing my proposal and dissertation.

Additional thanks go to Dr. Gregory Boyer, and Elizabeth Konopko at

SUNY ESF, as well as Dr. Michael Twiss at Clarkson University for access to

the data necessary to validate my satellite models for the Great Lakes.

I would also like to thank all of the researchers who have worked in

the ESRS lab, who provided input on these projects.

Finally, thanks go to my wife Dee, who has been very patient and

supporting as I have been pursuing this degree, and to our children, Anna,

James, and William, who have been great sources of joy and energy when it

was needed.

Richard H. Becker

11

REMOTE SENSING STUDIES FOR THE ASSESSMENT OF

GEOHAZARDS: TOXIC ALGAL BLOOMS IN THE LOWER

GREAT LAKES, AND LAND SUBSIDENCE

IN THE NILE DELTA

Richard H. Becker, Ph.D.

Western Michigan University, 2008

Remote sensing techniques provide valuable tools for assessing a

wide variety of environmental phenomena. They have been used for

monitoring and assessment of various types of geologic and environmental

hazards occurring on land, in the air, or in oceans. I present results from

two studies, the first of which examines the spatial and temporal

distribution of algal blooms in the Great Lakes; the second measures

subsidence in the Nile Delta.

In the first study, methodologies to investigate the extent and

distribution (temporally and spatially) of algal blooms in Lake Erie and

Lake Ontario are studied. Millions of people in the U.S and Canada rely

on the Great Lakes for drinking water, food, work, and recreation. Toxic

algal blooms present a hazard to the substantial number of communities

that draw water from the Great Lakes. Visible and infra red MODIS

satellite data are used to map the extent of algal blooms in these lakes.

Existing algorithms to retrieve chlorophyll concentrations are successfully

tested against in situ measurements from sampling cruises. Algorithms

are developed to identify the potentially toxic cyanobacterial blooms.

The second study examines subsidence in the Nile Delta. The

modern Nile Delta is the major agricultural production area for Egypt and

was formed from sediments supplied by at least 10 distinct distributary

channels throughout the Holocene. With an average elevation around a

meter above sea level and with a predicted rise in sea level of 1.8-5.9

mm/year the subsidence of the northern 30 km of the delta is a topic of

major concern to the Egyptian population and government. Ongoing

subsidence rates in the northeastern Nile Delta were estimated using

persistent scatterer radar interferometry techniques. The highest rates

( ~8 mm/yr; twice average Holocene rates) correlate with the distribution

of the youngest deposition, with older depositional centers subsiding at

slower rates of 2-6 mm/yr. Results are interpreted to indicate that: (1)

modern subsidence in the Delta is heavily influenced by the compaction of

the most recent sediments, and (2) the highly threatened areas are at the

terminus of the Damietta, where the most recent deposition has occurred.

TABLE OF CONTENTS

·ACKNOWLED GMENTS .......................................................................... 11

LIST OF TABLES ····················································································· Vl

LIST OF FIGURES . . . . . . . . .. . . . . . . . . .. . . . . . . . .. . . . . .. . . . . . . . . . . . . . . . . .. . . . .. . . . . . . . . . . . . . . . ... . . . . . . . Vll

CHAPTER

1. INTRODUCTION ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ......... ...... 1

1.1 Remot e Sensing of Geologic Haza rds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Diss ert a tion Ov erview....................................................... 6

2. AQUATIC REMOTE SENSING .............................. ....................... 8

2.1 Ov erview............................................................................ 8

2.2 Light a t the Sens or............................................................ 10

2.3 At mos pheric Mod eling....................................................... 12

2.4 Op tica l Components of Wa t er........................................... 14

3. SPATIAL AND TEMPORAL VARIATIONS OF ALGAL

BLOOMS IN THE LOWER GREAT LAKES ............................... .. 16

3.1 Abs t ra ct .............................................................................. 16

3.2 Int roduction .................................................... ................. .. 18

3.3 Ima ge Acquisition a nd Processing.................................... 20

3.4 Field Measurements .......................................................... 21

3.5 Sea WiF S vs. MOD IS Measurements of Chlorophyll-a..... 24

lll

CHAPTER

Table of Contents - Continued

3.6 Satellite vs. In Situ Measurements of Chlorophyll-a....... 27

3. 7 Identification and Mapping of Algal Blooms .. .............. ... 34

3.8 Locating Phycocyanin Containing Blooms ....................... 36

4. MAPPING CYANOBACTERIAL BLOOMS IN THE GREATLAKES USING MODIS................................................................... 39

4.1 Abstract.............................................................................. 39

4.2 Problem Description .......................................................... 40

4.3 Methodology....................................................................... 45

4.4 Field Measurements .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .. .... .. .. .. .. .. .... .. .. .. 4 7

4.5 Algorithm Development and Processing ofMODIS Data.................................................................................... 49

4.6 Results................................................................................ 52

5. RADAR INTERFEROMETRY........................................................ 62

5.1 Overview............................................................................ 62

5.2 Two-Pass Method............................................................... 67

5.3 Three-Pass Method............................................................ 67

5.4 Four-Pass Method.............................................................. 68

5.5 Multi Temporal Methods................................................... 69

5.6 Summary............................................................................ 69

IV

Table of Contents - Continued

CHAPTER

6. LAND SUBSIDENCE IN THE NILE DELTA: INFERENCES

FROM RADAR INTERFEROMETRY............................................ 72

6.1 Abstract.............................................................................. 72

6.2 Introduction . . . . .. . . . . . . . .. . . . ... . . . . . . .. . . . . . . . ... . . . . .. . . . . . . .. . . . . . .. . . . . . . . . . . . 73

6.3 Geology of the Nile Delta................................................... 76

6.4 Radar Interferometry . . . . . .. . . . . . . .. . . . . . .. ....... .. . . . . . . . .. . . . ... . . . . .. . . . . 79

6.5 Discussion and Results...................................................... 82

7. CONCLUSIONS.............................................................................. 88

REFERENCES .. ................................................. ........................................ 91

APPENDIX

SELECTED INTERMEDIATE INSAR PRODUCTS .......................... 112

V

LIST OF TABLES

4.1 Cruise dates and locations on which fluorometer horizontal

profiles were acquired....................................... . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 7

6.1 ERS-1 and ERS-2 scenes used in permanent scatterers study.

Temporal and perpendicular baselines are shown relative to

the scene acquired on 8/24/1995........................................................ 82

Vl

LIST OF FIGURES

2.1 Case I and case II waters................................................................... 9

2.2 Schematic showing components contributing to radiance measured at the sensor...................................................................... 11

3.1 Density scatter plot comparing chlorophyll-a concentration derived from MO DIS and Sea WiFS data.......................................... 25

3.2 MODIS Band 16 reflectance images showing relative atmospheric

contributions .. . . . . . . . . . . . . . .. . . . . . . . ... . . . . .. . . . .. . . . ... . . . .. . . . ... . . . ... . . . ... . . . . . . . .. . . . . . . .. . . 26

3.3 Satellite-derived chlorophyll-a concentrations plotted against in situ-:r:neasured chlorophyll-a concentrations..................................... 29

3.4 Comparison between satellite-derived chlorophyll-a concentration using case I (OC4v4, blue diamonds) and case II (Carder, pink squares) algorithms and in situ measured chlorophyll-a

concentrations . . . .. .. . . . . .. . . . . . . . . . . .. .. . . . . . .. . ... . . . .. . . . . .. . . ...... ... . . . . . . . . . . . . ......... .. . 30

3.5 SeaWiFS case I (OC4v4, blue diamonds) and case II (Carder, pink squares) derived concentrations vs. in situ results overlain over Sea WiFS validation results (gray squares) from http://seabass.gsfc.nasa.gov/matchup_results.html ........................ 32

3.6 In situ chlorophyll-a concentration (Parsons, blue diamonds, Welchmeyer, pink squares) versus depth ......................................... 33

3. 7 Progression of an algal bloom in the Oswego Harbor using a timeseries generated from Sea WiFS and MOD IS images acquired in August of 2004 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

3.8 Chlorophyll-a concentrations extracted from Sea WiFS images acquired in September of 2004.......................................................... 35

Vll

List of Figures - Continued

3.9 Chlorophyll-a concentrations extracted from a processed SeaWiFS image acquired on June 23, 2003 (left), and June 23, 2004 (right) . 36

3.10 Relative abundance of phycocyanin from in situ fluorescence measurements (July 13-16, 2004) overlain on top of relative

abundance of phycocyanin derived from Sea WiFS imagery (July 13 and 15, 2004) ................................. :...................................... 37

4.1 Landsat Thematic Mapper false color composite showing the location of Lake Erie, Lake Ontario and surrounding metropolitan areas............................................................................. 41

4.2 Absorption curves for green and blue green algae ........................... 43

4.3 Relative absorption curves for CDOM, water, green and blue green algae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.4 Cyanobacterial abundance from FluoroProbe profile (open blue circles) plotted vs. time since cruise start......................................... 54

4.5 Cyanobacterial abundance from FluoroProbe from September

2005 IFYLE cruise vs satellite-derived relative cyanobacterial concentrations from the MODIS scene of September 12, 2005 ..... 55

4.6 Abundances of cyanobacterial blooms in Lake Erie during Sept 2005 derived from MO DIS data ............... ........................................ 56

4. 7 Relative abundances of cyanobacterial blooms in Lake Erie

during August 2005 derived from MODIS data .............................. 57

4.8 Relative abundances of cyanobacterial blooms in Lake Erie (Aug 2004) derived from MODIS data (left) and contoured map generated from cell counts at 15 stations in the western basin of Lake Erie from Dyble (2008) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5.1 Geometry of repeat pass interferometry (after Sandwell and

Price, 1998)......................................................................................... 64

5.2 Sample interferogram generated from a pair of radar images acquired on 9-2-92 and 10-31-92 ....................................................... 66

Vlll

List of Figures - Continued

6.1 False color Landsat TM mosaic showing the Nile Delta, the Damietta and Rosetta branches, and urbanized centers (cities and villages) that appear in shades of purple ....................... 7 4

6.2(a) Mean velocities extracted from 14 ascending ERS-1 and ERS-2

scenes acquired over the period 1992-1999 and expressed as

vertical motion of permanent scatterers (negative velocities indicate downward motion) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

6.2(b) Average Holocene subsidence (Stanley and Warne, 1993a), and

Holocene thicknesses from individual borings (colored squares)

(Stanley et al., 1996) .. . . . . .. . . . . . . . ... . . . . . . .. . . . . . ... . . . . ... . . . . . ... . . . . . . .. . . . . .. . . . . .. . . . 83

IX

CHAPTER 1

INTRODUCTION

1.1 Remote Sensing of Geologic Hazards

Remote sensing techniques provide valuable tools for assessing a wide

variety of environmental phenomena. They have been used for monitoring

and assessment of various types of geologic and environmental hazards.

Examples of such applications on land include the use of interferometric

synthetic aperture radar (INSAR) studies to investigate motion associated

with earthquakes (e.g., Sandwell et al., 2002), subsidence related to mining

(e.g., Carnec and Delacourt, 2000) and to groundwater extraction (e.g.,

Hoffmann et al., 2001; Watson et al., 2002), and surface rise due to aquifer

recharge (Lu and Danskin, 2001). Landslides and slope stability have been

investigated using INSAR and visible techniques and data sets (e.g.,

Berardino et al., 2003; Hilley et al., 2004; Mason and Rosenbaum, 2002; Ostir

et al., 2003). Visible satellite data and satellite derived digital elevation

models have been used for assessment of mudflows from volcanic eruptions

(Ker le et al., 2003). Morphologic analyses of these data sets were used to

assess potential hazards form volcanic eruptions (Aydar et al., 2003).

Remote sensing applications have been quite useful in monitoring

1

geohazards in the atmosphere as well. The most obvious would be weather­

predictions and related hazards (Lynch, 2008) and monitoring spatial and

temporal variations of the ozone layer in the middle stratosphere (Stolarski

et al., 2006). Recent advances in remote extraction of precipitation data are

now enabling early warning applications in flood prediction (Pultz and

Scofield, 2002) and landslide risk assessment (Yang et al., 2006).

Remote sensing datasets and techniques have been used to address

geohazards in oceans and inland lakes. Remote sensing data sets have been

used to monitor the ongoing changes in the ocean circulation, including El

Nino events (Fu and Smith, 1996). They have also been used to identify the

overall abundances of phytoplankton blooms in the oceans and to estimate

primary productivity (Campbell et al., 2002). Additional applications include

examining the role phytoplankton plays in carbon cycling and its potential

impacts on the global climate (Schneider et al., 2008).

Remote sensing techniques have also been used to examine geohazards

at the interface of the land, atmosphere, and ocean systems. For example,

remote sensing data sets were used to examine the impacts of land use and

land cover change on water quality in the adjoining lakes and ocean systems

(Ucuncuoglu et al., 2006). They have also been used to assess the impacts of

tsunami on coastlines and vegetation. Examples include studies

investigating the impacts of the 2004 Indian Ocean tsunami (Belward et al.,

2007; Curran et al., 2007; Yang et al., 2007).

2

International institutions have recognized the importance of the potential

of remote sensing techniques in assessing geohazards. UNESCO developed a

Geohazards theme as a part of their Integrated Global Observation Strategy

(IGOS). One of the key components of this theme is the strengthening of their

existing Geosciences Application in Remote Sensing (GARS) Programme

(IGOS, 2003).

A significant advantage of using remote sensing data techniques is

that they can provide simultaneous measurements of observational

parameters over large areas, and can make repeat measurements on regular

schedules. With the deployment of more satellites and airborne instruments

over the past two decades, the availability of these data sets is becoming far

more widespread. Missions such as the National Aeronautic and Space

Administration's (NASA) Earth Observation System (EOS) provide global

coverage using multiple sensors on 30 different operational or planned

platforms (e.g., visible, infra red optical sensors, passive microwave sensors,

gravity, radar scatterometer) (NASA, 2006). Platforms such as NASA's

ASTER and the commercial SPOT satellite series provide higher resolution

(3-15 m) multispectral images and panchromatic stereo images. The

European Space Agency and the Canadian MDA corporation have several

operational radar satellites (ERS-2, ENVISAT, RADARSAT-2), as well as

archived data from previous missions starting in the 1990s (ERS-1,

RADARSAT-1). These and other platforms provide an extensive library of

ongoing acquisitions as well as archival data which could be used to

investigate a wide range of geohazards on land, in ocean water and in the

atmosphere.

3

I have chosen to examine how remote sensing techniques can be used

to examine two specific geohazards which have the potential to affect millions

of people. The first geohazard examines the spatial and temporal distribution

of algal blooms in the Great Lakes and the second examines subsidence in the

Nile Delta. Both exercises are applications that heavily rely on the use of

remote sensing datasets and techniques. The studies provide ways to better

understand and monitor ongoing processes (algal bloom development and

subsidence in the Delta) and have the potential for providing predictive tools

to be used by community leaders and policy makers to address the

investigated hazards.

In the first study, methodologies to temporally and spatially

investigate the extent and distribution of algal blooms in Lake Erie and Lake

Ontario were evaluated. Millions of people in the U.S. and Canada rely on the

Great Lakes for drinking water, food, work, and recreation. Toxic algal

blooms present a hazard to the substantial number of communities which

draw water from the Great Lakes, or use them for recreation (Dyble et al.,

2008). Visible and near infra red MOD IS and Sea WiFS satellite data were

used to map the distribution of algal blooms in these lakes. Existing

algorithms were used to extract chlorophyll concentrations from MODIS

scenes (Carder et al., 1999; O'Reilly et al., 1998) and results were compared

to in situ measurements collected during sampling cruises conducted during

the periods of acquisition of satellite measurements. The successful

algorithms were defined as that best reproduced the in situ measurements.

Algorithms were developed to identify the potentially toxic cyanobacterial

blooms. Again, cyanobacterial concentrations that were derived using my

4

algorithms were successfully tested against in situ measurements.

The second study examines subsidence in the Nile Delta. The modern

Nile Delta is the major agricultural production area for Egypt. It is formed

from sediments supplied throughout the Holocene by at least 10 distinct

distributary channels (Stanley and Warne, 1993b). With an average elevation

of a meter or so above sea level and a predicted rise in sea level of 1.8-5.9

mm/year (IPCC, 2007) the subsidence of the northern 30 km of the delta is a

topic of major concern to the Egyptian population and government.

Ongoing subsidence rates in the northeastern Nile Delta were

measured using persistent scatterer radar interferometry techniques (Hooper

et al., 2007; Hooper et al., 2004). Fourteen ERS-1 and ERS-2 satellite radar

images covering the period of 1992-1999 were used. It was found that the

highest rates (~8 mm/yr; twice average Holocene rates) correlate with the

distribution of the youngest deposits, whereas the older depositional centers

were found to subside at slower rates of 2-6 mm/yr. Results were interpreted

to indicate that: (1) modern subsidence in the Delta is heavily influenced by

the compaction of the most recent sediments, and (2) the highly threatened

areas are at the terminus of the Damietta branch, where the most recent

deposition has occurred.

These two problems are both good candidates for the use of remote

sensing datasets and techniques. Both cover large areas, hundreds of square

kilometers with observational parameters acquired over long periods (years).

Predictions made from the remote sensing data over the study areas have the

potential for providing guidelines to the policy makers when it comes to

measures to address the geohazards of concern. For example, in the case of

the algal bloom study, the success in monitoring the propagation of blooms

spatially and temporally could potentially affect decisions on the operations

at water intake facilities, recreational usage of inland lakes and even

government regulation regarding development plans. Similarly, results on

the Nile subsidence could potentially be used to orchestrate plans for

protecting the threatened areas in the Delta and in developing educated

plans for urban development of these areas.

1.2 Dissertation Overview

Chapters two through four of the dissertation cover topics related to

aquatic remote sensing, and summarize the results of the investigations that

were conducted in Lake Erie and Lake Ontario. Chapter two of this

dissertation provides a general overview of aquatic remote sensing principles,

techniques, and applications. Chapter three is modified from two articles

which have already been published and chapter four has been submitted for

consideration for publication.

