266
BOSTON UNIVERSITY GRADUATE SCHOOL OF ARTS AND SCIENCES Dissertation EARTH SYSTEM DATA RECORDS OF VEGETATION LEAF AREA INDEX FROM MULTIPLE SATELLITE-BORNE SENSORS by SANGRAM GANGULY B.Sc., Indian Institute of Technology, Kharagpur, 2004 M.Sc., Indian Institute of Technology, Kharagpur, 2004

Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

BOSTON UNIVERSITY

GRADUATE SCHOOL OF ARTS AND SCIENCES

Dissertation

EARTH SYSTEM DATA RECORDS OF VEGETATION LEAF AREA INDEX

FROM MULTIPLE SATELLITE-BORNE SENSORS

by

SANGRAM GANGULY

B.Sc., Indian Institute of Technology, Kharagpur, 2004M.Sc., Indian Institute of Technology, Kharagpur, 2004

Submitted in partial fulfillment of the

requirements for the degree of

Doctor of Philosophy

2008

Page 2: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Approved by

First Reader ________________________________________________Ranga B. Myneni, Ph.D.Professor of Geography and Environment

Second Reader ________________________________________________Yuri Knyazikhin, Ph.D.Research Professor of Geography and Environment

Third Reader ________________________________________________Alan H. Strahler, Ph.D.Professor of Geography and Environment

Fourth Reader ________________________________________________Mark A. Friedl, Ph.D.Professor of Geography and Environment

Fifth Reader ________________________________________________Compton J. Tucker, Ph.D.Research Scientist at NASA/Goddard Space Flight Center

ii

ii

Page 3: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Acknowledgements

I express a deep sense of gratitude to my advisors, Prof. Ranga Myneni and Prof.

Yuri Knyazikhin for providing me the opportunity to pursue my Ph.D. in the Department

of Geography and Environment, Boston University and for their valuable guidance,

encouragement and helpful counsel throughout my doctoral research work. In particular,

they exposed me to the relevant research topic, through rigorous scientific discussions

and timely suggestions. I take this opportunity to place on record my indebtedness to

them for their genuine interest in shaping my research career. It is indeed difficult to

phrase their benefaction in a few words.

I express my earnest acknowledgement to Prof. Robert Kaufmann, Prof. Sucharita

Gopal, Prof. Nathan Phillips, Prof. Mark Friedl and Prof. Bruce Anderson for sharing

with me their expertise on concepts of statistical techniques, GIS, ecological physiology

and physical climatology in the wider perspective of vegetation remote sensing and

climate interactions. I also express my sincere gratitude to Prof. Alan H. Strahler and

Prof. Curtis Woodcock for the captivating classes on advanced remote sensing

techniques. The teachings instilled in me the rationale of my scientific pursuit.

I sincerely appreciate the contributions of Dr. Ramakrishna Nemani and Dr.

Cristina Milesi (NASA Ames Research Center) for the research on climate and land use

change in the semi-arid lands; and to Dr. Nikolay Shabanov and Dr. Dong Huang for

helping me master the theoretical as well as the modeling domain of stochastic radiative

transfer theory. Thanks to Dr. Wenze Yang and Dr. Weile Wang for the numerous

intellectual and technical conversations during their stay here at Boston University.

iii

iii

Page 4: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

I extend my most sincere appreciation to my friends Arindam Samanta and Mitchell

Schull for the opportunity of working with them and their invaluable assistance in the

completion of the work. I offer my laudation to them for all the inspiring ideas and

relentless discussions on science and philosophy.

I also wish to express my special thanks to Aparna Kar, Leah Mehl, Sung Bae Jeon,

Qingling Zhang, Manish Verma, Maria, Changho So, Miguel Roman and all my other

friends for their very friendly returns, association, help and joint programs due to which

my stay with them turned out to be so rewarding and enjoyable.

Finally, I also owe my genuine deference to Mike Holmes, Laura Guild, Erin

Wnorowski, Christopher DeVits, Anastasia Stavropoulos, Linda Gregory and Sayaka

Yamaki for their succor in letting me avail of all the departmental facilities and also for

providing me every kind of administrative support at all times. A greater appreciation

goes to all the faculty and staff of the Department of Geography and Environment for

providing an excellent research environment as well as a cordial ambiance for cultivating

my individual growth. The experience is truly unforgettable.

Last but not least, no work could have been completed without the sustained and

undeterred motivation from my parents and their support and strong interest in everything

that I have had undertaken from my childhood. I also thank my brother Dr. Samrat

Ganguly for his constant support, encouragement and timely suggestions at all points of

my career.

iv

iv

Page 5: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

EARTH SYSTEM DATA RECORDS OF VEGETATION LEAF AREA INDEX

FROM MULTIPLE SATELLITE-BORNE SENSORS

(Order No. )

SANGRAM GANGULY

Boston University Graduate School of Arts and Sciences, 2008

Major Professor: Ranga B. Myneni, Professor of Geography and Environment

ABSTRACT

Monitoring of vegetation structure and functioning is critical to modeling terrestrial

material and energy cycles, ecosystem productivity and land use/land cover dynamics

within the general context of climate change. Satellite remote sensing is ideally suited for

vegetation monitoring as it provides multi-decadal observations at a range of spatio-

temporal scales. Consequently, there is now a pressing need to develop methodologies for

generating consistent Earth System Data Records (ESDRs) from multiple satellite

sensors.

Leaf Area Index (LAI) and Fraction of Photosynthetically Active Radiation (FPAR)

are sensor-independent measurable characteristics of vegetation. Multi-decadal global

data sets of these variables generated with a physically based algorithm and of known

accuracy are currently not available, although several short term research quality data sets

exist. Therefore, the objective of this research was to formulate and demonstrate the

performance of a synergistic approach for the retrieval of LAI/FPAR fields from

measurements by multiple sensors.

v

v

Page 6: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

An algorithm to generate consistent LAI values from the Moderate Resolution

Imaging Spectroradiometer (MODIS) and the Advanced Very High Resolution

Radiometers (AVHRR) was developed. The approach is based on the radiative transfer

theory of spectral invariants. The scale-dependent single scattering albedo and input data

uncertainties govern the constrains to consistent retrievals of LAI from data of multiple

sensors. A global monthly AVHRR LAI data set was generated from the corresponding

Normalized Difference Vegetation Index (NDVI) fields for the period July 1981 to

December 2006 using this algorithm.

An evaluation of generated data set through direct comparison to ground data and

inter-comparison with other LAI and surrogate data indicates good agreement both in

terms of absolute values and temporal variations. The utility of the derived data set is

demonstrated with a case study on characterizing climate and land use impacts on

vegetation in the semi-arid tropics. It was found that large portions of the densely

populated, tropical dry lands of the eastern hemisphere have experienced marked positive

trends in vegetation greenness over the period 1981-2006.

The results presented in this dissertation imbue confidence in the utility of this

seamless, consistent satellite data product for large scale terrestrial-biosphere modeling

and monitoring of global vegetation dynamics.

vi

vi

Page 7: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Table of Contents

Acknowledgements...........................................................................................................iii

ABSTRACT........................................................................................................................v

Table of Contents.............................................................................................................vii

List of Tables......................................................................................................................x

List of Figures...................................................................................................................xi

List of Abbreviations......................................................................................................xiv

1. Introduction....................................................................................................................1

1.1 Background................................................................................................................1

1.1.1 Climate-vegetation interaction............................................................................1

1.1.2 Long term vegetation monitoring with satellite data..........................................8

1.2 LAI/FPAR algorithm and products...........................................................................9

1.2.1 Definition of LAI and FPAR..............................................................................9

1.2.2 LAI/FPAR algorithms......................................................................................11

1.2.3 LAI and FPAR products from different sensors...............................................13

1.3 Objectives and structure of this dissertation............................................................14

1.3.1 Theoretical formulation for generating LAI and FPAR...................................16

1.3.2 Global LAI from AVHRR NDVI: production, evaluation and validation.......16

1.3.3 Vegetation changes versus climate/land use change in the semi-arid tropics. .17

2. Generating Leaf Area Index Earth System Data Record from multiple sensors:

Theory...............................................................................................................................20

2.1 Introduction..............................................................................................................20

2.2 Criteria for ensuring consistency.............................................................................24

2.3 Parameterization for canopy spectral reflectance....................................................26

2.3.1 Canopy spectral invariants................................................................................26

2.3.2 Canopy-ground interactions..............................................................................29

2.3.3 Generation of structural parameters..................................................................30

2.4 Algorithm adjustment for the sensors spatial resolution.........................................32

vii

vii

Page 8: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

2.5 Algorithm adjustment for the sensors spectral bandwidth......................................37

2.6 Data information content and observation uncertainty............................................40

2.7 Conclusions..............................................................................................................42

3. Generating Leaf Area Index Earth System Data Record from multiple sensors:

Implementation, analysis and validation.......................................................................54

3.1 Introduction..............................................................................................................54