The content of chapter three is modified from two studies. The first is a

study that was published in the proceedings of the 8th International

Conference on Remote Sensing for Marine and Coastal Environments (Becker

et al., 2005), and the second a study that was published in the Great Lakes

Research Reviews (Becker et al., 2006). The study utilizes existing

chlorophyll extraction algorithms to extract chlorophyll concentrations in the

optically complex waters of Lake Erie and Ontario. Comparisons between

algorithms were conducted; the ability to integrate data series from MODIS

and Sea WiFS was examined through comparisons of observations extracted

from contemporaneous MOD IS and Sea WiFS acquisitions. The study

evaluates the effects of varying atmospheric conditions and new

methodologies for extracting phycocyanin concentrations from reflectance

data were developed and applied to Lakes Erie and Ontario to assess the

distribution of blue-green algae blooms.

Chapter four addresses the same problem of extracting blue-green

algae concentrations from satellite imagery, but uses absorption spectra

instead, and has been submitted for consideration to the Journal of Great

Lakes Research.

Chapter five provides a general overview of radar interferometry

techniques. Chapter six covers the application of persistent scatterer radar

interferometry techniques for measuring subsidence in the Nile Delta. The

contents of chapter seven have been submitted to the journal Holocene for

consideration for publication.

Chapter seven provides a summary of the work, and suggestion for

future work in these two areas.

7

CHAPTER2

AQUATIC REMOTE SENSING

2.1 Overview

Monitoring water quality in large water bodies (e.g., lake, sea, and

ocean) spatially and temporally using traditional in situ based methods of

sampling at discreet stations is costly due to the high costs for ship time and

for analysis of collected samples. Moreover, in situ methods do not provide

detailed analyses on regional scales given the large extent and the spatial

and temporal variability of such dynamic systems. Remote sensing

techniques have long held great promise for making synoptic measurements

of the near surface properties for such extensive systems. By virtue of their

capabilities for repeatedly conducting simultaneous measurements with the

same observational parameters, remote sensing sensors could potentially

provide consistent measurements of selected chemical and physical

properties for surface water systems on regional scales. Examples of remotely

sensed properties that could be readily measured include temperatures,

concentrations for chlorophyll, suspended sediments, and total dissolved

solids (Kirk, 1994).

8

For the majority of the published literature on this topic (remote

sensing-based water composition), a distinction is being made between case I

and case II waters (Morel and Prieur, 1977) as shown in figure 2.1.

Case I

Case II

2 Suspended Mate� Figure 2.1 Case I and case II waters. Case I waters are dominated by

phytoplankton. Case II waters have a significant portion of colored dissolved

organic matter (CDOM) and suspended material.

Case I waters, the oceanic waters, are those in which phytoplankton is

the dominant optically active component (Fig. 2.1). For case I waters, the

9

optical properties are dominated by the pigment concentration. Most of the

existing algorithms for extracting water quality parameters were originally

derived for these optically simpler case I waters (Muller-Karger et al., 2005).

For case II or coastal and inland waters, the concentrations of these optically

active components do not co-vary. The optically active components of

phytoplankton, suspended organic and inorganic particulates (8PM), and

colored dissolved organic matter (CDOM or gelbstofl) vary independently.

The rivers discharging in lakes and coastal systems play an important role

delivering these components in the form of sediments and nutrients.

2.2 Light at the Sensor

In using space-borne observation, the property that is being measured

is the amount of light received by the sensor in discrete wavelength regions

(bands). The intensity oflight measured at the satellite is a function of the

wavelength dependant solar irradiance, and interactions with the

atmosphere, the surface of the water, and with the components dissolved and

suspended in the water (Fig. 2.2).

10

SUN

Figure 2.2 Schematic showing components contributing to radiance measured

at the sensor.

The upwelling radiance received at the sensor at the top of the

atmosphere (TOA), L 7 (-l), can be broken down into multiple components

(Hooker and McClain, 2000):

The first two components are atmospheric Rayleigh scattering from air

and scattering from aerosols, including aerosol-Rayleigh interactions. The

terms L ..... c and L9are radiances due to whitecaps and sun glint respectively.

Lw

is the upwelling radiance directly above the water surface and t(J) is the

diffuse transmittance of the atmospheric column.

11

To remove the variations in radiances that are related to varying

illumination angles and viewing geometries, L..,..

is converted to normalized

water-leaving radiance (nLw

) (Gordon and Clark, 1981), which is the water

leaving radiance corrected for the varying illumination angles, viewing

angles and sun-earth distance.

2.3 Atmospheric Modeling

Atmospheric contributions dominate (up to 90%) the signal received at

the satellite in aquatic remote sensing applications (Martin, 2004). All the

data that are included in this dissertation were corrected for atmospheric

contributions using algorithms included in the SeaDAS (Sea WiFS Data

Analysis System) software package (Patt et al., 2003). To enable the

implementation of these atmospheric corrections, SeaDAS uses ancillary

meteorological data (e.g., magnitude and direction of surface winds, water

vapor, and humidity) generated by the National Center for Environmental

Prediction (NCEP), and ozone concentrations from the Earth Probe Total

Ozone Mapping Spectrometer (EPTOMS). Wind speed and direction data are

used as input for algorithms determining whitecap radiance. Other

parameters from the meteorological data such as humidity are used as inputs

to determine the appropriate atmospheric model (Gordon and Wang, 1994).

12

The ozone data are used to calculate diffuse transmittance through the ozone

layer (Gordon, 1997).

The atmospheric contributions to the signal received by the satellite

are affected by Rayleigh scattering, whitecaps, interactions between Rayleigh

scattering and aerosols, and water in the atmosphere. For the MODIS

(Moderate Resolution Imaging Spectrometer) data that are being used in this

work, atmospheric corrections are processed using iterative near infra red

correction techniques (Gordon and Wang, 1994; Stumpf et al., 2003). This

technique first estimates and removes Rayleigh and whitecap atmospheric

contributions, leaving only the remaining atmospheric contributions from

aerosols and water (Gordon and Wang, 1994). Rayleigh and whitecap

components are solved for in SeaDAS iteratively using the near infra red

(NIR) bands (7 48 and 869 nm) and using over 25000 radiative transfer

simulations that are based on multiple MODTRAN (Moderate resolution

atmospheric Transmission) atmosphere models (Patt et al., 2003). Because

water heavily absorbs in the NIR wavelength region, L w is assumed to be

close to zero and L 7 (.-l) will consist only of L R (A.)+ LA (A) in this wavelength

region. Since L R(il) has already been modeled and removed, what is left

consists only of contributions from aerosols and Rayleigh-aerosol interactions.

The aerosol contribution to the NIR bands is iteratively fit to an aerosol

model, and LA

(l) is calculated for all bands (Stumpf et al., 2003).

13

2.4 Optical Components of Water

A large number of previous studies have been carried out to

characterize the optical properties of optically active components in water

bodies in order to determine the relative abundance of these components in

the water (such as Bukata et al., 1981; Clarke et al., 1970; Kahru and

Mitchell, 1998; Morel and Prieur, 1977 among many). These components

include the water, suspended sediments (suspended particulate matter, SPM

and suspended particulate inorganic matter, SPIM), colored dissolved organic

material (CDOM or gelbstofl), particulate organics (POM), and pigmented

materials (i.e., chlorophyll-a, phycocyanin) (Davies-Colley et al., 2003).

Two basic approaches have been used to address this issue. The first

involves extracting concentrations of the examined components from the

apparent optical properties (AOPs) such as the radiance or reflectance at the

sensor (e.g. O'Reilly et al., 1998; 2000; Vincent et al., 2004). Knowing the

reflectances of various components, the atmospherically corrected water­

leaving reflectance can be used to provide correlations with the

concentrations of the various components in the examined waters. Most of

the common chlorophyll algorithms used in case I waters are of this type (e.g.

O'Reilly et al., 1998). The drawback of this type of method in optically

complex (case II) waters is that the individual components concentrations are

14

non-linearly related to the measured apparent optical properties, requiring

significantly more data to successfully model.

The second approach relies on extracting theses concentrations from

the inherent optical properties (IOPs) (e.g. Bukata et al., 1998; Pegau et al.,

1999). IOPs refer to properties characterizing how a light field is modified by

physical processes of absorption and scattering independent of the geometric

properties of light field. Examples of the IOP include the total backscatter

and absorption (Bukata et al., 1998; Carder et al., 1999; Lee et al., 2002).

Several methods have been applied to retrieve inherent optical properties

from the apparent optical properties. Examples of these methods are those of

Maritorena (Maritorena et al., 2002), Carder (Carder et al., 1999) and Lee

(Lee et al., 2002). The application of these methods that allows the extraction

of total backscatter and absorption from the apparent optical properties is

enabled in the SeaDAS domain. Knowing the absorption and backscatter

characteristics of the individual components, total backscatter or absorption

can then be deconvolved to calculate the concentrations of individual

components contributing to the total absorption and backscatter.

15

CHAPTER 3

SPATIAL AND TEMPORAL VARIATIONS OF ALGAL BLOOMS IN THE LOWER GREAT LAKES

3.1 Abstract

As part of the MERHAB (Monitoring and Event Response for Harmful

Algal Blooms) Lower Great Lakes project, we are exploring the utility of

chlorophyll identification algorithms, and their use in mapping the spatial

and temporal distribution of algal blooms in Lake Erie and Lake Ontario. We

have processed Sea WiFS and MOD IS data acquired in the summers of 2003,

2004, and 2005 over the investigated lakes. The remotely-sensed data were

used to derive chlorophyll-a concentrations. The chlorophyll data were

compared with in situ measurements acquired largely from mooring stations,

buoys, and from cruise measurements. Comparisons were made between case

I and case II water satellite-based estimates of chlorophyll-a concentrations

and in situ cruise measurements for the 2003-2005 field seasons (July-August

2003, 2004, May-Sept 2005). We used algorithms which were developed for

case I waters (where chlorophyll content and other dissolved and suspended

materials are assumed to co-vary), and applied these algorithms to the more

optically complex waters of Lake Erie and Lake Ontario (Bukata et al., 1981;

O'Reilly et al., 2000). This approach has been successfully applied to Lake

16

Michigan waters (Lesht et al., 2002). We demonstrate similar applications

and successes in Lakes Erie and Ontario, where comparisons of satellite­

based chlorophyll-a concentrations with in situ chlorophyll-a measurements

indicate that these algorithms provide reasonably good estimates of

chlorophyll-a. We also demonstrate an even better correlation with

chlorophyll-a concentrations derived from the case II water algorithm. Using

a series of Sea WiFS and MOD IS scenes with minimal cloud cover, we

mapped annual re-occurrence (same location and time) of chlorophyll

distribution patterns as well as unique algal bloom events in Lakes Erie and

Ontario in the summers of 2003-2005. Our results indicate: (1) a significant

correlation between chlorophyll concentrations extracted from Sea WiFS and

from MODIS scenes acquired on the same day, especially for days with

minimal atmospheric contributions, (2) a general correspondence between

Sea WiFS-based chlorophyll-a concentrations and in situ chlorophyll

measurements, and (3) annual recurrence (same location and time) of

chlorophyll distribution patterns as well as unique algal bloom events in

Lakes Erie and Ontario.

We used the unique spectral characteristics of the phycocyanin

pigment to discriminate between phycocyanin containing blooms and other

blooms. These results have been validated has using phycocyanin

fluorescence data showing relative phycocyanin abundance.

17

3.2 Introduction

We evaluated the utility of SeaWiFS (Sea-viewing Wide Field-of-view

Sensor) and MODIS (Moderate Resolution Imaging Spectrometer) platforms

for mapping algal blooms as a first step towards the spectral identification of

the toxic blooms in the lower Great Lakes. Comparisons were made between

satellite-based estimates of chlorophyll-a concentrations and in situ

measurements for the 2003-2004 field seasons (July and August 2003, 2004).

We used algorithms which were developed for case I waters where

chlorophyll content and other dissolved and suspended materials are

assumed to co-vary, and applied these algorithms to the more optically

complex waters of Lake Erie and Lake Ontario (Bukata et al., 1981; O'Reilly

et al., 2000). This approach has been successfully applied to Lake Michigan

waters (Lesht et al., 2002). We demonstrate similar applications and

successes in Lakes Erie and Ontario, where comparisons of satellite-based

chlorophyll-a concentrations with in situ chlorophyll-a measurements

indicate that these algorithms provide reasonably good estimates of

chlorophyll-a in these lakes. Current and future work is focusing on

evaluating the case I chlorophyll algorithms over a wider array of

concentrations and time spans of the year to further examine the reliability

and accuracy of these methodologies. A case II model (Carder, 2003) is also

tested, and the results are significantly better than the case I models.

18

Monitoring the distribution of algal blooms in the Great Lakes area

using satellite data is often hindered by the cloud cover. The acquisition of a

continuous series of images with minimal cloud cover over an algal bloom of

interest is needed for proper mapping of the temporal and spatial distribution

of the bloom. The acquisition of a series of Sea WiFS or a series of MO DIS

images of such quality over the study area has proven to be difficult. To begin

addressing this problem, we will explore the possibility of integrating the

information (derived chlorophyll-a) contained in Sea WiFS and MO DIS

images. Chlorophyll-a concentrations extracted from MODIS and SeaWiFS

are compatible but are not identical, given the variations in number of, and

the wavelength region covered by, the bands used in extracting the respective

chlorophyll-a values. Nevertheless, we show that using a series of SeaWiFS

and MODIS scenes that are acquired at approximately the same time and

which show minimal cloud cover, the spatial extent of algal blooms of interest

could be readily measured through time. Finally, we use this methodology to

map the temporal and spatial patterns of algal blooms in Lakes Erie and

Ontario in the summers of 2003 and 2004, and to identify the individual and

annually occurring algal bloom events.

19

3.3 Image Acquisition and Processing

We examined SeaWiFS and MODIS (raw and chlorophyll-a

concentrations) data for the summers of 2003 and 2004 over Lakes Erie and

Ontario. Approximately 120 Sea WiFS scenes (2003, 2004) and 20 MOD IS

scenes (2004) were processed using SeaDAS software and applying the

OC4V4 (SeaWiFS), and the OC3 (MODIS) algorithms (O'Reilly et al., 2000)

for the calculation of water leaving radiance and chlorophyll-a

concentrations. The OC3 and OC4V4 are techniques which extract empirical

relationships between ratios of reflectances from three (OC3) or four (OC4v4)

wavelength regions in the visible portion of the spectra and observed

chlorophyll-a concentrations (O'Reilly et al., 2000).

Chlorophyll-a concentrations were also derived using Carder case II

semi-analytic algorithm in SeaDAS (Carder et al., 1999) and results were

compared to those extracted using the OC3 and OC4V4 techniques. The

default iterative Near IR atmospheric models were utilized. Images were

brought to a common projection using commercial image processing software

(ENVI) to allow comparisons to be made between scenes and between remote

sensing and in situ observations.

20

3.4 Field Measurements

3.4.1 ln Situ Sampling and Measurements

Samples were collected and processed for station cruises on the RV

Guardian and RV Limnos during the summers of 2003, 2004 and 2005 by Dr.

Boyer's research group at the State University at New York School of

Environmental Science and Forestry (SUNY ESF). Fluorescence data were

acquired by Dr. Boyer's group at SUNY ESF during 2004 and 2005 and by

Dr. Michael Twiss at the University of Binghamton in 2005.

Discrete water samples (lL) were collected at 1 m depth and filter onto

4 7-mm Whatman 934-AH glass fiber filters for phycocyanin and chlorophyll

determinations. The concentration of chlorophyll-a was determined by

fluorescence using a Turner Designs TD-700 fluorometer equipped with a

chlorophyll excitation (436 nm) and emission (680 nm) filter set and a

mercury blue light source (Welshmeyer, 1994). The concentration of

chlorophyll-a was also computed using spectrophotogrammetry by analysis of

chlorophyll pigments by the trichromatic technique described by Parsons

(Parsons, 1963).

Phycocyanin concentrations were estimated fluorometrically using a

modification of the method of Abalde et al (1998) and Siegelman and Kycia

21

(1983) as detailed in Konopko (2007). Phycocyanin was extracted from these

filters by freezing the samples at -21 deg C and thawing at 4 deg C three

times in 10 mM phosphate buffer (pH 6.8) under dim light. The extract was

clarified by centrifugation at 22,000 x g for 15 min, and the phycocyanin

concentrations in the supernatant determined by fluorescence using a Turner

Designs 10-AU fluorometer equipped with a 577 nm band pass excitation

filter and a 660 nm cutoff emission filter with a cool while light source.

3.4.2 Fluorometer Measurements

Sampling surveys were conducted in Lake Erie and Lake Ontario in

2004 and 2005. Horizontal fluorometer profiles were obtained throughout the

surveys and water samples were occasionally collected for analysis and

calibration of the fluorometer data.

A 45 kg steel "fish" was towed by the RIV Lake Guardian at a speed of

8-11 knots at a depth of 1 m. Water was pumped through a Teflon™-lined

polyethylene tubing into a 15 L polyvinylchloride "ferry box" that provided an

integrated water sample at a horizontal transect resolution of 0.5 km.

During the Sept 2005 cruise a FluoroProbe (bbe Moldeanke, GmbH),

was submersed inside the ferry box and measured temperature and

phytoplankton concentrations. The FluoroProbe uses five different excitation

22

wavelength ranges (450, 525, 570, 590 and 610 nm) sequentially to illuminate

a sample. It detects fluorescence at 680 nm. Based on the different

fluorescence responses to light at the individual excitation wavelengths, the

concentration of different algal groups is calculated based on a gaussian fit.