3.2 Production of the LAI dataset..................................................................................56

3.2.1 Input satellite data and land cover classification map......................................56

3.2.2 Achieving consistency with the Terra MODIS LAI products..........................58

3.2.3 Global LAI production.....................................................................................60

3.2.4 LAI data dissemination.....................................................................................61

3.3 Comparison with other satellite LAI products.........................................................61

3.3.1 Assessment of AVHRR LAI for the tuning year..............................................61

3.3.2 Comparison with MODIS C5 LAI data............................................................63

3.3.3 Comparison with CYCLOPES LAI product....................................................64

3.4 Comparison with climate variables.........................................................................66

3.4.1 LAI variation with surface temperature in the northern latitudes.....................66

3.4.2 LAI variation with precipitation in the semi-arid tropics.................................68

3.4.3 Canonical correlation analysis..........................................................................70

3.5 Validation with ground data....................................................................................72

3.5.1 Validation with field/plot-level observations...................................................73

3.5.2 Validation with fine resolution LAI maps........................................................75

3.6 Conclusions..............................................................................................................76

4. Climate and land use impacts on vegetation in the semi-arid tropics: A case study

with the LAI data set.......................................................................................................94

4.1 Introduction..............................................................................................................94

4.2 Data and methods....................................................................................................96

4.2.1 Normalized Difference Vegetation Index.........................................................96

4.2.2 Leaf Area Index................................................................................................96

viii

viii

Page 9: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

4.2.3 Precipitation......................................................................................................97

4.2.4 Land cover........................................................................................................98

4.2.4 Ancillary data....................................................................................................98

4.2.5 Methods............................................................................................................99

4.3 Results and discussion...........................................................................................100

4.3.1 Climate-driven increases in LAI.....................................................................100

4.3.2 Land use change-driven increases in LAI......................................................102

4.4 Conclusions............................................................................................................105

5. Concluding remarks..................................................................................................118

Appendices......................................................................................................................128

Appendix A: Analytical expression of the “S” problem.............................................128

Appendix B: Ancillary tables......................................................................................130

List of Journal Abbreviations.......................................................................................133

Bibliography...................................................................................................................135

Curriculum Vitae.............................................................................................................154

ix

ix

Page 10: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

List of Tables

Table 3.1 Global scale analysis of LAI differences across different biomes for different months and for annual maximum LAI....………………………………………………...90

Table 3.2 Relationship between LAI, NDVI and temperature anomalies in Eurasia (EA) and North America (NA) during the time period 1982 to 2006 and 1982 to 1999………91

Table 3.3 Relationship between annual maximum LAI, NDVI and three wettest month precipitation anomalies in the semi-arid tropics from 1981 to 2006…………………….92

Table 3.4 Validation of AVHRR LAI with fine resolution LAI maps over different sites. Mean () and dispersion () are measured with respect to valid retrievals over both field and satellite measured values…………………………………………………………….93

Table 4.1 Change in LAI, precipitation, food production, fertilizer use, irrigated land fraction, population, and per capita gross domestic product for 19 countries from 1981 to 2006……………………………………………………………………………………..117

Table B1 BELMANIP sites used for CYCLOPES LAI and AVHRR LAI inter-comparison……………………………………………………………………………...130

Table B2 Summary of MODIS LAI field campaigns used for validation with AVHRR LAI……………………………………………………………………………………...131

List of Figures

x

x

Page 11: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 1.1 Flowchart of the LAI retrieval process............................................................19

Figure 2.1 Schematic plot of the photon-canopy interactions...........................................44

Figure 2.2 Canopy interceptances i0 and i0,S as a function of LAI.....................................45

Figure 2.3 Recollision probabilities calculated by fitting equations for the black soil (BRFBS, aBS and tBS) and “S” (aS, rS, and tS) problems to their simulated values.......46

Figure 2.4 Values of the energy conservation relationships rBS+tBS+aBS as a function of single scattering albedo and LAI...............................................................................47

Figure 2.5 The spectral invariant approximation to the BRF superimposed on the MODIS LUTs entries for the BRF..........................................................................................48

Figure 2.6 Relative spectral response function in the red and NIR spectral intervals for the NOAA AVHRR-16 and MODIS sensors............................................................49

Figure 2.7 The ratio (p) for red and NIR spectral bands for AVHRR and MODIS........50

Figure 2.8 Reflectance of vegetated surface in the red-NIR spectral plane......................51

Figure 2.9 Distribution of acceptable LAI values corresponding to the full range of possible values of the radius (squares) and to a specific value of the radius (triangles)...................................................................................................................52

Figure 2.10 Correlation between LAI values retrieved using the MODIS (horizontal axis) and AVHHR (vertical axis) modes of the algorithm for dark (sur,RED=sur,RED=0.05), medium (sur,RED=sur,RED=0.16) and bright (sur,RED=sur,RED=0.26) backgrounds........53

Figure 3.1 The difference (RMSE) between MODIS and AVHRR LAI values (vertical axis on the left side and solid lines) and the Retrieval Index (vertical axis on the right side and dashed lines) as a function of the relative uncertainty =RED/RED and single scattering albedo at red spectral band.............................................................79

Figure 3.2 Flowchart showing global LAI production......................................................80

Figure 3.3 Average annual LAI difference (AVHRR minus MODIS) for the year 2001. 81

Figure 3.4 Panel (a) shows comparison between MODIS and AVHRR LAI for the year 2001 (blue color) and 2002 (red color) for different vegetation classes....................82

Figure 3.5 Panels (a) to (d) shows comparison between CYCLOPES and AVHRR LAI values for years 2001 to 2003 and different vegetation classes over the BELMANIP

xi

xi

Page 12: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

site database...............................................................................................................83

Figure 3.6 Trends in AVHRR LAI for the growing season, April to October, for the region 40o N to 70o N, for the periods 1982 to 1999 (panel (a)) and 1982 to 2006 (panel (b))..................................................................................................................84

Figure 3.7 Standardized April to October anomalies of AVHRR LAI (green), AVHRR NDVI (blue), and GISS Temperature (red dashed line) for Eurasian and North American needleleaf forests (panels (c) and (d)) and tundra (panels (a) and (b)) from 1982 to 2006..............................................................................................................85

Figure 3.8 Standardized anomalies of annual peak AVHRR LAI (green line), annual peak AVHRR NDVI (blue line) and annual peak (three wettest month CRU+TRMM) precipitation (red dashed line) for the semi-arid regions (panels (a)-(d)) from 1981 to 2006...........................................................................................................................86

Figure 3.9 Correlation between standardized time series of the first canonical factor (CF-1, panels (a) and (c)) and second canonical factor (CF-2, panels (b) and (d)) with NINO3 and AO indices in the northern and tropical/subtropical regions.................87

Figure 3.10 Panel (a) shows a comparison of AVHRR LAI with field measurements for the six major vegetation classes. Altogether 44 field data values were used (Table B2 of Appendix B).....................................................................................................88

Figure 3.11 Histograms from fine resolution LAI maps (blue color) and AVHRR LAI (red color) over different sites...................................................................................89

Figure 4.1 Panel (a) shows colormap of peak annual NDVI climatology. Panel (b) shows the spatial frequency distribution (number of pixels) of peak annual NDVI climatology values in the range 0.12-0.55...............................................................107

Figure 4.2 Percentage distribution of IGBP land cover classes (Panel (a)) and frequency distribution of bare (red), herbaceous (blue) and tree (black) cover from MODIS VCF map - expressed as percentage of total number of pixels (Panel (b)) for peak annual NDVI climatology range of 0.12-0.55.........................................................108

Figure 4.3 Colormap of percentage change in mean peak annual LAI between decade 1 (1981-1990) and decade 2 (1995-2006)...................................................................109

Figure 4.4 Colormap of percentage change in mean peak annual precipitation (mm/yr) between decade 1 (1981-1990) and decade 2 (1995-2006).....................................110

Figure 4.5 Plots of peak annual LAI anomaly and peak annual rainfall anomaly for 8 representative countries in the study region – Australia, Botswana, India, Mali,

xii

xii

Page 13: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Niger, Saudi Arabia, South Africa and Turkey.......................................................111

Figure 4.6 Monthly average temperature and monthly precipitation trend for India in the 1980’s and 1990’s....................................................................................................112

Figure 4.7 Long term mean monthly LAI (red line) and percent trend in monthly LAI (green line) for India over the time period 1981-2006............................................113

Figure 4.8 Panel (a) shows cross plot between area with increased LAI (pixels with statistically significant (pvalue<0.05) positive LAI trend during the period 1981-2006) within each major Indian state and change in net irrigated area over the same period. Panel (b) is a similar plot as panel (a) but with all pixels with positive LAI trend during the same period...................................................................................114

Figure 4.9 Percentage change in mean peak annual LAI as in Fig. 4.3 for the semi-arid districts of Mandsaur (State: Madhya Pradesh), Kota (State: Rajasthan), Jhalawar (State: Rajasthan) and Ujjain (State: Madhya Pradesh) in India.............................115

Figure 4.10 The percentage of cropped area that is irrigated (blue bar) and percentage of irrigated land utilizing groundwater (green bar) for each of the major Indian states (Panel (a)). Panel (b) tabulates the fertilizer consumption (kg/Ha) and fraction irrigated sown area (in %) for different states (year 2002-2003) and shows the corresponding regression relation (orange squares represents states).....................116