The FluoroProbe distinguishes phytoplankton into four groups: 1)

Chlorophyta and Euglenophyta, 2) phycocyanin (PC)-rich Cyanobacteria, 3)

Heterokontophyta and Dinophyta, and 4) Cryptophyta and phycoerythrin

(PE)-rich Cyanobacteria based on fluorescence peaks specific to the

phytoplankton types (Beutler et al., 2003; 2002). All data were time stamped

and archived on a field computer (Panasonic, model CF-29). Time-stamped

geographic coordinates were obtained simultaneously from the ship

navigational system.

During the remaining 5 cruises, water was drawn by gravity from the

ferry box to two Turner Designs 10-AU fluorometers, equipped with flow cells

and either a chlorophyll or phycocyanin filter sets (340 to 500 nm excitation,

>665 nm emission) and phycocyanin (excitation 577 nm, emission 660 nm).

Water flow rate was controlled by the head pressure generated by a 5.5 cm x

1.3 m plastic tube. Air bubbles introduced during rough waters were also

removed by this tube. Readings from the fluorometers were collected at one­

minute intervals and linked to the ship's GPS system. This corresponded to

approximately 300 m intervals when the ship was in motion at its normal

cruising speed of 10 knots.

23

3.5 Sea WiFS vs. MOD IS Measurements of Chlorophyll-a

Sea WiFS and MO DIS images are routinely acquired approximately an

hour apart over the study area (SeaWiFS around 12:00 pm and MODIS Aqua

at 13:00 local time). A comparison between chlorophyll-a concentration

extracted from SeaWiFS and MODIS Aqua was conducted for 20 pairs of

scenes with each pair acquired approximately one hour apart in the summer

of 2004. We found a correlation between chlorophyll-a concentrations derived

from SeaWiFS and MODIS data (Fig. 3.1). The figure displays the range of

observed variations in the degree of correspondence between the Sea WiFS

and MODIS data sets. The highest correlation between SeaWiFS and MODIS

data was observed for the pair of scenes acquired on 8/2/04 (correlation

coefficient of 0.65) and the lowest correlation was for the pair of images

acquired on the previous day (on 8/1/04) (correlation coefficient of 0.48). In

addition, a least squares linear fit between the OC3 and OC4 satellite derived

data for 8/2/04 had a slope of 0.78 (perfect fit slope=l), where the least

squares linear fit for the data from 8/1/04 had a slope of 0.28, much further

from the ideal 1:1 relationship. The correlation coefficient for the 6/23/04 data

was .55, with a slope of 0.33, so again there is a correlation, but one where a

bias is introduced, and the OC4 is underestimated relative to the OC3. The

9/23/04 had a correlation coefficient of .55, and a slope of the linear fit of 0.74,

much closer to an ideal fit. Because such differences occurred in a relatively

24

short time period (1 day), we suspect that these variations largely reflect

temporal variations in atmospheric contributions and are less likely to be

related to variations in water characteristics. The variations cause a general

decorrelation between the images, and also a bias towards higher OC3 values

relative to OC4v4 derived values.

u 0

>­.c a.

0

08/02/04

11.25

7.50

3.75 -'.l�f.:,_. r?t:_-· __

0.00-+-....;.,..����������-+ 0,00 3.75 7.50 11.25 15.0

09/23/04

11.25

7.50

3.75

0.00 3.75 7.50 11.25 15.00

0

.c

u Vl LL

15.00 -f--L......L-'---'-L......L-'---'-L......L----'---'-L......L----'---+ 15.00 -f--L......L-'---'-L......L-'---'-L......L----'---'-L......L----'---+

ro Q)

Vl

06/23/04 08/01/04

11.25 11.25

7.50 . :_ ·�\\t:.- :7.50

3.75 3.75

o.oo-+--�����������--+- o.oo-+--�����������--+-

o.oo 3.75 7.50 11.25 15.00 0.00 3.75 7.50 11.25

MODIS Chlorophyll-a (µg/L)(OC3)

Figure 3.1 Density scatter plot comparing chlorophyll-a concentration

derived from MODIS and SeaWiFS data. Areas in shades of red and blue

represent the largest and the smallest densities of picture elements,

respectively.

25

15.00

08/02/04 09/23/04

06/23/04 08/01/04

Figure 3.2 MODIS Band 16 (846 nm) reflectance images showing relative

atmospheric contributions. Brighter intensities show increased atmospheric

contribution (e.g., increased cloud cover).

Atmospheric contributions will cause reflectance values for clear water

to exceed zero in the wavelength regions: 745-785 nm and 845-885 nm. These

spectral regions are covered by MODIS bands 15 and 16 and SeaWiFS bands

7 and 8 (Land and Haigh, 1996; Sturm and Zibordi, 2002). Reflectance over

clear water ranged from 0.010 to 0.014 on 08/02/04 and varied from 0.010 to

0.030 on 08/01/04. Figure 3.2 shows MODIS band 16 images acquired over

Lake Erie on 06/23/04, 08/01/04, 08/02/04, and 09/23/04. The brightest image,

the one displaying the highest reflectance, and the least correspondence

between the derived SeaWiFS and MODIS chlorophyll-a concentration is the

image acquired on 8/01/04. The scene that was acquired on 8/02/04 is among

the darkest images (lowest reflectance), and shows the highest

26

correspondence between the derived chlorophyll-a concentrations (Fig. 3.1).

Thus application of improved atmospheric models (Ruddick et al., 2000) over

inland waters such as Lakes Erie and Ontario should enhance the observed

correlations.

Our results indicate that chlorophyll-a concentration extracted from

MO DIS and Sea WiFS scene pairs acquired around the same time and under

favorable atmospheric conditions can be integrated together to aid analyses.

3.6 Satellite vs. In Situ Measurements of Chlorophyll-a

In situ chlorophyll-a measurements using multiple measurement

techniques were compared with chlorophyll-a concentrations extracted from

Sea WiFS data (Fig. 3.3); the remote sensing and in situ measurements were

acquired around the same approximate date (within 48 hours). Inspection of

figure 3.3 suggests that chlorophyll concentrations from Sea WiFS correlate

well with in situ measurements (pink: Welchmeyer; blue: Parson) and shows

that the Sea WiFS concentrations are more or less similar to the in situ

measurements in magnitude. In situ chlorophyll-a measurements were also

compared with chlorophyll-a concentrations extracted from case I and case II

algorithms from SeaWiFS (Figs. 3.4 and 3.5); the remote sensing and in situ

measurements were acquired around the same approximate date (within 48

hours). Figure 3.4 shows that the Carder case II algorithm does provide a

27

better fit to the in situ data (R2 of 0.78 for Carder method vs R2 of 0.551 for

the OC4v4 method, and slope of the relationship closer to 1). Figure 3.5

shows the data (spectral and in situ) that were used to calibrate and derive

the case I water algorithms for open oceans; the inner diagonal solid lines

mark± 35% agreement relative to the 1:1 central solid line and the dashed

lines are the 1:5 and 5:1 lines encompassing the data set

(http://seabass.gsfc.nasa.gov/matchup_results.html). For comparison, we

plotted our in situ chlorophyll-a data and the chlorophyll-a data derived from

Sea WiFS for the same locations and using case I and case II algorithms.

Inspection of figure 3.5 suggests that all of our case I samples plot within the

envelope defined by the 1:5 and 5:1 lines and for case II samples, the majority

of them plot within the envelope defined by the ± 35% lines. These results

suggest that the application of case I water algorithms provide reasonable

approximations for the chlorophyll-a content in Lakes Erie and Ontario and

that case II algorithms provide more precise estimates.

28

===

1

·c

C,

-�

25 • El<tracted Chi-a Parsons

• Extracted Chi-a Welcluneyer

20 --1:1

----- Linear (Extracted Chi-a Parsons)

-- - -- Linear (E;.ttacted Chi-aWelchmeyer)

15

• ♦

10

••

,,

y = 0.673x -t 1.534 R2 = 0.780 _.·

■ -✓

_,... /.-</= 0.52x+ 1.182

,'

/ 2 =-

o------�----�-----�-----------

0 5 10 15 20

In situ Chi-a (1t9/L)

Figure 3.3 Satellite-derived chlorophyll-a concentrations plotted

against in situ-measured chlorophyll-a concentrations. Blue diamonds

represent measurements made using Parsons technique (Parsons,

1963), whereas those measured using Welschmeyer technique

(Welschmeyer, 1994) are represented by pink squares. Linear fits for

both relationships are displayed, showing a better correlation with

measurements made using the Welschmeyer technique.

29

Differences between the Sea WiFS-based and in situ chlorophyll-a

concentrations extracted using case II algorithms could be attributed to: (1)

20 ,-------------------------------------.

18

16

14

-�12

<1> 10 .::::

:t:: 8

6

4

2

0

♦ Swa-V\Ms - OC4v4

■ Sea-\'Vifs Ca1der

----- Lin war (Sea-VVifs - OC4v4)

- - - Linear (Sea-\'\/ifs Ca1der)

--1:1 iperf9ctMatch)

0 2 4 6

'I= 0.814, +0.323.,.. R' = 0 901 ;­

. ;, 6'

6' ,

, ..... _.,.,, .., ,;: 0.599� + 1.807

, - R"= 0.648

✓ ,,, ,,

.,.. ,,

. .,, ..

8 10 12 In situ Chi-a (llg!L)

14 15 18

Figure 3.4 Comparison between satellite-derived chlorophyll-a

concentration using case I (OC4v4, blue diamonds) and case II (Carder,

pink squares) algorithms and in situ measured chlorophyll-a

concentrations. Linear fits for both relationships are displayed showing

a better correlation with the Carder-derived chlorophyll-a

concentrations.

30

20

differences in "sample size" (1 km.2 for satellite vs. point samples for ground

measurements), (2) variations in chlorophyll concentrations with depth that

are reflected in the depth-integrated satellite measurements, but not in the

in situ measurements, and (3) differences in the time of acquisition. Satellite

measurements are not always acquired at the same time during which the

ground sampling campaign is being conducted.

In figure 3.6, the chlorophyll-a concentrations obtained for seven water

samples from Lake Erie, acquired on July 13, 2004 (location: Lat: 41.816 N;

Lon: 82.983 E), at depths ranging from Oto 7 m, and analyzed using the

Parsons (blue symbol) and Welchmeyer (pink symbol) protocols are compared

to concentrations extracted from Sea WiFS data (yellow line). The figure

shows the large variations of in situ measurements with depth and

demonstrates how the Sea WiFS chlorophyll-a concentration could represent a

depth integrated measurement of the in situ chlorophyll-a concentrations.

31

l lzlc1r,-11 hyll

101 ◊ Sea-Wifs - OC4v4

□ Sea-Wifs Carder •

• • • 1:1

• •

-

10-� ------------------Lo-:: 10

° i0

1 10-

ln 1 u

Figure 3.5 SeaWiFS case I (OC4v4, blue diamonds) and case II (Carder, pink

squares) derived concentrations vs. in situ results overlain over Sea WiFS

validation results (gray squares) from

http://seabass.gsfc.nasa.gov/matchup results.html. Data from both case I and

case II algorithms fall within validation data set ranges.

32

-

-

-

■ •

■ •

- •

■ •

I I

• Extracted Chl-a Parsons(ug/L)

• Extracted Chl-a (ug/L)

Sea-Wifs

■ •

Figure 3.6 In situ chlorophyll-a concentration (Parsons, blue

diamonds, Welchmeyer, pink squares) versus depth. Figure also

shows the Sea WiFS-derived chlorophyll-a (yellow line).

33

3. 7 Identification and Mapping of Algal Blooms

Examination of chlorophyll-a concentration images (extracted from

Sea WiFS and MO DIS) acquired in summers of 2003 and 2004 over Lakes

Erie and Ontario revealed several algal blooms. Blooms were noted as early

as early June in Lake Ontario, and through mid-October in Lake Erie. As

described earlier, we integrated inferences from Sea WiFS and MO DIS data

to provide the most complete cloud free coverage. For example, mapping of a

bloom which was noted in the Oswego Harbor at the beginning of August

2004 was enabled using a time series of chlorophyll-a images extracted from

a combined set of Sea WiFS and MO DIS data sets. The assembled time series

provided a temporal coverage with minimal cloud cover (Fig. 3.7).

>6

µg/L

0.0

µg/L

Figure 3. 7 Progression of an algal bloom in the Oswego Harbor using a time

series generated from Sea WiFS and MOD IS images acquired in August of 2004.

34

Generally the blooms in Lake Erie originate in the western basin and

propagate into the central portion of the Lake (Fig. 3.8). Although these

blooms usually cover much of the central and western basins of Lake Erie,

they do not propagate into the eastern basin (Fig. 3.8). An algal bloom in the

western basin of Lake Erie coincided with a reported cyanobacterial bloom in

the same area (Rinta-Kanto et al., 2005). The only bloom that was observed

in the eastern basin of Lake Erie throughout the investigated period is a

bloom that occurred in the same approximate location in years 2003 and

2004, around the end of June (Fig. 3.9).

Figure 3.8 Chlorophyll-a concentrations extracted from Sea WiFS images acquired

in September of 2004. These are indicative of the propagation of an algal bloom in

the Western and Central Basins of Lake Erie; the inferred bloom (from satellite

data) coincided with a reported cyanobacterial bloom.

35

Figure 3.9 Chlorophyll-a concentrations extracted from a

processed SeaWiFS image acquired on June 23, 2003 (left), and

June 23, 2004 (right). Image shows blooms in the eastern parts

of Lake Erie. This late June bloom is the only bloom identified in

the Eastern Basin of the lake.

3.8 Locating Phycocyanin Containing Blooms

>6

µg/L

0.0

µg/L

One of the primary goals of the MERHAB project is to be able to

discriminate between potentially harmful and non-harmful algal blooms in

the lower Great Lakes. Microcystis species and other toxin-producing

cyanobacteria have been observed in Lake Erie (Brittain et al., 2000; Rinta­

Kanto et al., 2005). As Microcystis is a cyanobacteria containing the

phycocyanin pigment it should be possible to rule out the possible toxicity of

many blooms by the absence of phycocyanin. We developed an algorithm to

allow the identification of the location of potentially toxic cyanobacterial

blooms from Sea WiFS satellite data, based on spectral characteristics of

phycocyanin pigment.

To accomplish this goal, Sea WiFS images were processed to remote

sensing reflectance using SeaDAS 4.8. A mixing model was adopted based on

the following end member components: chlorophyll-a, phycocyanin and water

36

(Vincent et al., 2004), chlorophyll and water (extracted from Sea WiFS images

in areas of >8 µg/L chlorophyll-a and low phycocyanin as determined from

fluorescence data), and water with chlorophyll-a levels below 1 µg/L.

Figure 3.10 Relative abundance of phycocyanin from in situ fluorescence

measurements (July 13-16, 2004) overlain on top of relative abundance of

phycocyanin derived from Sea WiFS imagery (July 13 and 15, 2004).

This phycocyanin abundance extracted from Sea WiFS was compared

with relative abundances of phycocyanin based on fluorescence data acquired

during July 2004 (Fig. 3.10). Our success in spectrally distinguishing

phycocyanin is consistent with the earlier findings in the western Lake Erie

using remote sensing techniques (Vincent et al., 2004). Despite the broad

wavelength regions covered by the TM Landsat bands, and the absence of TM

band(s) in the wavelength region affected by the phycocyanin (PC) (refer to

37

section 4.2 for further details), Vincent et al. (2004) successfully developed

algorithms to detect PC from Landsat TM data for mapping cyanobacterial

blooms in Lake Erie using ratios of reflectance data. We are currently

refining our mixing model to include better end members for high turbidity

and high CDOM areas in an attempt to remove some false positive

identification and increase the areas where the model resolves.

38

CHAPTER4

MAPPING CYANOBACTERIAL BLOOMS IN THE GREAT LAKES USING MODIS

4.1 Abstract

Microcystis and other toxin-producing cyanobacteria have been

documented in Lake Erie and Ontario in the last several years. We developed

algorithms to discriminate potentially toxic cyanobacterial blooms from other

harmless blooms and to extract relative phycocyanin abundances from

Moderate Resolution Imaging Spectrometer (MODIS) satellite data. Lee's

Quasi-Analytical Algorithm is used to calculate total absorption and

backscatter from the 250 m, 500 m and 1 km bands of MOD IS scenes. A

nonnegative least square algorithm was then utilized to unmix relative

concentrations of green algae, blue-green algae, and CDOM and sediments

combined in lake waters using published absorption spectra for these

components. MODIS-derived cyanobacterial concentrations and/or bloom

distributions from 10 scenes acquired in the summers of 2004 and 2005 were

successfully verified against contemporaneous calibrated measurements of

pigments that were acquired from along track fluorometer measurements

from six cruises, and three additional cyanobacterial blooms reported in the

scientific literature between 2002 and 2006. The developed methodologies

39

could potentially be used to develop cost-effective practical screening methods

for rapid detection of, and warning against, cyanobacterial blooms in the

lower Great Lakes.

4.2 Problem Description

Lake Erie (surface area: >25,000 km2; average depth: 20 m) and Lake

Ontario (surface area: 19,000 km2; average depth: 86 m) are two of the

largest 20 lakes in the world (Fig. 4.1). The Niagara River carries the outflow

from Lake Erie into Lake Ontario and accounts for over 85% of the total

inflow to Lake Ontario (Shear et al., 1995). The outflow of Lake Ontario is

carried by the St. Lawrence River which flows into the Atlantic Ocean.

Millions of citizens of the U.S and Canada rely on Lakes Erie and Ontario

and the remaining Laurentian lakes for drinking and recreation purposes.

Increasing reports over the past years of occurrences of harmful algae in the

Great Lakes demonstrated the vulnerability of these resources and the need

for developing methodologies for expedited detection of the spatial

distribution and propagation of these blooms on regional scales (Anderson et

al., 2002; Hallegraeff, 1993).

40

44°

N

42°

N

84°

W

Detroit �

· Windsor-1,

Toledo

80°

W 76°

W

Lake Erie

Erie·

Cleveland

Figure 4.1 Landsat Thematic Mapper false color composite showing the location of Lake Erie, Lake Ontario and surrounding metropolitan areas.