List of Abbreviations

AO Arctic Oscillation

xiii

xiii

Page 14: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

ASTER Advanced Spaceborne Thermal Emission and Reflection Radiometer

AVHRR Advanced Very High Resolution Radiometer

AVIM Atmospheric-Vegetation Interaction Model

BGC Biogeochemical Cycles

BRF Bi-Directional Reflectance Factor

CCA Canonical Correlation Analysis

CCM Community Climate Model

CEOS Committee on Earth Observation Satellites

CRU Climate Research Unit

DAAC Distributed Active Archive Center

DGVM Dynamic Global Vegetation Model

DN Digital Number

ENVISAT Environmental Satellite

ENSO El Niño-Southern Oscillation

ESDR Earth System Data Record

ETM Enhanced Thematic Mapper

FPAR Fraction of Photosynthetically Active Radiation Absorbed by Vegetation

GCOS Global Climate Observing System

GIMMS Global Inventory Monitoring and Modeling Studies

GISS Goddard Institute for Space Studies

xiv

xiv

Page 15: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

IBIS Integrated Biosphere Simulated Model

IGBP International Geosphere-Biosphere Programme

IPCC Intergovernmental Panel on Climate Change

LAI Leaf Area Index

LPJ Lund-Potsdam-Jenna

LPV Land Product Validation

LSM Land Surface Model

LUT Look-Up Table

MAP Mean Annual Precipitation

MERIS Medium Resolution Imaging Spectrometers

MISR Multiangle Imaging Spectroradiometer

MODIS Moderate Resolution Imaging Spectroradiometer

NASA National Aeronautics and Space Administration

NCAR National Center for Atmospheric Research

NDVI Normalized Difference Vegetation Index

NIR Near-Infrared

NOAA National Oceanic and Atmospheric Administration

NPOESS National Polar-orbiting Operational Environmental Satellite System

NPP NPOESS Preparatory Project

ORNL Oak Ridge National Laboratory

PAL Pathfinder AVHRR Land

xv

xv

Page 16: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

PC Principal Component

POLDER Polarization and Directionality of the Earth’s Reflectances

PSF Point Spread Function

RI Retrieval Index

RMSE Root Mean Square Error

RT Radiative Transfer

SeaWiFS Sea-viewing Wide Field-of-view Sensor

SiB Simple Biospheric Model

SM Simple Ratio

SPOT Systeme Pour l’Observation de la Terre

SRF Spectral Response Function

SSR Spectral Surface Reflectance

TEM Terrestrial Ecosystem Model

TRAC Tracing Radiation and Architecture of Canopies

TRMM Tropical Rainfall Measuring Mission

VCF Vegetation Continuous Fields

VIIRS Visible/Infrared Imager/Radiometer Suite

Chapter 1

xvi

xvi

Page 17: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

1. Introduction

1.1 Background

Vegetation covers almost 75% of the Earth’s land surface. Its character, structure

and functioning is critical to modeling material and energy cycles in our climate system.

Therefore, long-term monitoring of vegetation is a topical theme in light of present

concerns about changing climate. Satellite remote sensing provides the ideal data for

monitoring changes in land surface characteristics at a range of scales with sufficient

spatial and temporal resolution. The advances in remote sensing, both in theory and

instrumentation, have paved the way for better understanding the partitioning of radiative

energy between the Earth’s surface and atmosphere (Tucker, 1986; Diner et al., 1999;

Justice et al., 1998). Consequently, studies on the retrieval of biophysical variables that

act as a proxy to the amount of vegetation on the land surface have gained momentum in

recent decades.

1.1.1 Climate-vegetation interaction

Numerous terrestrial-biosphere models use land surface properties such as land

cover type, leaf area index (LAI), fraction of photosynthetically active radiation (0.4-

0.7mm) absorbed by vegetation (FPAR), roughness length, albedo, etc. as key inputs for

simulation of land surface processes. Such processes are important components of the

climate system and their impact on climate variability has been documented in a number

of sensitivity studies (Mintz 1984; Dickinson 1986; Sellers et al., 1986; Bonan, 1998).

xvii

xvii

Page 18: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

These and other studies have shown that successful modeling of mass and energy

exchange between the terrestrial biosphere and the atmosphere requires an accurate

description of the ecological state coupled with appropriate biogeochemical controls.

Vegetation and climate interact at time scales from seconds to hundreds of years (Sellers

et al., 1995; Pielke et al., 1998). The impact of vegetation on climate as well as its

response to climate perturbations at short to longer time scales is well documented in

literature.

1.1.1.1 Vegetation impact on climate

The impact/feedback of vegetation on climate through exchanges of energy, water,

momentum, CO2, and other important atmospheric gases is progressively becoming a

vital component in the realm of climate-vegetation interactions. The IPCC (IPCC 2007)

has identified changes in ecosystem and vegetated land cover as a potentially important

climate feedback. Changes in species composition and ecosystem structure alter albedo,

surface roughness, stomatal conductance, rooting depth, and nutrient availability, which

in turn alter land surface energy fluxes and hydrological and biogeochemical cycles

(Panel on Climate Change Feedbacks, 2003).

a) Stomata feedback: Stomatal conductance and the amount of leaf area play an

important role in the partitioning of net radiation into sensible and latent heat fluxes,

constituting a major component in vegetation feedback to climate. Stomata provide the

pathway to facilitate the exchange of mass and energy between the atmosphere and the

cellular tissues inside the leaf. Stomatal conductance usually decreases with elevated

levels of CO2, while photosynthesis increases. Several climate model simulations

xviii

xviii

Page 19: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

incorporating elevated CO2 levels and decreased stomatal conductance show a decreased

latent heat flux, increased sensible heat and surface warming over large vegetated areas

(Sellers et al., 1996). Considering the direct response of stomata to changes in

environmental CO2 levels, it is difficult to quantify this feedback at a global scale and

assess the magnitude of it relative to other climate feedbacks. This is mostly due to the

uncertainties related to the physiological processes operating at the scale of a leaf as

compared to those operating at the canopy or landscape level. In addition, most studies

incorporate short term response of plants to CO2 as opposed to long term acclimation to

higher CO2. This is also directly related to the amount of leaf area, as changing the leaf

area will most certainly alter the physiological response of stomata (canopy conductance)

and the allocation of carbon, further altering the net exchange of heat and moisture

between the atmosphere and plants.

b) Leaf area feedback: The leaf area feedback, alongside the stomatal response to

factors like increased CO2 and moisture content, also plays a significant role in altering

the land surface climate, including temperature and transpiration by changing the albedo

and sensible and latent heat fluxes (Schwartz et al., 1999; Fitzjarrald et al., 2001; Cleland

et al., 2007). This feedback has been predominantly characterized by the seasonal

dynamics or phenology of leaves (onset and senescence of leaves). Observations over the

broadleaf deciduous trees of eastern United States show that the springtime air

temperatures are distinctly different (for a period of less than a few weeks) after the onset

of leaves, and is directly related to increased rates of transpiration upon leaf emergence

that cools and moistens the air (Schwartz and Karl, 1990; Schwartz, 1992, 1996). A

xix

xix

Page 20: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

similar pattern was also observed in the deciduous trees of west central Canada (Hogg et

al., 2000). Leaf area and phenology are critical in regulating surface climate and are

incorporated in various climate models. Buermann et al., (2001) showed that higher leaf

over vegetated regions in summer increased the evaporative demand (provided there is

adequate soil water), thus resulting in the cooling of surface temperatures and

precipitation increases. Incorporating the seasonality of leaf area in the NCAR CCM3

climate model, Dickinson et al., (1998) demonstrated that the midsummer precipitation is

significantly reduced and surface temperatures are increased over the extratropical

continents in the Northern Hemisphere due to lower leaf area.

c) Albedo feedback: The albedo over vegetated land surfaces can vary from very

low values (10-15% over humid tropical forests) to somewhat larger values (15-20% over

herbaceous vegetation) (Hartmann, 1994). The vegetation albedo is a function of plant

structural and optical properties and the leaf area index (Dickinson, 1983). As the amount

of leaf area increases, light absorption increases inside a canopy and the reflection from

background soil decreases. This results in an overall decrease in surface albedo. A similar

situation is observed in the case of snow in tree-covered landscapes. Land surface with

fully covered snow has high values of surface albedo (~0.8). The high albedo is masked

in the presence of trees and can reach significantly lower values (~0.2-0.4) depending on

vegetation cover type. Several climate model studies have found that the presence of

boreal forests warm climate as compared to the tundra, thus contributing to a significant

positive vegetation feedback (Bonan et al., 1992; Douville and Royer, 1996). This

phenomena is further manifested in the fact that both the forest and tundra ecosystems

xx

xx

Page 21: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

significantly differ in net radiation partitioning (sensible and latent heat fluxes). The

classic study by Charney, (1975) showed that in the sub-tropical desert-margin regions,

heat loss due to high albedos from decreased vegetation leads to a horizontal temperature

gradient and induces an atmospheric circulation which maintains the sinking motion of

dry air masses. As a result, there is a further reduction in precipitation and henceforth

creates a positive feedback that possibly explains recurrence of droughts and

desertification.

d) Surface roughness feedback: The roughness of a canopy surface determines the

degree of coupling between plants and the atmosphere, and in principal influences the

partitioning of sensible and latent heat fluxes between the land surface and the overlying

air (Lambers et al., 2000). An increase in vegetation height and leaf area augment the

transfer of sensible and latent heat away from the surface, while exerting a larger drag

force on the atmospheric boundary layer (Sellers et al., 1997). Similarly, the height of

vegetation cover affects the convergence of water vapor transport and rainfall

distribution, thus altering the near-surface climate (Sud et al., 1988).