Conditions favoring the formation of the Harmful Algal Blooms (HABs)

are not completely understood. It is believed that a number of factors (e.g.,

sunlight, nutrients, grazing pressure, water temperature, water depth, water

flow, filter feeders) could collectively promote their development (Downing et

al., 2001; Havens, 2008; Paerl, 2008). Cyanobacterial blooms occur in warm

stable water masses where the species can maximize light capture and

nutrient uptake through the formation of gas vacuoles to regulate buoyancy

(Reynolds et al., 1987). Filter feeders (e.g., invasive zebra mussels; Dreissena

polymorpha) discriminate against toxic cyanobacteria in feeding, giving them

a competitive advantage (Vanderploeg et al., 2001).

In the Lower Great lakes, the toxin producing cyanobacterial species

which have been observed include Microcystis, Anabaena and Planktothrix

41

species (Boyer, 2006; Rinta-Kanto et al., 2005). Cyanobacterial blooms have

been reported in the western basin of Lake Erie in late summer to fall of 2000

through 2007. The most significant levels of microcystin toxicity (14.3 to 20.0

µg L-1) were associated with high chlorophyll-a (4 to 40 µg L-1) blooms and

was detected at the mouth of Maumee River in the summers of 2003 and

2004 (Rinta-Kanto et al., 2005) at stations in the western basin and at the

mouth of the Sandusky River. In Lake Ontario a toxic bloom of Microcystis

approached the World Health Organization (WHO) guideline value of 1 µg

microcystin-LR L-1 in 2003 (Boyer et al., 2004; Makarewicz et al., 2006).

Given the potential health risks associated with the blooms and the large

areal extent of these occurrences throughout Lakes Erie and Ontario,

regional and cost effective monitoring solutions should be developed to

provide adequate warning for the impacted populations. In this manuscript

we describe one such approach that takes advantage of the spectral

properties of these blooms as observed in the readily available, frequently

acquired remote sensing Moderate Resolution Imaging Spectrometer

(MODIS) data sets.

The toxic cyanobacteria in the Great Lakes not only possess the

chlorophyll-a pigments like all other phytoplankton, but also have the

accessory pigment phycocyanin (Fay, 1983) that strongly absorbs in the 620

to 630 µm wavelength region (Bryant, 1981; Sathyendranath et al., 1987),

giving the blue-green algae in Lakes Erie and Ontario their characteristic

42

C 0

0 (/)

.D.

Ctl

Q) >

� a> 0::

-Microcy

-Gree1

450 500 55:J 6JO

Wavelength (nm)

650 700

Landsat TM

MERIS

MODIS 500m

MODIS 1km

MODIS 250m

Figure 4.2 Absorption curves for green and blue green algae. Also shown are positions of MOD IS, MERIS and Landsat TM bands.

color. Because all of the toxin-producing species in the Great Lakes are

cyanobacterial, the accessory pigments associated with these blooms can

potentially be used to identify toxic blooms. Earlier attempts to use satellite

data to remotely map the distribution of toxic algae in the Great Lakes and

elsewhere taking advantage of the spectral signature introduced by the

phycocyanin pigment (PC) are limited. One of these methods utilized band

ratios of reflectance values extracted from Landsat TM images (Vincent et

al., 2004). A second method, a semi-analytical algorithm, was applied to

Medium Resolution Imaging Spectrometer (MERIS) reflectance images

43

(Simis et al., 2005). Vincent's approach takes advantage of the high spatial

resolution provided by the Landsat TM imagery, but is disadvantaged by the

infrequent visits (at best every 7 days) of the Landsat system over the study

area and by the absence of bands of sufficient spectral resolution over the

critical wavelength regions to distinguish the phycocyanin absorption feature

from the chlorophyll-a absorption feature. Specifically, there are no Landsat

bands that are positioned over the maximum phycocyanin absorptions in the

620-630 nm wavelength region (Fig. 4.2). The figure shows that this is not the

case with the MERIS data; the phycocyanin absorption features are covered

by MERIS band 6. However, MERIS's temporal availability (three day

repeat, but only acquired when ordered) and the relatively high commercial

cost of the data makes it a less practical system for routine screening and

monitoring purposes.

Monitoring systems that are based on airborne or space-borne (e.g.,

Hyperion) hyperspectral reflectance algorithms (Dekker et al., 1991; Gons et

al., 1992; Jupp et al., 1994) are costly and provide limited spatial coverage

(USGS, 2008). Previous attempts using the 1 km MODIS bands were

discounted as viable methods for discriminating cyanobacterial blooms

(Kutser et al., 2006a), as these bands do not adequately sample the

phycocyanin absorption feature.

In this study, we report the first MODIS-based algorithms which use

absorption spectra to identify potentially toxic cyanobacteria by

44

distinguishing the phycocyanin absorption feature from the chlorophyll

absorption features. Unlike previous attempts, that solely used the 1 km or

250 m bands (Kutser et al., 2006a; Kutser et al., 2006b), our approach

integrates the spectrally wider 250 m MODIS band 1 with the 1 km and 500

m bands enabling the discrimination between the green and cyanobacterial

blooms (Fig. 4.2). A MODIS-based algorithm is advantageous for the

following reasons: (1) the data are free and rapidly available, making the

developed methodologies more appealing to users; (2) reliable atmospheric

models are in place enabling comparisons to be made between MODIS­

derived water compositions extracted from scenes acquired under varying

atmospheric conditions; (3) the revisit time is short (daily coverage from

Aqua) allowing monitoring of spatial and temporal propagation of blooms

with confidence; (4) archival MODIS data are available since July of 2002

facilitating the identification of recurrent seasonal development and

propagation of blooms; and (5) the MODIS bands are placed over critical

wavelength regions enabling the discrimination between the phycocyanin and

chlorophyll absorption (Fig. 4.2).

4.3 Methodology

In this study we developed an algorithm that is based on a non­

negative linear least squares algorithm to unmix the prevailing component

45

concentrations in the waters of Lakes Erie and Ontario both spatially and

temporally. The investigated components include: Colored Dissolved Organic

Matter (CDOM), suspended sediments, green algae, and blue-green algae.

The adopted unmixing procedure was based on the individual components of

the examined waters and on inherent optical property (IOP) of the system.

The !OP-based algorithm provides linear relationships between components

concentrations and the observed properties (Bukata et al., 1998). The

developed algorithm was applied to selected MODIS data sets that were

acquired over the study areas in July and August, 2004, and July, August,

and September, 2005. The selected scenes were acquired around the time

periods during which our field campaigns were conducted (within one week of

the cruise) to enable comparisons to be made between relative cyanobacterial

and other phytoplankton concentrations extracted from space-borne and field

observations. The fluorescence data derived from along track horizontal

profiles acquired during cruises were calibrated against individual on-station

measurements. We further tested our algorithms by comparing space-borne

observations from additional MODIS scenes (late August 2005) to field

observations acquired in the study area (e.g. Dyble et al., 2008) for the

western basin of Lake Erie.

46

4.4 Field Measurements

Sampling surveys were conducted in Lake Erie and Lake Ontario in

2004-2005 (Table 4.1). Four offshore Lake Erie surveys and two Lake Ontario

surveys are reported here and compared with observations made from

MODIS data. Horizontal fluorometer profiles were obtained throughout the

surveys and water samples were occasionally collected for analysis and

calibration of the fluorometer data.

Cruise

Name

MELEE-VIII

Taste &Odor

MELEE-IX

MELEE-X

Taste &Odor

IFYLE

Lake

Erie

Ontario

Erie

Erie

Ontario

Erie

Departure Port and Date

Colborne, ON 7/12/2004

Burlington, ON

8/30/2004

Colborne, ON 7/11/2005

Amherstburg, ON 8/22/2005

Burlington, ON

8/29/2005

Cleveland, OH

9/7/2005

Return Port

and Date

Colborne, ON 7/16/2004

Burlington, ON

9/3/2004

Colborne, ON 7/15/2005

Burlington, ON 8/26/2005

Burlington, ON

9/2/2005

Cleveland, OH

9/11/2005

Instruments

used to measure PC

10-AU

10-AU

10-AU

10-AU &Hydrolab

10-AU &Hydrolab

Fluoroprobe

Table 4.1 Cruise dates and locations on which fluorometer horizontal profiles were acquired.

A 45 kg steel "fish" was towed by the RIV Lake Guardian at a speed of

8-11 knots at a depth of 1 m. Water was pumped through a Teflon™-lined

polyethylene tubing into a 15 L polyvinylchloride "ferry box" that provided an

47

integrated water sample at a horizontal transect resolution of 0.5 km. During

the September 2005 cruise a FluoroProbe (bbe Moldeanke, GmbH), was

submersed inside the ferry box and measured temperature and

phytoplankton concentrations. The FluoroProbe distinguishes phytoplankton

into four groups: 1) Chlorophyta and Euglenophyta, 2) phycocyanin (PC)-rich

Cyanobacteria, 3) Heterokontophyta and Dinophyta, and 4) Cryptophyta and

phycoerythrin (PE)-rich Cyanobacteria (Beutler et al., 2003; 2002). All data

were time stamped and archived on a field computer (Panasonic, model CF-

29). Time-stamped geographic coordinates were obtained simultaneously

from the ship navigational system. During the remaining 5 cruises, water

was drawn by gravity from the ferry box to a fluorometer (Turner Designs,

model 10-AU) which measured chlorophyll and phycocyanin abundance

(Konopko, 2007).

Discrete water samples (1 L) were collected at 1 m depth and filtered

onto 47-mm Whatman 934-AH glass fiber filters for phycocyanin and

chlorophyll determinations and calibration of the fluorometer data.

Chlorophyll-a was determined fluorometrically using the method of

Welschmeyer (1994). Phycocyanin concentrations were estimated

fluorometrically using a modification of the method of Abalde et al. (1998)

and Siegelman and Kycia (1978). Phycocyanin was extracted from these

filters and the phycocyanin concentrations determined by fluorescence as

described in Konopko (2007).

48

4.5 Algorithm Development and Processing of MODIS Data

MODIS images were selected to provide the best spatial coverage of

the lakes and with minimum cloud coverage. Preference was given to scenes

where the lakes were viewed closer to nadir (in the center of the MO DIS

swath) to maximize spatial resolution and minimize atmospheric

contributions. MODIS data scenes were processed in SeaDAS 5.2 software

package. Atmospheric corrections were conducted using the Iterative Near

Infra Red Atmosphere model (Patt et al., 2003). Atmospherically corrected

scenes were then processed using the developed algorithms to extract

compositional information (e.g., cyanobacterial and other phytoplankton

abundances).

The quasi analytical methods of Lee et al. (2002) (Lee QAA method)

was selected to solve for the total absorption (aT) and total back scatter at

each location from the reflectance data (Gordon et al., 1975) (equation 4.1):

(4.1)

Where R(0-,A) is the below surface reflectance, bb is the total backscatter, aT is

the total absorption, and f is a constant that depends on the nature of the

incident light field and volume scattering function.

The Lee QAA method was selected because it makes no a priori

assumptions of phytoplankton or CDOM absorption either in shape or

magnitude, has limited empirical relationships, has been successfully tested

49

with non case I datasets, and works with multi-spectral data. When the

water leaving radiance in any bands fell below a threshold value of 0.01

mW• cm-2 • µ.m- 1 • sr-1, the absorption calculations from the QAA algorithm

proved unreliable, producing artificially high very noisy absorption values.

These locations were masked in the results as unusable.

The absorption component is directly related to the concentration of

the individual colored components, through the following relationship

(Bukata et al., 1998):

ar(A,) = I at (,,l)ci + aw

(4.2)

i=l

Where aT is the total absorption, ai and Ci are the absorption and

concentration of each component, and aw is the absorption of pure water.

Knowing the absorption spectra of these components, this system of linear

equations was then solved for component concentrations using a non-negative

least squares technique (Lawson and Hanson, 1974).

The individual absorption spectra for the components in Lake Erie and

Ontario were extracted from published data: (1) green algae from Kirk (1994),

(2) blue-green Microcystis from Kuster (2006b), (3) water from Pope and Fry

(1997), and (4) CDOM and sediment are represented by an average spectra

generated from the spectra of these two endmembers (reported in Binding et

al., 2008) because the spectra of CDOM and sediment are quite similar.

50

Binding (2008) found that acnoM and asediment in the western basin of Lake

Erie are best represented by fitting an exponential function:

a= e<-s(,l-440)) (4.3)

where Sis best represented by a value of 0.011 nm-1 for suspended sediments,

and by 0.0161 nm-1 for CDOM.

The endmember spectra for CDOM/sediment, blue-green algae, green

algae, and water were resampled to MODIS bands using the MODIS relative

. spectral response functions (Fig. 4.3). The figure shows that the blue-green

algae possess absorption spectra that can be distinguished from those of the

C 0

:;::; Cl.

0 (/) .0 ro

(I) >

� Q) IY

450 500 550 600

Wavelength (nm)

-.-water

-CDOM+Sediment

..,.... Blue-Green

-Green

650 700

Figure 4.3 Relative absorption curves for CDOM, water, green and blue green

algae. Absorption curves resampled to MODIS bands

51

green algae using the MODIS bands, largely due to absorption differences

observed in MODIS bands centered at 645 and 555 nm.

Knowing the total absorption for each image pixel and the individual

absorption spectra for the four selected endmembers, we then solved for the

concentrations of these endmembers using all MO DIS bands centered

between 468 and 678 nm and applying a non-negative linear least squares

technique implemented in MATLAB. This method was adopted to avoid

physically meaningless negative solutions (i.e., negative concentrations).

Solutions for each pixel were evaluated by calculating the root mean squared

error (RMSE) between the modeled total absorption values to the satellite­

derived total absorption values. Those with an RMSE value greater than

0.08m-1 (approximately 20% of the aT at 645 nm in the western basin) were

considered invalid and rejected. The solution for each endmember is a

relative concentration of that endmember.

4.6 Results

Results (relative cyanobacterial concentrations) were re-projected to

UTM zone 17, and compared to along-track fluorometer results from six

cruises (four in Lake Erie and two in Lake Ontario) conducted in 2004-2005

(Table 4.1). We first demonstrate the correspondence between the patterns of

blue-green algae abundances extracted from the September 7-11 2005 cruises

52

and the relative MODIS-derived cyanobacterial concentrations. We then

extract a linear relationship between the FlouroProbe derived cyanobacterial

abundances and the MODIS derived cyanobacterial abundances and use it to

represent the relative MODIS cyanobacterial concentrations as

concentrations. The calibrated algae abundance data from the FluroProbe

used for this relationship is only available in the September 7-11 2005 cruise.

The horizontal profile extracted from the FluoroProbe fluorimeter

measurements from the September 2005 cruise was compared with the

satellite-based relative cyanobacteria concentrations extracted from the

temporally closest satellite image with good coverage of the lake (Fig. 4.4).

The scene was acquired on September 12, 2005, one day following the end of

the five day cruise. The figure displays data as a function of time (first 48

hours of the cruise) and covers the western basin and the first section of the

central basin. The selected sections of the cruise encompass the full range of

measured variations in FluoroProbe phycocyanin concentrations.

53

1.40 - l 1 /

15 1.20

� 2 13 �

1.00

fl

::i

11 u...

5 O.EO

E

....,

� o.w

■ Cv<1nubdckria Aburdance from MODIS

O Cv<1nubacleriil Aburdance Frum FluoruP·ut;e

•.)

"'

� ;:;

7 C

E

� 0.40

0 "'

-�

,.,,

,:i

C

x

F �

3 C

::, 0.70 1

-1

0 1:J 15 20 25 30 35 40 45 Time of aquisition (hours f-om start of cruise)

Figure 4.4 Cyanobacterial abundance from FluoroProbe profile (open blue circles)

plotted vs. time since cruise start. Corresponding MODIS derived cyanobacterial

concentrations extracted for each the FluoroProbe location from the September

12, 2005 image are shown with red squares.

Inspection of figure 4.4 shows the agreement between the MODIS­

derived relative cyanobacteria concentrations and the fluorometer

concentrations for the September 2005 bloom events. This correlation is also

observed on a plot of cyanobacterial abundance from FluoroProbe vs. MODIS

derived relative cyanobacterial concentrations (Fig 4.5). A correlation coefficient of

R2: 0.658 was observed despite the differences in acquisition time between

the two sets (up to 5 days). Variations are to be expected between the field

and satellite-based algal distributions given that blooms will have moved

54

1.0

0.9

0.8

CJ) 0.7

/{= 0.032x 1- 0.084

+---------------.1�-----/--�-�·�'•8�---•

0.6

0.5

0.4

0.3

0.2 ,!;;

.E 0.1

0.0

0 5 10 15 20 25 30

Cyanobacterial Abundance from Flouroprobe (as JL9/L of chlorophyll)

Figure 4.5 Cyanobacterial abundance from FluoroProbe from September 2005

IFYLE cruise vs satellite-derived relative cyanobacterial concentrations from the

MODIS scene of September 12, 2005.

between the time the satellite and the FluoroProbe measurements were made.

The linear relationship was used to represent the relative MODIS

cyanobacterial concentrations in units of µg/L. The results are displayed in

figure 4.6.

In this figure the MODIS-derived cyanobacterial concentrations for the entire

lake are expressed in the same units as the fluorometer derived value, in

units of µg/L as chlorophyll-a. Inspection of figure 4.6 shows the agreement

between the MODIS derived relative cyanobacteria concentrations and the

fluorometer concentrations for the September 2005 bloom events. Figure 4.4

demonstrates this correlation for a limited number of cruise line sections in

55

the western and central Basins, whereas Figure 4.6 shows that the

correlation applies to the remaining sections for the September 2005 cruise

as well.