1.1.1.2 Vegetation response to climate

The global dynamics of vegetation is primarily controlled by the prevailing climatic

constraints, in addition to other factors like edaphic conditions (nutrient availability and

soil texture) and anthropogenic disturbances. The response of vegetation to changes in

climate is evident across different spatial/temporal scales and climatic factors like

temperature, radiation and precipitation impose a varsity of complex limitations to plant

xxi

xxi

Page 22: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

growth in different parts of the world (Churkina et al., 1998; Nemani et al., 2003). Some

of the more prominent research on studies of vegetation response to climate are:

a) Recent reports indicate seasonal swings in green leaf area of about 25% in a

majority of the Amazon rainforests, with the seasonal cycle timed to the seasonality of

solar radiation in a manner that is suggestive of anticipatory and opportunistic patterns of

net leaf flushing during the early to mid part of the light rich dry season and net leaf

abscission during the cloudy wet season (Huete et al., 2006; Myneni et al., 2007).

b) The spatial patterns of the dominant modes of interannual co-variability in

Northern Hemisphere vegetation greenness, surface temperature and precipitation were

reported by Buermann et al., (2003) to be strongly linked to large scale circulation

anomalies such as the El Niño Southern Oscillation and the Arctic Oscillation.

c) Several researchers have reported on a progressively greening trend and longer

growing seasons in the northern hemisphere (Myneni et al., 1997; Zhou et al., 2001;

Slayback et al., 2003) and relaxation of climatic constraints to plant growth during the

1980s and 1990s that resulted in anomalous vegetation productivity (Nemani et al.,

2003). Most recent studies have also reported a decreasing trend in growing season

vegetation activity in the boreal forests of Southern Alaska, Canada and in the interior

forests of Russia (Angert et al., 2005; Goetz et al., 2005);

d) The availability of water critically limits plant growth in the semi-arid tropical

regions of the world, especially in grasslands where precipitation received in the wet

months is the primary driver of plant growth (Hickler et al., 2005; Prince et al., 2007).

Several recent studies on greening and increased precipitation have been reported in the

xxii

xxii

Page 23: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Sahelian region (Herrmann et al., 2005; Seaquist et al., 2006). The productivity of crops

and pastures in most of semi-arid and arid regions of Africa is predictably associated with

the El Niño Southern Oscillation and the North Atlantic Oscillation (Stige et al., 2006).

Their results suggest reduced food production in Africa if the global climate changes

toward more El Niño-like conditions, as most climate models predict. An extensive study

over the African savannas also show that the maximum woody cover in savannas

receiving a mean annual precipitation (MAP) of less than ~650 mm is inhibited by, and

increases linearly, with MAP (Sankaran et al., 2005).

e) Scholze et al., (2006) quantify the risks of climate-induced changes in key

ecosystem processes during the 21st century by forcing a dynamic global vegetation

model (DGVM) with multiple scenarios from 16 climate models. Their results show high

risk of forest loss in Eurasia, eastern China, Canada, Central America, and Amazonia,

with forest extensions into the Arctic and semiarid savannas. Malhi et al., (2008)

discusses the recent trends and fate of the Amazonian rainforests to projected risks of a

drying and changing climate.

These and other studies emphasize the fact that climate-vegetation interactions are a

vital component in defining the present as well as future needs of quantifying global land

cover change and terrestrial productivity in the context of climate change.

1.1.2 Long-term vegetation monitoring with satellite data

The Advanced Very High Resolution Radiometers (AVHRR) onboard the NOAA

series satellite platforms 7 to 16 provided the first long term global time series of data

xxiii

xxiii

Page 24: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

suitable for vegetation sensing (Tucker et al., 2005). The NASA Moderate Resolution

Imaging Spectroradiometer (MODIS) and Multi-angle Imaging SpectroRadiometer

(MISR) onboard Terra and Aqua platforms started delivering higher quality spectral and

angular measurements since February 2000 (Justice et al., 2002). These records are

expected to be extended by the planned Visible/Infrared Imager Radiometer Suite

(VIIRS) instrument on the NPOESS Preparatory Project (NPP) (Murphy et al., 2006).

Other long-term sources of data for vegetation monitoring include the Sea-viewing Wide

Field-of-view (SeaWiFS), Systeme Pour ‘Observation de la Terre (SPOT)

VEGETATION, and Environmental Satellite (ENVISAT) Medium Resolution Imaging

Spectrometers (MERIS).

A meaningful monitoring of vegetation requires a seamless consistent long-term

vegetation record obtained from multiple instruments but this is challenging because of

sensor differences and methodological issues (Van Leeuwen et al., 2006; Vermote and

Saleous, 2006; Brown et al., 2006). The scientific challenges include modeling the highly

variable radiative properties of global vegetation, scaling, and atmospheric correction of

data. The sensor related issues pertain to differences in sensor spectral characteristics,

spatial resolution, calibration, measurement geometry and data information content. The

consistency among biophysical variables derived from different sensors is thus a critical

issue in establishing the proper consensus for vegetation monitoring over a period of

several decades.