82°

W 80°

W

0 12.525 50 75 100

Kilometers

42°

N

41°

N

Figure 4.6 Abundances of cyanobacterial blooms in Lake Erie during Sept

2005 derived from MODIS data. Overlain on satellite images are along track fluorescence data from temporally overlapping cruises.

MODIS derived relative cyanobacterial abundances were also

compared with phycocyanin fluorescence values from the remaining cruises

in the summer of 2004 and 2005 (Table 4.1). Because the fluorometer data

for these cruises were uncalibrated and reported in raw flouresence, we chose

to present our unmixing results as relative cyanobacterial abundances.

Another approach that we could have adopted is to apply the linear

relationship extracted from the September cruise data and report results in

units of µg/L as chlorophyll-a. However, such an approach could be inaccurate

given the reported variations in phycocyanin abundance (relative to algae

56

concentration) over the course of a season (Simis et al., 2007). Figure 4. 7

shows along track relative phycocyanin fluorescence data overlain on relative

cyanobacterial abundance derived from MODIS data. Again, the agreement

in bloom distribution between the two data sets is evident. Although not

shown, this is also true for the July 2004 sampling event, where the patterns

seen in the fluorometer data are evident in the satellite data. In one case, the

July 2005 cruise, the MODIS-derived concentrations showed a possible low

level bloom throughout the entire western basin of Lake Erie, whereas the

fluorescence data showed more variations in concentrations: higher towards

the mouth of the Maumee River, lower towards the north-central portion of

the western basin. Simis (2007) showed that the absorption spectra of

cyanobacteria can change over the course of a season as the amount of

82°

W 80°

W

0 12.525 50 75 100

KIiometers

Figure 4. 7 Relative abundances of cyanobacterial blooms in Lake Erie during August 2005 derived from MODIS data. Overlain on satellite images are along track fluorescence data from temporally overlapping cruises.

57

42°

N

41°

N

phycocyanin changes. This may provide one explanation for why the

technique worked well in August and September of 2004 and 2005, but not as

well in July of 2004.

Applications of our algorithms did not produce false positives when

compared with fluorometer data. No cyanobacterial phycocyanin-bearing

blooms were noted from the MODIS data acquired during two cruises over

Lake Ontario, consistent with the fluorometer measurements. The latter

showed negligible phycocyanin concentrations across the entire lake with one

exception, the bay of Quinte, a long and narrow bay (length: >100 km; width:

1-3 km on average) that cannot be resolved from the coarse MODIS data.

The developed algorithms were also successful in identifying

phycocyanin blooms in the western basin of Lake Erie that were reported by

other researchers. The MODIS-derived relative phycocyanin patterns

matched the field-based contoured cell count patterns reported by Dyble

(2008) (Fig. 4.8). Although not shown, the MODIS-derived phycocyanin

concentration patterns also matched those reported by Vincent (2004) for the

September 2002 bloom in the western basin, the blooms in August of 2003

(Rinta-Kanto et al., 2005), and others in 2006 (Bridgeman, personal

communication).

58

83°

W

42°

N

�4 _..:\_ Kilometers f

Figure 4.8 Relative abundances of cyanobacterial blooms in Lake Erie (August

2004) derived from MODIS data (left) and contoured map generated from cell

counts at 15 stations in the western basin of Lake Erie from Dyble (2008).

Despite the reported success identifying toxic algal blooms, one has to

be aware of the potential uncertainties that could be introduced given the

complexities of the investigated systems, the encountered technical

difficulties, and the adopted assumptions and approximations. These factors

could introduce uncertainties in the bloom patterns and concentrations as

attempts are made to extend the developed methodologies spatially (beyond

Lake Erie and Ontario) and temporally (2002 through 2006). Next, we

provide examples of such complexities, uncertainties, and approximations.

There are difficulties in correlating and calibrating MODIS data to

field data. Simultaneous acquisition of in situ and satellite data is not always

possible. Fluorometer measurements are acquired over a period of 4-6 days,

whereas the satellite data represent a single point in time. The sample size

for MO DIS (pixel size at nadir: 250 m to 1 km) exceeds that of the samples

acquired in the field. The sample depth differs as well, MODIS integrates

measurements across a water column of varying depth depending on the

59

concentration of the different absorbers and scatterers and the wavelength of

the light (Bukata et al., 1998) whereas field samples are generally acquired

at a particular depth (in this case 1.5 m below the surface for fluorometer

data). These difficulties are compounded by the fact that wind pattern can

significantly affect the movement of cyanobacterial blooms laterally

(Kanoshina et al., 2003) and the ability of cyanobacteria to regulate their

vertical position in the water column (Reynolds et al., 1987).

Compositionally distinct endmembers can display similar absorption

spectra as is the case with CDOM and sediments. In such cases, it is best to

represent the spectrally similar endmembers as a single endmember, making

it difficult to estimate the proportions of these two components. One or more

of the endmembers could experience natural variations, such as the

variations observed amongst individual cyanobacterial species making the

representation of such similar, yet spectrally variable population, by a single

spectrum, an oversimplification. There are conceivable potential spatial and

temporal variations in endmember compositions and associated spectral

signatures. One such potential variation was described above, the seasonal

variation in phycocyanin composition. Similarly, the CDOM and sediment

spectra can vary with the value of S varying from .007 to .018 nm-1 (refer to

equation 3), variations which can lead to significant spectral differences in

the shorter (412 nm and 443 nm) wavelengths. Locally, one or more

60

endmembers can go undetected resulting in fictitious solutions given the

unsatisfactory selection of endmembers for a particular location.

Unmixing of spectral components works best when extensive field data

are available for correlating and calibrating the sensor-derived component

concentrations, when all of the endmembers are well characterized spatially

and temporally and are spectrally orthogonal (distinct). The developed

methodologies could potentially be used to develop cost-effective practical

screening method for rapid detection of, and warning against, cyanobacterial

blooms in the lower Great Lakes.

61

CHAPTERS

RADAR INTERFEROMETRY

5.1 Overview

Radar interferometry is a technique that uses multiple radar images to

infer topography, and to detect subtle temporal changes in topographic relief.

It has been demonstrated that changes on the order of several cm/yr and as

small as 0.1 mm/yr (under optimum conditions) could be readily measured

using this technique (Massonnet and Feigl, 1998). This technique has been

used to map deformation and fault slip from earthquakes (Sandwell et al.,

2002), mine subsidence (Carnec and Delacourt, 2000), aquifer compaction

from pumping (Burbey, 2003), and landslides (Amelung and Day, 2002), as

well as seasonal changes in surface topography due to groundwater

extraction and recharge (Hoffmann et al., 2001). The ideal setting for

applying these techniques is in arid areas, where interferences from

vegetation and atmospheric effects are minimal. Over the past decade,

advances in radar interferometry applications and techniques have led to

successful applications in temperate regions and over a wider range of

climatic conditions and landscapes.

62

The Synthetic Aperture Radar (SAR) Interferometry technique exploits

the information contained in the phase of two or more acquired over the same

location. The technique makes use of the difference in phase (interferometric

phase) between two radar scenes to determine the exact differences in range

from the satellite, and subsequently determines the precise x, y, and z

location of the reflector, enabling the extraction of topography or subtle

changes in topography.

Figure 5.1 shows the basic configuration of a pair of images used in

repeat pass interferometry. p is the range to a target from the satellite

reference position, and p +b p is the range to the same target acquired in the

second pass. B is the baseline, or physical distance between the location of

the satellite in the first and second pass. 0 is the look angle, and a is the

angle between the baseline vector and the tangent plane. It is then possible

to define bp as a function of B, 0, a, p, and A, the wavelength of the radar

beam. bp is proportional to the phase difference component of the radar

return <j>, measured during the two radar acquisitions:

<j>=(4n/A)*bp (5.1)

The common terminology for the reference scene is the master scene, and the

repeat scene is the slave scene.

63

ref

erence

spheroid

Figure 5.1 Geometry of repeat pass interferometry (after Sandwell and Price, 1998).

The factors which contribute to· phase differences between two radar

scenes include topography, deformation, and atmospheric effects.

<j>= <l>':t'opo + $def + <l>atm+ <!>noise, (5.2)

The phase <j> is recorded cyclically from -n<<j><n, so there is by default

an ambiguity in determining p from <j>.

64

This process is described extensively by Gabriel and Massonnet

elsewhere (Gabriel et al., 1989; Massonnet and Feigl, 1998) .

There are three basic families of radar interferometry techniques

currently in use and under development. These are the basic 2-4 pass

differential INSAR (DINSAR) techniques, as well as two classes of multi­

temporal techniques which use numbers of scenes ranging from tens to

hundreds. The multi-temporal techniques are expansions and refinements of

the basic 2-4 pass techniques. They repeat many of the same steps, and then

extract usable information from results which are ambiguous in the 2-4 pass

techniques.

The basis for all of these techniques is the generation of an

interferogram. To generate an interferogram, the two scenes need to be co­

registered in radar-space. This means that the slave scene (or a subset

thereof) has to be co-registered to the master scene (or a subset of the master

scene). This processing is done in the Delft object-oriented radar

interferometric software (DORIS), a public domain radar interferometry

software application (Kampes et al., 2003). With DORIS, the orbits are used

to provide an initial estimate of the registration, and then the images are

iteratively correlated using the cross correlation amplitude of the radar

signal in individual subsets of the radar images. The radar images can be

filtered (optional) to improve the registration. The slave image is then re-

65

sampled to the master image. The interferogram is then calculated from the

co-registered images as the dot-product of the complex images (Fig. 6.2). This

step is repeated for every interferogram that is generated. Any interferogram

generated like this will have a phase component related to the curvature of

the earth's surface. The curvature is then calculated and removed before any

further processing is done.

"'#1 Phase (Band 1 :9-6·92·10·11 ·92·1nterfero c,nt) temp7

. �

Figure 5.2 Sample interferogram generated from a pair of radar images acquired on

9-2-92 and 10-31-92.

The second necessary component of these techniques is phase

unwrapping. In the above interferogram (Fig. 5.2), a repeating ripple pattern

is evident; the ripples trend from the lower left to the upper right. This is

caused by the phase ranging from -n<q><n, cyclically. In order for this to be

turned into a measure of range, the cycles have to be added together, so that

66

the phase numbers then extend from Oto ~20n, instead of the original cyclical

distribution (-n«j><n). This is called unwrapping, and is a major challenge in

interferometry. The method we use (snaphu) is described in full in Chane and

Zebker (2002). If correlation between scenes is low or if coherence in the

interferogram is low due to changes in the scene through time or increased

perpendicular baseline, then this step becomes almost impossible.

5.2 Two-Pass Method

In the two-pass INSAR technique an interferogram is generated from

two scenes which span a deformation event (the master is acquired before,

the slave after). <l>atm and <!>noise are assumed to be negligible. <)>Topo is calculated

from a digital elevation model (DEM) which has been registered with the

master scene and subtracted from <j> to yield <!>def. This method is fairly simple,

but it relies heavily on the availability of a high quality DEM and excellent

registration between the DEM and the master. Any error in the DEM or in

registration will cause ambiguities in detecting and mapping deformation.

5.3 Three-Pass Method

In three-pass interferometry, instead of a DEM being used, an

unwrapped interferogram is used to remove the <)>Topo component. Noise and

67

atmospheric contributions are again considered to be negligible. An

interferogram from a co-registered master and slave with a very small

temporal and spatial baseline (i.e., 1 day) is generated. An additional slave

image on the other side of the deformation (also with a small spatial baseline)

is co-registered to the same master, and an interferogram is generated. The

interferogram from the first pair is unwrapped, scaled to match the second

baseline, and re-wrapped. The second interferogram is subtracted from the

first, removing the topographic phase, leaving <!>def.

5.4 Four-Pass Method

In the four-pass interferometry, a master-slave pair with very small

temporal and spatial baselines is acquired before and after the deformation

event. Each slave is co-registered to the appropriate master. An

interferogram is generated for each pair. One interferogram is unwrapped

and re-sampled to match the radar coordinates of the other pair. It is then

scaled and re-wrapped. The second interferogram is subtracted from the first,

yielding <!>def. This method has several distinct advantages over the three-pass

method. The <!>noise due to de-correlation is significantly reduced. All four

scenes do not need to share the same small baseline range, but pairs can be

selected to minimize spatial and temporal baselines. This significantly

increases the detection of the resultant deformation. 68

5.5 Multi Temporal Methods

The multi-temporal methods generate a far higher number of

interferogram pairs, throughout a deformation event. By making a large

number of interferograms, and with educated assumptions about the nature

of the deformation (i.e. linear deformation) and of the atmospheric

contributions, the errors associated with the solution can be minimized. In

the Small Baseline Subset (SBAS) approach, patches of coherent data are

processed. In the Persistent/Permanent scatterer techniques, individual

objects (single rooftops, etc.) which are exceptionally good scatterers that stay

coherent over long times and spatial baselines are used instead.

The selection of the family of techniques to be used depends on data

quality, environment, deformation type, availability of scenes, and processing

time. The technique which we have focused on in the Nile Delta exercise is

the Persistent Scatterer techniques as outlined by Hooper (Hooper et al.,

2007; Hooper et al., 2004) and Kampes (Kampes, 2006). The following

summarizes the main characteristics of each of these techniques, as well as

the advantages and disadvantages of each method.

5.6 Summary

• Basic 2, 3 or 4 Pass DINSAR

o 2 to 4 scenes required

69

o Good coherence between scenes mandatory

o Atmospheric effects assumed negligible

o Two-pass

• Needs supplemental DEM information

• Subject to inaccuracies in DEM

• Pair needs to bracket deformation event

• DEM needs to be accurately radar coded

o Three-pass

• Subject to atmospheric effects

• One pair needs to be very close together in time (1 day) on

one side of the deformation event, the third scene has to

be on the other side of the event

• Difficult to maintain correlation over long times (years)

• Scenes registered to common master

o Four-pass

• 4 scenes, pairs of scenes from before and after the

deformation event

• Each pair needs small baseline, good correlation

• Allows for longer time for deformation to occur

• Scenes co-registered as pairs, pairs co-registered to each

other.

• Multi-temporal techniques:

o Small Baseline Techniques (SBAS)

• 10-20 scenes and more are used

• Assumes areas of good coherence in interferogram

• Entire scenes need not be coherent

• Scenes spatially resampled to one common scene, directly

or through cascading sequence (Refice et al., 2003)

• Only useful for gradual deformations (i.e., subsidence)

70

• Examples can be found in (Lanari et al., 2004a; Lanari et

al., 2004b)

• High number of interferograms generated

• Processing time intensive

o Permanent Scatterer Techniques

• >6 (Persistent Scatterer) to >40 (Permanent Scatterer)

scenes

• Good coherence at individual points

(Permanent/Persistent Scatterers)

• Deformation can be gradual (subsidence) or sharp

(faulting)

• Non-linear estimate of deformation

• Scenes spatially resampled to one common scene, directly

or through cascading sequence

• Very high number of interferograms generated

• Processing time intensive

• Examples include: (Ferretti et al., 2000, 2001; Hooper et

al., 2007; Hooper et al., 2004)

71

CHAPTER6

LAND SUBSIDENCE IN THE NILE DELTA: INFERENCES FROM RADAR INTERFEROMETRY

6.1 Abstract

The Nile Delta has formed by progression of a complex system of deltaic

fans throughout the Pleistocene, with the modern delta being formed from

sediments supplied by at least 10 distinct distributary channels throughout

the Holocene. With an average elevation of a meter or so above sea level and

a predicted rise in sea level of 1.8-5.9 mm/yr (IPCC, 2007) the subsidence of

the northern delta is becoming a topic of major concern both to the Egyptian

population and the government. The Nile Delta is home to more than 50

million people and the major agricultural production area for Egypt. We

evaluated the modern rates of subsidence of sections of the northeastern Nile

Delta (a total length of 110 km, up to 50 km from the coastline) using

persistent scatterer radar interferometry techniques applied to 14 ERS-1 and

ERS-2 scenes. The area covered includes the present active depocenter of the

Damietta promontories, and the nearby Mendesian depocenter that was

active up to recent times (up to 2500 years ago). The highest subsidence rates

( ~8 mm/yr; twice the average Holocene rates) do not correlate with the

distribution of the thickest Holocene sediments, but rather with the

72

distribution of the youngest depositional centers (major deposition occurred

between ~3500 years bp and present) at the terminus of the Damietta

branch. The adjacent slightly older (8000- 2500 years bp) Mendesian branch

depositional center is subsiding at slower rates of 2-6 mm/yr. Results are

interpreted to indicate that: (1) modern subsidence in the Delta is heavily

influenced by the compaction of the most recent sediments, and (2) the highly

threatened areas are at the terminus of the Damietta and possibly the

Rosetta branches, where active deposition is occurring.

6.2 Introduction

Deltas worldwide witness phases of progradation and destruction, phases

that are largely related to the interplay between fluvial processes, sea level

changes and coastal erosion at the mouth of the river. Deltas prograde when

fluvial processes dominate and retreat when coastal erosion intensifies. Many

of the world's large deltas are subsiding, in large part due to compaction and

isostatic response to loading by thick depositional sequences (Stanley et al.,

1996). Over the past 7000 years (Holocene) many deltas all over the world

have been witnessing an overall constructional phase. Recently (over the past

tens of years), there have been indications that this trend is being reversed in

a number of these deltas world-wide.

73

Modification of free-flowing rivers for energy generation or irrigation

projects can accelerate changes in delta plains and coasts including delta

subsidence due to the deprivation of the coastal systems of the silt and clay

which accumulate in deltas. Increased coastal erosion, and accelerated net

subsidence of deltas under the weight of the thick delta deposits combined

with rising sea level also exacerbates deltaic changes (Kay, 1993).

The Nile Delta (Fig. 6.1), the area of this study is witnessing a

30°

E

.,.. Alexandria

I .. ,,.Rosetta --

Branch Mahala

,..•.

0 25 50 75 100 N

-c:::=-K■il■o■mi=e =te=r=s=--• -i-

32°

E

31.5°

N

31°

N

30.5°

N

Figure 6.1 False color Landsat TM mosaic showing the Nile Delta, the Damietta and Rosetta branches, and urbanized centers (cities and villages) that appear in shades of purple. The thickest Holocene deposits are located north of the black dashed line (Stanley and Warne, 1993a). The red box

outlines the area covered in Figure 6.2. Inset (lower left corner) shows the location of Nile Delta, Nile River, and the Aswan High Dam.

74

destructional phase largely related to anthropogenic activities (damming,

water diversion, etc.) by the nations of the Nile Basin throughout the last

several centuries, with the most drastic changes occurring in the last 100

years. Egypt increased its water storage capacity by constructing dams (e.g.,

Aswan Dam, Aswan High Dam) and impounding water behind dams in

extensive artificial reservoirs (e.g., Lake Nasser: total capacity: 1.6 x 1011 m3)

(Said, 1993). Egypt is embarking on diversion projects (e.g., Toshka project,

El Salam Canal) to channel large volumes of Nile River water to reclaim

lands in the Western Desert and Sinai. As a result, water discharged at the

Rosetta and Damietta channels (Fig. 6.1) has been dramatically reduced

(1950s: 40 x 109 m3/yr; 2000s: < 5 x 109 m3/yr) (Frihy and Lawrence, 2004)

altering the balance that existed between sediment influx, sea level rise,

coastal erosion, and subsidence. For example, since 1972 the retreat of the

Damietta promontory has increased to alarming rates of up to 52 m/yr (Frihy

and Lawrence, 2004).

The pioneering work by Stanley, Warne, and co-workers (Krom et al.,

2002; Stanley and Warne, 1993b, 1998; Stanley, 2001; Stanley et al., 2004;

Warne and Stanley, 1993) on the subsidence of the Nile Delta throughout the

Holocene, and the accelerated subsidence of the area following the erection of

the Aswan High Dam and subsequent cut off of sediment supply, has received

a great deal of attention from the scientific community as well as extensive

media coverage. Eighty-seven radiocarbon-dated cores across the northern

75

delta were used to interpret the interactions between sea-level changes,

climatic oscillations, subsidence, and sediment transport. Average subsidence

rates (0.5 to 5 mm/yr for the past 7500 years) were calculated from

radiocarbon dated sediments from wells in the northern delta plain.

There are several uncertainties associated with using the average

Holocene subsidence rates to infer precise estimates of modern subsidence:

(1) subsidence rates for deltas change with time (Meckel et al., 2007) and

thus, average Holocene values do not necessarily represent shorter temporal

scales (modern) rates, (2) spatial heterogeneities in subsidence rates might

not be adequately captured using a limited number of samples collected from

widely spaced wells drilled in the northern delta, (3) uncertainties associated

with the 14C ages may be large as the position of dated samples in the

stratigraphic sequence may have been altered in many cases resulting in

underestimating the subsidence rates calculated from these samples (Said,

1993; Stanley and Hait, 2000; Stanley, 2001). Persistent scatterer radar

interferometry techniques were applied over large sections of the

northeastern sections of the Delta to overcome these problems.

6.3 Geology of the Nile Delta

The modern Nile Delta sits on the Nile Cone, a depositional feature

resulting from more than 5 million years of discharge and deposition of 3500

76

m of sediment by the Nile and Paleonile drainage systems (Sestini, 1989).

The Pleistocene section consists of up to 800 m of deltaic sands with minor

clay layers (RIGW, 1992). During the late Pleistocene, from roughly 35 to 18

ka, the region of the modern Nile Delta was a seasonally active alluvial plain

with braided stream channels. This was followed between 15 and 8 ka by a

period of rapid sea level rise, which reworked much of these sediments. Some

8000 years ago, a slowing down in sea level rise prompted the development of

the modern Nile Delta (Said, 1990; 1993). The Delta prograded up to 10 m/yr

into the Mediterranean through accretion of silts and clays (1-7mm/yr) giving

rise to thicknesses of up to 60 m of Holocene deltaic fluvial/marine deposits

(Stanley and Warne, 1993a).

Throughout the Holocene the centers of deposition continued to shift

from one spot to another along the delta coastline depending on which

channel(s) were active at any particular time (Stanley et al., 2004). For

example, from 8000 years ago, the eastern sections of the Nile delta were

building out along the Tanitic and Mendesian branches (Fig. 6.2a). The

Tanitic branch deposited material underneath what is now the Manzala

Lagoon, and prograded across to present day Port Said (Fig. 6.2). The

Mendesian branch formed a depositional center under the western portion of

the Manzala Lagoon and prograded to the northeast to merge with the

depositional center of the Tani tic branch. The supply of sediments by these

two branches gave rise to up to 40 m thick sections of Holocene deposits,

77

currently underlying the Manzala Lagoon. Around 1500 to 2500 years ago

these two branches silted up allowing the Damietta branch to become the

largest contributor of Nile sediment flux (Stanley and Warne, 1993a).

In addition to natural phenomena, and as early as 5000 bp, human

intervention has played an increasing role in modifying the natural flow

system and flood patterns. By the middle of the 19th century, earlier smaller

individual projects (e.g., raised levees) quickly gave way to larger irrigation

canal systems to irrigate large sectors of the Nile Delta (Sestini, 1989). Up to

that time, the annual floods still delivered considerable silt to the delta, a

situation that rapidly changed beginning in the late 19th century and

continued throughout the 20th century. Major engineering projects were

implemented including dams and barrages culminating in the construction of

the Aswan High Dam in 1964. Prior to the construction of the latter,

approximately 9.5x106 metric tons of sediment were deposited onto the Nile

Delta floodplain each year, an amount equivalent to a ~ 1 mm layer of silt

across the Delta. This added thickness partially balanced continued

subsidence of the Delta. Following the construction of the Aswan High Dam,

the sediment load no longer reaches the delta, instead it settles in Lake

Nasser behind the dam (Said, 1993).

78

6.4 Radar Interferometry

Radar interferometry has been successfully used to map small (as

small as 0.1 mm/yr) (Massonnet and Feigl, 1998) deformation and fault slip

from earthquakes (Sandwell et al., 2002), mine subsidence (Carnec and

Delacourt, 2000), aquifer compaction from pumping (Burbey, 2003), and

landslides (Amelung and Day, 2002), as well as seasonal changes due to

groundwater extraction and recharge (Hoffmann et al., 2001).

Interferometric synthetic aperture radar (InSAR) processing makes

use of the difference in phase between two radar scenes to determine precise

differences in range to a target and to subsequently determine the exact

surface location, and subtle changes in topography. In conventional three­

pass InSAR interferometry, the deformation is calculated from the difference

between the unwrapped phase of a reference interferogram (no deformation

between scene acquisitions) and a pair spanning the deformation (Massonnet,

1996). Successful implementation of this method relies on maintaining a high

correlation for pairs of scenes over the time period of the deformation. If

changes in scatterers occur between passes, such as those introduced by

growing vegetation, phase will vary randomly and the process of extracting

phase differences indicative of subsidence will be jeopardized. Because the

Nile Delta is highly vegetated, and the variations we are examining are long

term variations compared to those introduced by growing vegetation, the

79

conventional three-pass interferometry technique was inadequate for the

study area (Aly et al., 2005).

We took advantage of recent refinements in persistent scatterer radar

interferometry techniques (Hooper et al., 2007; Hooper et al., 2004) that are

now enabling successful applications over a wider-range of physiographic and

atmospheric conditions. To date, applications of the related permanent

scatterer techniques in the Nile Delta have been restricted to two major

cities, Mahala and Mansura, in central Nile Delta (Fig. 6.1) (Aly et al., 2005).

We applied the persistent scatterer technique (Hooper et al., 2007; Hooper et

al., 2004) to investigate the spatial variations in subsidence rates across large

sections (area: ~2400 km2) of the heavily vegetated northeastern delta (Fig.

6.2). The adopted persistent scatterer method generates multiple three-pass

interferograms, but restricts the phase unwrapping and analysis to pixels

containing individual scatterers which dominate individual pixels, and

remain stable over the time period of interest. Because the signal from these

scatterers is much larger than the random noise from a large number of

small scatterers, the variance in phase in these pixels reflects the underlying

deformation (Ferretti et al., 2000; Hooper et al., 2004). In the northern delta,

these scatterers include buildings, individual outcrops, rocks, or utility poles.

In applying this technique, we took advantage of the extensive distribution of

small buildings and scattered dwellings, present largely in villages and cities

(areas appearing in shades of violet in Fig. 6.1) across the delta. The Stanford

80

Method for Persistent Scatterers (StAMPS) algorithm was adopted; this

method uses the spatial correlation of the interferogram phase to identify

such pixels (Hooper et al., 2007). For the majority of the investigated area,

the density of the permanent scatters ranged from 4 to 30 permanent

scatterers/km2 enabling the estimation of atmospheric contributions

(Colesanti et al., 2003).

Synthetic aperture radar (SAR) scenes covering the period 1992-1999

were acquired from the European Remote Sensing satellites (ERS-1 and

ERS-2; Table 6.1). In the selection of scenes, attempts were made to pick

scenes with minimal differences in their spatial baselines (perpendicular

baseline <500m) and which are temporally as evenly distributed as possible

throughout the investigated period. Fourteen scenes in track 436 were

processed, and mean line of sight velocities were first calculated and then

converted to equivalent vertical motions.

ERS precise orbit information was obtained from Delft Institute for Earth­

Oriented Space Research (Scharroo and Visser, 1998). All of the

interferometric processing was conducted using: (1) the Delft Object-oriented

Radar Interferometric Software (DORIS) (Kampes et al., 2003), (2) the

Repeat Orbit Interferometry Package (ROI_PAC) from JPL, and (3) StAMPS

(Hooper et al., 2004).

81

Temporal Acquisition Perpendicular baseline Date Satellite Baseline (m) (days) Track

19920601 ERS-1 331 -1179 436

19930517 ERS-1 60 -829 436

19931004 ERS-1 77 -689 436

19931108 ERS-1 300 -654 436

19950720 ERS-2 -35 169 436

19950823 ERS-1 100 -1 436

19950824 ERS-2 0 0 436

19960111 ERS-2 129 140 436

19960530 ERS-2 -341 280 436

19970515 ERS-2 150 630 436

19980709 ERS-2 -333 1050 436

19980813 ERS-2 186 1085 436

19990311 ERS-2 64 1295 436

19990520 ERS-2 -61 1365 436

19991007 ERS-2 -190 1505 436

Table 6.1 ERS-1 and ERS-2 scenes used in permanent scatterers study. Temporal and perpendicular baselines are shown relative to the scene acquired on 8/24/1995.

6.5 Discussion and Results

Results of permanent scatterer analysis applied to the 13 ascending

scenes are shown in figure 6.2a. A large number (>10,000) of persistent

scatterers were identified and used to constrain modern subsidence rates in

the study area. Subsidence rates represented on this figure are relative to the

average motion of the permanent scatterers in the southwestern portion of

82

31.5°

E

Total Holocene Thickness (After Stanley, 1996)

• • •

Kilorneters -- -

0 4 8 12 16

Modern Relatlve Motion (mm/yr) Persistent Scatterer

■ ■ D

32°

E

Average Holocene Subsidence (After Stanley, 1993a)

Elevation (m)

SRTM 3 Arc Second

31.5°

N

31°

N

Figure 6.2 (a) Mean velocities extracted from 14 ascending ERS-1 and ERS-2 scenes acquired over the period 1992-1999 and expressed as vertical motion of permanent scatterers (negative velocities indicate downward motion). Areas of highest modern subsidence include areas at the terminus of the Damietta branch that are underlain by significant (>5 m) amounts of young (<3500 years) sediments (areas shaded in white). Also shown in white solid lines are the current Damietta branch, the paleo Mendesian and Tanitic branches, and shown in black dashed lines, the paleo-coastlines positions at 4000, 3500, 2000 and 1500 years bp (Coutellier and Stanley, 1987; Stanley and

Warne, 1993a). (b) Average Holocene subsidence (Stanley and Warne, 1993a), and Holocene thicknesses from individual borings (colored squares) (Stanley et al., 1996). High modern subsidence rates correlate with the distribution of younger (<3500 yr) Holocene sediments rather than thick Holocene sediments. Radar data provided by the European Space Agency.

83

the study area, an area that experienced minimal subsidence (<0.5mm/yr)

(Stanley and Warne, 1993a) throughout the Holocene. Previous estimates of

average Holocene subsidence rates were based on data from 87 wells across

the northern delta, 22 of which were drilled in the study area (Fig. 6.2b). The

subsidence rates that we report and their spatial distribution differ

considerably from those previously reported for the Holocene (Emery et al.,

1988; Stanley and Warne, 1993a). Modern rates are generally high (up to 8

mm/yr) compared to previously calculated average Holocene rates (0.5 to 4.5

mm/yr). The highest modern rates are observed over the youngest (<3500

year old) sediments under the terminus of the Damietta branch (Fig. 6.2a).

Moderate rates (4-6 mm/yr) are calculated around the Manzala Lagoon (Fig.

6.2), a depocenter that was active more than 3500 years ago.

The highest subsidence rates were calculated at the terminus of the

Damietta branch suggesting a causal relationship between rapid subsidence

and the distribution of young sediments that were probably deposited during

the recent progradation of the Damietta branch and have been undergoing

compaction and isostatic subsidence since then. This hypothesis is supported

by the general spatial correspondence between rapidly subsiding areas and

those underlain by thick, young ( <3500 years old) sediments. Figure 6.2

shows a correspondence between the area subtended by a contour defining

domains of 5 or more meter thick young deposits and the areas undergoing

rapid subsidence (clusters of orange, red, and yellow dots on Fig. 6.2a). The

84

contour was defined using data extracted from radiocarbon dated cores and

sediment thicknesses from 22 drilled wells in the study area (Stanley et al.,

1996). Only cores where the 14C age sequence increases sequentially with

depth were used. The locations of these cores are represented as squares color

coded to total Holocene depth in figure 6.2b.

The proposed hypothesis is consistent with the documented evolution of

the depositional environments over the past 3500 years in the northeast

delta. The Damietta branch has prograded some 40 km to the northeast over

the past 3500 years as the coastline has migrated to the north (Fig. 6.2a). The

highest modern subsidence rates (6-8 mm/yr) correspond to areas of this most

recent deposition along the Damietta branch between the paloeshoreline of

3500 years bp (Coutellier and Stanley, 1987) and the modern shoreline (Fig.

6.2a).

The proposed hypothesis is also consistent with the documented evolution

of the depositional environments for the time period preceding the past 3500

years. If modern subsidence is largely controlled by compaction of younger

sediments, one should expect to observe moderate subsidence rates in areas

of slightly older (>3500 years old) sediments that were transported by the

earlier Mendesian and Tanitic branches and deposited in the area occupied

by Manzala Lagoon. In these areas, only a thin cover of younger sediments

(1-4 meters) were deposited by the earlier Mendesian and Tanitic branches

85

(Stanley et al., 1996). Our results indicate that this is indeed the case;

moderate (2-6 mm/yr) subsidence rates are observed in the area occupied by

the Mendesian depocenter (area outlined by dashed white line: Fig. 6.2a).

Unfortunately the examined radar scene does not cover the Tanitic

depocenter. We suggest that these large thicknesses of the Holocene

sediments in the Manzala Lagoon must have witnessed early periods of rapid

compaction and subsidence followed by a progressive decline in subsidence

rates with time as sediments became less prone to compaction. Such an

interpretation could explain why the highest average Holocene rates are

reported for areas with the thickest Holocene sediment in the Manzala

Lagoon (Stanley and Warne, 1993a), whereas the modern subsidence rates

we report for these areas are moderate.

It is unlikely that the observed modern subsidence was caused by tectonic

movements, as there have not been any reported earthquakes (ISC, 2001)

along the previously identified faults (Frihy, 2003; Stanley, 1988) in the area

or in any other location within the study area throughout the investigated

period (1992-1999). Our findings do not preclude seismic activities occurring

on larger timescales. The lack of any recent (1992-1996) seismic activities

also argues against the observed displacements being related to horizontal

motion along faults. The observed modern displacements are interpreted here

to be vertical in nature and are largely attributed to compaction and isostatic

adjustment.

86

Our findings in the Nile Delta are consistent with those reported from

deltas elsewhere. Modern subsidence rates in the Mississippi are interpreted

to indicate that compaction of sediments constitutes a substantial source of

subsidence in deltaic systems, especially during the earlier periods following

deposition, and that average Holocene subsidence rates do not necessarily

predict current subsidence rates. Compaction rates in the Mississippi Delta

system within the first tens of years to centuries can range from 5 mm/yr to

values as high as 10 mm/yr in organic rich peat systems, considerably higher

than the estimated average Holocene subsidence rates of <3 mm/yr

(Tornqvist et al., 2008). Mekel et al. (2007) modeled the compaction of

Holocene sediments in the Mississippi Delta; his simulations predicted rapid

compaction of deposits shortly after deposition, with progressive decline in

compaction with increasing age of sediments.

In the northern delta, it has been assumed that the areas underlain by

the highest sediment thickness are currently subsiding the fastest, and would

be impacted most severely by projected sea level rises. This study shows that

the sediment age rather than the thickness might be the dominant factor

controlling subsidence in the Nile Delta. Results indicate that areas around

the city and harbor of Damietta may be the most vulnerable to inundation.

The conditions described for the Nile Delta are typical of many of the world's

deltas.

87

CHAPTER7

CONCLUSIONS

In this manuscript, I demonstrated methodologies and applications

addressing two specific problems pertaining to mapping the abundance of

algal blooms in the Great Lakes, and land subsidence in the Nile Delta.

Specifically, I investigated methodologies to temporally and spatially

investigate the extent and distribution of algal blooms in Lake Erie and Lake

Ontario, and proposed a method for discriminating between potentially toxic

and non-toxic blooms. Visible and near infra red MO DIS and Sea WiFS

satellite data were used to map the distribution of algal blooms in these

lakes. Existing algorithms for mapping blooms were applied and results

were compared. Moreover, new methodologies were developed to distinguish

between the non-toxic blooms and the potentially toxic cyanobacterial blooms.

Cyanobacterial concentrations that were derived using my algorithms were

successfully tested against in situ measurements.

These results show great promise for establishing cost-effective,

remotely-based monitoring systems which could be used by various

municipalities to secure water intake systems and the recreational areas in

Lakes Erie and Ontario. To achieve this goal, we now have in place an

automated system which produces chlorophyll images (using case II Carder

algorithms applied to MODIS data) and posts the results on the web

(http://www.esrs.wmich.edu/MERHAB-LGL) for all interested parties

Future steps will include automation of the cyanobacterial

88

identification method to provide rapid access to results and to make these

available to scientists and city managers. Better refinement of the spectral

properties of the endmembers is also planned. Currently, Microcystis is the

only cyanobacteria included in the model. Plans are underway to better

represent the known variations in toxic species in the lakes, including

Planktothrix, and Anabaena. The model should be also tested and refined in

other locations in the Great Lakes (e.g., Saginaw Bay) and in other settings

where cyanobacterial blooms were reported (e.g., Baltic Sea).

In the second study I examined subsidence in the northeastern portion

of the Nile Delta. Because of its low average elevation, and a predicted rise in

sea level of 1.8-5.9 mm/year (IPCC, 2007) the potential for inundation of this

region of the delta is high. This provides a significant concern to the Egyptian

population and government. I calculated modern subsidence rates in the

northeastern Nile Delta using persistent scatterer radar interferometry

techniques, and found that the modern subsidence rates are as high as 8

mm/year. The highest rates are twice the average Holocene rates, and

correlate with the distribution of the youngest deposits. Slower rates of 2-6

mm/yr. were found under the older depositional centers. These results were

interpreted to indicate that the modern subsidence in the Delta is heavily

influenced by the compaction of the most recent sediments, and that the

highly threatened areas are at the terminus of the Damietta branch, where

the most recent deposition has occurred.

This information about the ongoing modern distribution of subsidence

rates is useful for Egyptian officials in the Damietta region. Prior to my work

it had been assumed that the maximum subsidence rates were about 5

89

mm/yr, approximately half of the maximum rate we report in this study.

Moreover, the previous studies predicted that the highest subsidence rates

were underneath the Manzala Lagoon, whereas our study shows that it is

around the Damietta region. Naturally, the high subsidence rates under

Damietta city and its surroundings raises the concern of inundation in this

area.

Future work in this area should include verification of these

calculations. Geodetic stations could be established in or near Damietta city

and the harbor. The radar interferometry study should be extended both

spatially to cover larger portions of the northern Nile Delta region, and

temporally, to include data later than 1999. Future studies could take

advantage of the L-band radar instrumentation on the ALOS satellite. The

later provides better penetration of vegetation canopy, and could potentially

provide more coherent applications compared to those based on the C band of

ERS-1 and ERS-2. As longer acquisition periods of radar data are achieved

better precision in results will be accomplished.

90

REFERENCES

Abalde, J., Betanocourt, L., Torres, E., Cid, A., and Barwell, C., 1998,

Purification and characterization of phycocyanin from the marine

cyanobacterium Synechococcus: Plant Science, v. 16, p. 109-120.

Aly, M.H., Giardino, J.R., and Klein, A.G., 2005,-Local Subsidence in Greater

Mahala, Egypt, Detected by Radar Interferometry: Eos Trans. AGU,

Fall Meet. Suppl., v. 86.

Amelung, F., and Day, S., 2002, InSAR observations of the 1995 Fogo, Cape

Verde, eruption: Implications for the effects of collapse events upon

island volcanoes: Geophysical Research Letters, v. 29.

Anderson, D.M., Glibert, P.M., and Burkholder, J.M., 2002, Harmful algal

blooms and eutrophication: Nutrient sources, composition, and

consequences: Estuaries, v. 25, p. 704-726.

Aydar, E., Gourgaud, A., Ulusoy, I., Digonnet, F., Labazuy, P., Sen, E.,

Bayhan, H., Kurttas, T., and Tolluoglu, A.U., 2003, Morphological

analysis of active Mount Nemrut stratovolcano, eastern Turkey:

evidences and possible impact areas of future eruption: Journal of

Volcanology and Geothermal Research, v. 123, p. 301-312.

Becker, R., Sultan, M., Atkinson, J., Boyer, G.L., and Konopko, E., 2005,

Spatial and Temporal Variations of Algal Blooms in the Lower Great

91

Lakes, 8th International Conference on Remote Sensing for Marine

and Coastal Environment: Halifax, Nova Scotia, Altarim.

Becker, R., Sultan, M., Boyer, G., and Konopko, E., 2006, Mapping Variations

Of Algal Blooms In The Lower Great Lakes: Great Lakes Research

Reviews, v. 7, p. 14-17.

Belward, A.S., Stibig, H.J., Eva, H., Rembold, F., Bucha, T., Hartley, A.,

Beuchle, R., Khudhairy, D., Michielon, M., and Mollicone, D., 2007,

Mapping severe damage to land cover following the 2004 Indian Ocean

tsunami using moderate spatial resolution satellite imagery:

International Journal of Remote Sensing, v. 28, p. 2977-2994.

Berardino, P., Costantini, M., Franceschetti, G., Iodice, A., Pietranera, L.,

and Rizzo, V., 2003, Use of differential SAR interferometry in

monitoring and modelling large slope instability at Maratea

(Basilicata, Italy): Engineering Geology, v. 68, p. 31-51.

Beutler, M., Wilstshire, K.H., Arp, M., Kruse, J., Reineke, C., Moldaenke, C.,

and Hansen, U.-P., 2003, A reduced model of the fluorescence from the

cyanobacterial photosynthetic apparatus designed for the in situ

detection of cyanobacteria: Bioch. Bioph. Acta v. 1604, p. 33-46.

Beutler, M., Wilstshire, K.H., Meyer, B., Moldaenke, C., Li.iring, C.,

Meyerhofer, M., Hansen, U.-P., and Dau, H., 2002, A fluorometric

method for the differentiation of algal populations in vivo and in situ:

Photosyn. Res, v. 72, p. 39-53.

92

Binding, C.E., Jerome, J.H., Bukata, R.P., and Booty, W.G., 2008, Spectral

absorption properties of dissolved and particulate matter in Lake Erie:

Remote Sensing of Environment, v. 112, p. 1702-1711.

Boyer, G., 2006, Toxic Cyanobacteria in the Great Lakes: More than Just the

Western Basin of Lake Erie: Great Lakes Research Reviews, v. 7, p. 2-

7.

Boyer, G.L., Satchwell, M.F., Hotto, Y., A., X., Lewis, T., Oles, M.,

Makarewicz, J.C., and Watson, S.B., 2004, Occurrence of

Cyanobacterial Toxins in Lake Ontario, USA, 6th Internat Conf. Toxic

Cyanobacteria: Bergen Norway.

Brittain, S.M., Wang, J., Babcock-Jackson, L., Carmichael, W.W., Rinehart,

K.L., and Culver, D.A., 2000, Isolation and characterization of

microcystins, cyclic heptapeptide hepatotoxins from a Lake Erie stain

of Microcystis aeruginosa: J. Great Lakes Res, v. 26, p. 241-249.

Bryant, D.A., 1981, The Photoregulated Expression of Multiple Phycocyanin

Species - A General Mechanism for the Control of Phycocyanin

Synthesis in Chromatically Adapting Cyanobacteria: European

Journal of Biochemistry, v. 119, p. 425-429.

Bukata, R.P., Jerome, J.H., Bruton, J.E., Jain, S.C., and Zwick, H.H., 1981,

Optical Water-Quality Model of Lake-Ontario .1. Determination of the

Optical-Cross-Sections of Organic and Inorganic Particulates in Lake­

Ontario: Applied Optics, v. 20, p. 1696-1703.

93

Bukata, R.P., Jerome, J.H., Kondratyev, K.Y., and Pozdnyakov, D., 1998,

Optical properties and remote sensing of inland and coastal waters:

Boca Raton Fla, C. R. C. Press.

Burbey, T.J., 2003, Use of time-subsidence data during pumping to

characterize specific storage and hydraulic conductivity of semi­

confining units: Journal of Hydrology, v. 281, p. 3-22.

Campbell, J., Antoine, D., Armstrong, R., Arrigo, K., Balch, W., Barber, R.,

Behrenfeld, M., Bidigare, R., Bishop, J., Carr, M.E., Esaias, W.,

Falkowski, P., Hoepffner, N., Iverson, R., Kiefer, D., Lohrenz, S.,

Marra, J., Morel, A., Ryan, J., Vedernikov, V., Waters, K., Yentsch, C.,

and Yoder, J., 2002, Comparison of algorithms for estimating ocean

primary production from surface chlorophyll, temperature, and

irradiance: Global Biogeochemical Cycles, v. 16.

Carder, K.L., Chen, F.R., Lee, Z.P., Hawes, S.K., and Kamykowski, D., 1999,

Semianalytic Moderate-Resolution Imaging Spectrometer algorithms

for chlorophyll a and absorption with bio-optical domains based on

nitrate-depletion temperatures: Journal of Geophysical Research­

Oceans, v. 104, p. 5403-5421.

Carnec, C., and Delacourt, C., 2000, Three years of mining subsidence

monitored by SAR interferometry, near Gardanne, France: Journal of

Applied Geophysics, v. 43, p. 43-54.

94

Chen, C.W., and Zebker, H.A., 2002, Phase unwrapping for large SAR

interferograms: Statistical segmentation and generalized network

models: IEEE Transactions on Geoscience and Remote Sensing, v. 40,

p. 1709-1719.

Clarke, G.L., Ewing, G.C., and Lorenzen, C.J., 1970, Spectra of backscattered

light from the sea from aircraft as a measure of chlorophyll

concentration.: Science, v. 167, p. 1119-1121.

Colesanti, C., Ferretti, A., Novali, F., Prati, C., and Rocca, F., 2003, SAR

monitoring of progressive and seasonal ground deformation using the

permanent scatterers technique: IEEE Transactions on Geoscience and

Remote Sensing, v. 41, p. 1685-1701.

Coutellier, V., and Stanley, D.J., 1987, Late Quaternary Stratigraphy and

Paleogeography of the Eastern Nile Delta, Egypt: Marine Geology, v.

77, p. 257-275.

Curran, P.J., Dash, J., and Llewellyn, G.M., 2007, Indian Ocean tsunami:

The use of MERIS (MTCI) data to infer salt stress in coastal

vegetation: International Journal of Remote Sensing, v. 28, p. 729-735.

Davies-Colley, R., Vant, W.N., and Smith, D.G., 2003, Colour and Clarity of

Natural Waters: Science and Management of Optical Water Quality:

Caldwell, NJ, Blackburn Press.

95

Dekker, A.G., Malthus, T.J., and Seyhan, E., 1991, Quantitative modeling of

inland water-quality for highresolution mss systems: IEEE

Transactions on Geoscience and Remote Sensing, v. 29, p. 89-95.

Downing, J.A., Watson, S.B., and McCauley, E., 2001, Predicting

cyanobacteria dominance in lakes: Canadian Journal of Fisheries and

Aquatic Sciences, v. 58, p. 1905-1908.

Dyble, J., Fahnenstiel, G.L., Litaker, R.W., Millie, D.F., and Tester, P.A.,

2008, Microcystin concentrations and genetic diversity of Microcystis in

the lower Great Lakes: Environmental Toxicology, v. 9999, p. NA.

Emery, K.O., Aubrey, D.G., and Goldsmith, V., 1988, Coastal Neo-Tectonics

of the Mediterranean from Tide-Gauge Records: Marine Geology, v. 81,

p. 41-52.

Fay, P., 1983, The Blue-Greens: London, Arnold.

Ferretti, A., Prati, C., and Rocca, F., 2000, Nonlinear subsidence rate

estimation using permanent scatterers in differential SAR

interferometry: Ieee Transactions on Geoscience and Remote Sensing,

v. 38, p. 2202-2212.

-, 2001, Permanent scatterers in SAR interferometry: IEEE Transactions on

Geoscience and Remote Sensing, v. 39, p. 8-20.

Frihy, 0., 2003, The Nile Delta-Alexandria Coast: Vulnerability to Sea-Level

Rise, Consequences and Adaptation: Mitigation and Adaptation

Studies for Global Change, v. 8, p. 115-138.

96

Frihy, 0., and Lawrence, D., 2004, Evolution of the modern Nile delta

promontories: development of accretional features during shoreline

retreat: Environmental Geology, v. 46, p. 914-931.

Fu, L.L., and Smith, R.D., 1996, Global ocean circulation from satellite

altimetry and high-resolution computer simulation: Bulletin of the

American Meteorological Society, v. 77, p. 2625-2636.

Gabriel, A.K., Goldstein, R.M., and Zebker, H.A., 1989, Mapping Small

Elevation Changes over Large Areas - Differential Radar

Interferometry: Journal of Geophysical Research-Solid Earth and

Planets, v. 94, p. 9183-9191.

Gons, H.J., Kromkamp, J., Rijkeboer, M., and Schofield, 0., 1992,

Characterization of the Light-Field in Laboratory Scale Enclosures of

Eutrophic Lake Water (Lake Loosdrecht, the Netherlands):

Hydrobiologia, v. 238, p. 99-109.

Gordon, H.R., 1997, Atmospheric correction of ocean color imagery in the

Earth Observing System era: Journal of Geophysical Research­

Atmospheres, v. 102, p. 17081-17106.

Gordon, H.R., Brown, O.B., and Jacobs, M.M., 1975, Computed relationships

between the inherent apparent optical properties of a flat

homogeneous ocean: Applied Optics, v. 14, p. 417-427.

97

Gordon, H.R., and Clark, D.K., 1981, Clear Water Radiances for Atmospheric

Correction of Coastal Zone Color Scanner Imagery: Applied Optics, v.

20, p. 4175-4180.

Gordon, H.R., and Wang, M.H., 1994, Retrieval of Water-Leaving Radiance

and Aerosol Optical-Thickness over the Oceans with Seawifs - a

Preliminary Algorithm: Applied Optics, v. 33, p. 443-452.

Hallegraeff, G.M., 1993, A Review of Harmful Algal Blooms and their

Apparent Global Increase: Phycologia, v. 32, p. 79-99.

Havens, K.E., 2008, Cyanobacteria blooms: effects on aquatic ecosystems, in

Hudnell, H.K., ed., Cyanobacterial Harmful Algal Blooms: State of the

Science and Research Needs: Advances in Experimental Medicine and

Biology: New York, NY, Springer, p. 950.

Hilley, G.E., Burgmann, R., Ferretti, A., Novali, F., and Rocca, F., 2004,

Dynamics of slow-moving landslides from permanent scatterer

analysis: Science, v. 304, p. 1952-1955.

Hoffmann, J., Zebker, H.A., Galloway, D.L., and Amelung, F., 2001, Seasonal

subsidence and rebound in Las Vegas Valley, Nevada, observed by

synthetic aperture radar interferometry: Water Resources Research, v.

37, p. 1551-1566.

Hooker, S.B., and McClain, C.R., 2000, The calibration and validation of

SeaWiFS data: Progress in Oceanography, v. 45, p. 427-465.

98

Hooper, A., Segall, P., and Zebker, H., 2007, Persistent scatterer

interferometric synthetic aperture radar for crustal deformation

analysis, with application to Volcan Alcedo, Galapagos: Journal of

Geophysical Research-Solid Earth, v. 112.

Hooper, A., Zebker, H., Segall, P., and Kampes, B., 2004, A new method for

measuring deformation on volcanoes and other natural terrains using

InSAR persistent scatterers: Geophysical Research Letters, v. 31.

IGOS, 2003, IGOS Geohazards Theme Report.

IPCC, 2007, Climate Change 2007: The Physical Science Basis. Contribution

of Working Group I to the Fourth Assessment Report of the

Intergovernmental Panel on Climate Change in Solomon, S., Qin, D.,

Manning, M., Chen, Z., Marquis, M., Averyt, K.B., M.Tignor, and

Miller, H.L., eds.: Cambridge, United Kingdom and New York, NY,

USA., Cambridge University Press.

ISC, 2001, International Seismological Centre On-line Bulletin,

http://www.isc.ac.uk, Volume 2008: Thatcham, United Kingdom,

Internatl. Seis. Cent.

Jupp, D.L.B., Kirk, J.T.O., and Harris, G.P., 1994, Detection, Identification

and Mapping of Cyanobacteria - Using Remote-Sensing to Measure the

Optical-Quality of Turbid Inland Waters: Australian Journal of Marine

and Freshwater Research, v. 45, p. 801-828.

99

Kahru, M., and Mitchell, B.G., 1998, Spectral reflectance and absorption of a

massive red tide off southern California: Journal of Geophysical

Research-Oceans, v. 103, p. 21601-21609.

Kampes, B., 2006, Radar Interferometry: Persistent Scatterer Technique,

Springer.

Kampes, B.M., Hanssen, R., and Perski, Z., 2003, Radar Interferometry with

Public Domain Tools, Fringe 2003: Frascati, Italy.

Kanoshina, i., Lips, U., and Leppanen, J., 2003, The influence of weather

conditions (temperature and wind) on cyanobacterial bloom

development in the Gulf of Finland (Baltic Sea): Harmful Algae, v. 2, p.

29-41.

Kay, R., 1993, Deltas of the World: New York, American Society of Civil

Engineers, p. 138.

Kerle, N., Froger, J.L., Oppenheimer, C., and De Vries, B.V., 2003, Remote

sensing of the 1998 mudflow at Casi ta volcano, Nicaragua:

International Journal of Remote Sensing, v. 24, p. 4791-4816.

Kirk, J.T.O., 1994, Light & Photosynthesis in Aquatic Ecosystems:

Cambridge, Cambridge University Press.

Konopko, E., 2007, Development of a Flow-through Fluorometric System for

the Detection of Phycocyanin in the Lower Great Lakes: Syracuse, NY,

State University of New York College of Environmental Science and

Forestry.

100

Krom, M.D., Stanley, J.D., Cliff, R.A., and Woodward, J.C., 2002, Nile River

sediment fluctuations over the past 7000 yr and their key role in

sapropel development: Geology, v. 30, p. 71-74.

Kutser, T., Metsamaa, L., Strombeck, N., and Vahtmae, E., 2006a,

Monitoring cyanobacterial blooms by satellite remote sensing:

Estuarine Coastal and Shelf Science, v. 67, p. 303-312.

Kutser, T., Metsamaa, L., Vahtmae, E., and Strombeck, N., 2006b, Suitability

of MOD IS 250 m resolution band data for quantitative mapping of

cyanobacterial blooms: Proceedings of the Estonian Academy of

Sciences Biology Ecology, v. 55, p. 318-328.

Lanari, R., Mora, 0., Manunta, M., Mallorqui, J.J., Berardino, P., and

Sansosti, E., 2004a, A small-baseline approach for investigating

deformations on full-resolution differential SAR interferograms: IEEE

Transactions on Geoscience and Remote Sensing, v. 42, p. 1377-1386.

Lanari, R., Zeni, G., Manunta, M., Guarino, S., Berardino, P., and Sansosti,

E., 2004b, An integrated SAR/GIS approach for investigating urban

deformation phenomena: a case study of the city of Naples, Italy:

International Journal of Remote Sensing, v. 25, p. 2855-2862.

Land, P.E., and Haigh, J.D., 1996, Atmospheric correction over case 2 waters

with an iterative fitting algorithm: Applied Optics, v. 35, p. 5443-5451.

Lawson, C.L., and Hanson, R.J., 1974, Solving Least Squares Problems:

Englewood Cliffs, NJ, Prentice Hall.

101

Lee, Z.P., Carder, K.L., and Arnone, R.A., 2002, Deriving inherent optical

properties from water color: a multiband quasi-analytical algorithm for

optically deep waters: Applied Optics, v. 41, p. 5755-5772.

Lesht, B.M., Stroud, J.R., McCormick, M.J., Fahnenstiel, G.L., Stein, M.L.,

Welty, L.J., and Leshkevich, G.A., 2002, An event-driven

phytoplankton bloom in southern Lake Michigan observed by satellite:

Geophysical Research Letters, v. 29.

Lu, Z., and Danskin, W.R., 2001, InSAR analysis of natural recharge to

define structure of a ground-water basin, San Bernardino, California:

Geophysical Research Letters, v. 28, p. 2661-2664.

Lynch, P., 2008, The origins of computer weather prediction and climate

modeling: Journal of Computational Physics, v. 227, p. 3431-3444.

Makarewicz, J.C., Boyer, G.L., Guenther, W., Arnold, M., and Lewis., T.W.,

2006, The occurrence of cyanotoxins in the nearshore and coastal

embayments of Lake Ontario: Great Lakes Research Reviews, v. 7, p.

25-29.

Maritorena, S., Siegel, D.A., and Peterson, A.R., 2002, Optimization of a

semianalytical ocean color model for global-scale applications: Applied

Optics, v. 41, p. 2705-2714.

Martin, S., 2004, An introduction to Ocean Remote Sensing: Cambridge, UK,

Cambridge University Press, 426 p.

102

Mason, P.J., and Rosenbaum, M.S., 2002, Geohazard mapping for predicting

landslides: an example from the Langhe Hills in Piemonte, NW Italy:

Quarterly Journal of Engineering Geology and Hydrogeology, v. 35, p.

317-326.

Massonnet, D., 1996, Tracking the Earth's surface at the centimetre level: an

introduction to radar interferometry: Nature & Resources, v. 32, p. 20-

29.

Massonnet, D., and Feigl, K.L., 1998, Radar interferometry and its

application to changes in the earth's surface: Reviews of Geophysics, v.

36, p. 441-500.

Meckel, T.A., Ten Brink, U.S., and Williams, S.J., 2007, Sediment compaction

rates and subsidence in deltaic plains: numerical constraints and

stratigraphic influences: Basin Research, v. 19, p. 19-31.

Morel, A., and Prieur, L., 1977, Analysis of variations in ocean color:

Limnology and Oceanography, v. 22, p. 709-722.

Muller-Karger, F.E., Chuanmin, H., Andrefouet, S., Varela, R., and Thunell,

R., 2005, The Color of the Coastal Ocean and Applications in the

Solution of Research and Management Problems, in Miller, R.L., Del

Castillo, C., and McKee, B.A., eds., Remote Sensing of Coastal Aquatic

Environments, Volume 7: Remote Sensing and Digital Image

Processing: Dordrecht, The Netherlands, Springer, p. 345.

103

NASA, 2006, Earth Science Reference Handbook: A guide to NASA's Earth

Science Program and Earth Observing Satellite Mission, in Parkinson,

C.L., Ward, A., and King, M.D., eds.: Washington D.C., National

Aeronautics and Space Administration.

O'Reilly, J.E., Maritorena, S., Mitchell, B.G., Siegel, D.A., Carder, K.L.,

Garver, S.A., Kahru, M., and McClain, C., 1998, Ocean color

chlorophyll algorithms for SeaWiFS: Journal of Geophysical Research­

Oceans, v. 103, p. 24937-24953.

O'Reilly, J.E., Maritorena, S., Siegel, D.A., O'Brien, M.C., Toole, D., Chavez,

F.P., Strutton, P., Cota, G.F., Hooker, S.B., McClain, C.R., Carder,

K.L., Muller-Karger, F., Harding, L., Magnuson, A., Phinney, D.,

Moore, G.F., Aiken, J., Arrigo, K.R., Letelier, R., and Culver, M., 2000,

Sea WiFS Postlaunch Calibration and Validation Analyses, Part 3.

Ostir, K., Veljanovski, T., Podobnikar, T., and Stancic, Z., 2003, Application

of satellite remote sensing in natural hazard management: the Mount

Mangart landslide case study: International Journal of Remote

Sensing, v. 24, p. 3983-4002.

Paerl, H.W., 2008, Nutrient and other environmental controls of harmful

cyanobacterial blooms along the freshwater-marine continuum, in

Hudnell, H.K., ed., Cyanobacterial Harmful Algal Blooms: State of the

Science and Research Needs: Advances in Experimental Medicine and

Biology: New York, NY, Springer, p. 950.

104

Parsons, T.T.a.J.D.H.S., 1963, Discussion of spectrophotometric

determination of marine-plant pigments, with revised equations for

ascertaining chlorophylls and carotenoids.: J. Mar. Res, v. 21, p. 155-

163.

Patt, F.S., Barnes, R.A., Eplee, R.E.J., Franz, B.A., Robinson, W.D., Feldman,

G., Bailey, S.W., Gales, P.J., Werdell, P.J., Wang, M., Frouin, R.,

Stumpf, R.P., Arnone, R., Gould, R.W., Martinolich, P.M.,

Ransibrahmanakul, V., O'Reilly, J.E., and Yoder, J., 2003, Algorithm

Updates for the Fourth SeaWiFS Data Reprocessing, NASA Tech.

Memo. 2003--206892, Volume 22, NASA Goddard Space Flight Center,

Greenbelt, Maryland, p. 7 4.

Pegau, W.S., Zaneveld, J.R.V., Barnard, A.H., Maske, H., Alvarez-Borrego, S.,

Lara-Lara, R., and Cervantes-Duarte, R., 1999, Inherent optical

properties in the Gulf of California: Ciencias Marinas, v. 25, p. 469-

485.

Pope, R.M., and Fry, E.S., 1997, Absorption spectrum (380-700 nm) of pure

water .2. Integrating cavity measurements: Applied Optics, v. 36, p.

8710-8723.

Pultz, T.J., and Scofield, R.A., 2002, Applications of remotely sensed data in

flood prediction and monitoring: report of the CEOS Disaster

Management Support Group flood team: 2002 IEEE International

Geoscience and Remote Sensing Symposium. 24th Canadian

105

Symposium on Remote Sensing. Proceedings (Cat.

No.02CH37380) I 2002 IEEE International Geoscience and Remote

Sensing Symposium. 24th Canadian Symposium on Remote Sensing.

Proceedings (Cat. No.02CH37380), p. 768-70 vol.2 I 6 vol.xcvi+3694.

Refice, A., Bovenga, F., and Nutricato, R., 2003, Stepwise Approach to INSAR

Processing of Multi temporal Datasets, Fringe 20003, Volume SP-550:

Frascati, Italy, ESA.

Reynolds, C.S., Oliver, R.L., and Walsby, A.E., 1987, Cyanobacterial

dominance: the role of buoyancy regulation in dynamic lake

environments: New Zealand J ournalof Marine and Freshwater

Research, v. 21, p. 379-390.

RIGW, 1992, Nile Delta, Hydrogeological Map of Egypt: El Kanater el

Khariya, Research Institute for Groundwater (RIGW), Ministry of

Public Works and Water Resources (MPWWR).

Rinta-Kanto, J.M., Oullette, A.J.A., Twiss, M.R., Boyer, G.L., Bridgeman,

T.B., and Wilhelm, S.W., 2005, Quantification of Toxic Microcystis spp.

during the 2003 and 2004 blooms in Western Lake Erie:

Environmental Science and Technology.

Ruddick, K.G., Ovidio, F., and Rijkeboer, M., 2000, Atmospheric correction of

Sea WiFS imagery for turbid coastal and inland waters: Applied Optics,

V. 39, p. 897-912.

106

Said, R., 1990, The Geology of Egypt: Rotterdam, Netherlands, A. A.

Balkema, 734 p.

-, 1993, The River Nile: Geology, Hydrology and Utilization, Pergamon

Press.

Sandwell, D.T., and Price, E.J., 1998, Phase gradient approach to stacking

interferograms: Journal of Geophysical Research-Solid Earth, v. 103, p.

30183-30204.

Sandwell, D.T., Sichoix, L., and Smith, B., 2002, The 1999 Hector Mine

earthquake, southern California: Vector near-field displacements from

ERS InSAR: Bulletin of the Seismological Society of America, v. 92, p.

1341-1354.

Sathyendranath, S., Lazzara, L., and Prieur, L., 1987, Variations in the

Spectral Values of Specific Absorption of Phytoplankton: Limnology

and Oceanography, v. 32, p. 403-415.

Scharroo, R., and Visser, P., 1998, Precise orbit determination and gravity

field improvement for the ERS satellites: Journal of Geophysical

Research-Oceans, v. 103, p. 8113-8127.

Schneider, B., Bopp, L., Gehlen, M., Segschneider, J., Frolicher, T.L., Cadule,

P., Friedlingstein, P., Doney, S.C., Behrenfeld, M.J., and Joos, F., 2008,

Climate-induced interannual variability of marine primary and export

production in three global coupled climate carbon cycle models:

Biogeosciences, v. 5, p. 597-614.

107

Sestini, G., 1989, The Implications of Climate Changes for the Nile Delta:

Nairobi, Kenya, UNEP/OCA.

Shear, H., Fuller, K., and Wittig, J., 1995, The Great Lakes An

Environmental Atlas and Resource Book/Les Grands Lacs, atlas

ecologique et manuel des ressources: Toronto and Chicago, United

States Environmental Protection Agency and Government of Canada.

Siegelman, H.W., and Kycia, J.H., 1978, Algal biliproteins, in Hellebrust, J.,

and Craigie, J.S., eds., Handbook of Phycological Methods:

Physiological and Biochemical Methods: Cambridge, UK, Cambride

University Press, p. 71-80.

Simis, S.G.H., Peters, S.W.M., and Gons, H.J., 2005, Remote sensing of the

cyanobacterial pigment phycocyanin in turbid inland water: Limnology

and Oceanography, v. 50, p. 237-245.

Simis, S.G.H., Ruiz-Verdu, A., Dominguez-Gomez, J.A., Pena-Martinez, R.,

Peters, S.W.M., and Gons, H.J., 2007, Influence of phytoplankton

pigment composition on remote sensing of cyanobacterial biomass:

Remote Sensing of Environment, v. 106, p. 414-427.

Stanley, D.J., 1988, Subsidence in the Northeastern Nile Delta - Rapid Rates,

Possible Causes, and Consequences: Science, v. 240, p. 497-500.

Stanley, D.J., and Hait, A.K., 2000, Deltas, radiocarbon dating, and

measurements of sediment storage and subsidence: Geology, v. 28, p.

295-298.

108

Stanley, D.J., McRea, J.E., Jr., and Waldron, J.C., 1996, Nile delta drill core

and sample database for 1985-1994: Mediterranean basin (MEDIBA)

program, Smithsonian Contributions to the Marine Sciences, v. 37, p.

1-428.

Stanley, D.J., and Warne, A.G., 1993a, Nile Delta: Recent Geological

Evolution and Human Impact: Science, v. 260, p. 628-634.

-, 1993b, Sea-Level and Initiation of Predynastic Culture in the Nile Delta:

Nature, v. 363, p. 435-438.

-, 1998, Nile delta in its destruction phase: Journal of Coastal Research, v.

14, p. 794-825.

Stanley, J.D., 2001, Dating modern deltas: Progress, problems, and

prognostics: Annual Review of Earth and Planetary Sciences, v. 29, p.

257-294.

Stanley, J.D., Warne, A.G., and Schnepp, G., 2004, Geoarchaeological

interpretation of the canopic, largest of the relict Nile Delta

distributaries, Egypt: Journal of Coastal Research, v. 20, p. 920-930.

Stolarski, R.S., Douglass, A.R., Gupta, M., Newman, P.A., Pawson, S.,

Schoeberl, M.R., and Nielsen, J.E., 2006, An ozone increase in the

Antarctic summer stratosphere: a dynamical response to the ozone

hole: Geophysical Research Letters I Geophysical Research Letters, p.

1-4.

109

Stumpf, R.P., Arnone, R., Gould, R.W., Martinolich, P.M., and

Ransibrahmanakul, V., 2003, A partially coupled ocean-atmosphere

model for retrieval of water-leaving radiance from seawifs in coastal

waters, NASA Tech. Memo. 206892, National Aeronautics and Space

Administration, Goddard Space Flight Center, Greenbelt, MD.

Sturm, B., and Zibordi, G., 2002, SeaWiFS atmospheric correction by an

approximate model and vicarious calibration: International Journal of

Remote Sensing, v. 23, p. 489-501.

Tornqvist, T.E., Wallace, D.J., Storms, J.E.A., Wallinga, J., van Dam, R.L.,

Blaauw, M., Derksen, M.S., Klerks, C.J.W., Meijneken, C., and

Snijders, E.M.A., 2008, Mississippi Delta subsidence primarily caused

by compaction of Holocene strata: Nature Geosciences, v. 1, p. 173-176.

Ucuncuoglu, E., Arli, 0., and Eronat, A.H., 2006, Evaluating the impact of

coastal land uses on water-clarity conditions from Landsat TMIETM +

imagery: Candarli Bay, Aegean Sea: International Journal of Remote

Sensing, v. 27, p. 3627-3643.

Vanderploeg, H.A., Liebig, J.R., Carmichael, W.W., Agy, M.A., Johengen,

T.H., Fahnenstiel, GL., and Nalepa, T.F., 2001, Zebra mussel

(Dreissena polymorpha) selective filtration promoted toxic Microcystis

blooms in Saginaw Bay (Lake Huron) and Lake Erie: Canadian

Journal of Fisheries and Aquatic Sciences, v. 58, p. 1208-1221.

110

Vincent, R.K., Qin, X.M., McKay, R.M.L., Miner, J., Czajkowski, K., Savino,

J., and Bridgeman, T., 2004, Phycocyanin detection from LANDSAT

TM data for mapping cyanobacterial blooms in Lake Erie: Remote

Sensing of Environment, v. 89, p. 381-392.

Warne, A.G., and Stanley, D.J., 1993, Archaeology to Refine Holocene

Subsidence Rates Along the Nile Delta Margin, Egypt: Geology, v. 21,

p. 715-718.

Watson, K.M., Bock, Y., and Sandwell, D.T., 2002, Satellite interferometric

observations of displacements associated with seasonal groundwater in

the Los Angeles basin: Journal of Geophysical Research-Solid Earth, v.

107.

Welschmeyer, N.A., 1994, Fluorometric analysis of chlorophyll a in the

presence of chlorophyll b and pheopigments: Limnology and

Oceanography, v. 39, p. 1985-1992.

Yang, H., Adler, R., and Huffman, G., 2006, Evaluation of the potential of

NASA multi-satellite precipitation analysis in global landslide hazard

assessment: Geophysical Research Letters I Geophysical Research

Letters, p. 1-5.

Yang, M.D., Su, T.C., Hsu, C.H., Chang, K.C., and Wu, A.M., 2007, Mapping

of the 26 December 2004 tsunami disaster by using FORMOSAT-2

images: International Journal of Remote Sensing, v. 28, p. 3071-3091.

111

APPENDIX

SELECTED INTERMEDIATE INSAR PRODUCTS

112

Figure Al Three Pass interferogram generated from scenes acquired 8-24-1995

and 8-23-1995 prior to persistent scatterer selection. Good coherence is seen due

to small spatial and temporal baselines

113

Figure A.2 Three Pass interferogram generated from scenes acquired 8-24-1995

and 1-11-1996. Coherence is maintained in areas around cities and buildings.

These areas are good candidates for persistent scatterers.

114

t •

Stanley et al 1993

Figure A.3 Sample subsidence rate calculation for location in Damietta City

before atmospheric corrections are applied. Error bars show 1 standard deviation

from average of all nearby points within 1 km 2. Y axis shows vertical position

real relative to entire scene average using the 5-15-1997 scene as the reference.

115