Among the aforesaid biophysical variables, LAI and FPAR are recognized as two of

the most important variables representative of vegetation structure and functioning that

xxiv

xxiv

Page 25: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

are commonly derived from satellite data (Running et al., 1986). The availability of data

from multiple sensors in the recent decade allows for rich spectral and angular sampling

of the radiation field reflected by vegetation canopies, thus increasing the potential for

obtaining accurate estimates of these biophysical variables. Long-term records of LAI

and FPAR are required by various terrestrial biosphere models, like the Terrestrial

Ecosystem Model (TEM) (Melillo et al., 1993), Biome-BGC (Running and Gower,

1991), Simple Biospheric Model (SiB) (Sellers et al., 1986), Integrated Biosphere

Simulated Model (IBIS) (Foley et al., 1996), Lund-Potsdam-Jena (LPJ) dynamic global

vegetation model in Land Surface Model (LSM) (Bonan et al., 2003) and the

Atmospheric-Vegetation Interactive Model (AVIM) (Jinjun et al., 1995), to investigate

the response of ecosystems to changes in climate, carbon cycle, land cover and land use.

1.2 LAI/FPAR algorithm and products

1.2.1 Definition of LAI and FPAR

Leaf area index (LAI) is defined as the one-sided green leaf area per unit ground

area in broadleaf canopies and as half the total needle surface area per unit ground area in

coniferous canopies. LAI characterizes the functioning surface area of a vegetation

canopy (Myneni et al., 2002). The interactions between the vegetation surface and the

atmosphere, for example., radiation exchange, transpiration rates, precipitation

interception, momentum and gas exchange, is predominantly determined by the total leaf

area (Monteith and Unsworth, 1990). An increase in leaf area, for example, increases the

xxv

xxv

Page 26: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

uptake of CO2 from the atmosphere due to greater sunlight absorption and hence results

in increased canopy conductance and transpiration rates (Field and Mooney, 1983; cf.

Section 1.1.1.1 on leaf area feedback). Field measurements of LAI include hemispherical

photography and optical instruments like TRAC, LAI-2000 or LI-COR (Chen et al.,

1997). Satellite remote sensing enables derivation of LAI globally at desired spatial

resolution and temporal frequency with algorithms based on the physics of radiative

transfer.

Another parameter that characterizes the energy absorption capacity of a vegetation

canopy is FPAR, defined as the fraction of photosynthetically active radiation (0.4 -

0.7μm) absorbed by the vegetation canopy. FPAR depends on the incident radiation field,

architecture and optics of the canopy, and the reflectance of the soil background. FPAR is

well related to NDVI and usually increases with ground cover and plant leaf area

(Myneni and Williams, 1994). It is one of the fundamental parameters used to estimate

net primary production and for modeling of terrestrial carbon processes (Seller et al.,

1986; Pitman, 2003; Knorr et al., 2005). Similar to LAI, FPAR has also been identified as

one of the fundamental terrestrial state variables in the context of global change studies

(GCOS 2nd Adequacy report).

1.2.2 LAI/FPAR algorithms

There is considerable literature on the estimation of LAI from vegetation indices

like the Normalized Difference Vegetation Index (NDVI) and simple ratio (Asrar et al.,

1984; Chen and Cihlar, 1996). In particular, Sellers et al., (1996) introduced an empirical

xxvi

xxvi

Page 27: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

algorithm that calculated FPAR as a function of the simple ratio. Lu and Shuttleworth

(2002) used this definition of FPAR and approximated the relationship between LAI and

FPAR to be exponential (Monteith and Unsworth, 1990) for evenly distributed

vegetation. Strong positive correlations were found between LAI and NDVI for various

vegetation types (Myneni et al., 1997), as well as with simple ratio in coniferous forests

(Chen and Cihlar, 1996). Site-specific NDVI-LAI empirical relationships have been used

in various ecosystems (Fassnacht et al., 1997; Colombo et al., 2003), but with limited

success when applied across sites and vegetation classes.

The sensitivity of NDVI to LAI is controlled by the relationship between NDVI and

fractional vegetation cover when LAI is in the range of about 2 to 4 (Carlson and Ripley,

1997). Steltzer and Welker (2006) incorporated fractional cover of photosynthetic

vegetation for multiple species into the exponential NDVI-LAI model for a regional scale

analysis, and suggested that species composition affects the NDVI-LAI relationship

through leaf-level properties (leaf optics, leaf structure and orientation) and canopy-level

structural properties that influence the vertical and horizontal distribution of leaf area

within a canopy. It is thus clear that NDVI-LAI empirical relationships do vary across

different species and are sensitive to canopy structure and fractional ground cover. These

empirical relationships can also vary both seasonally and inter-annually with respect to

phenological development of the vegetation. Thus, a relationship established between

LAI and NDVI in a particular year may not be applicable in other years (Wang et al.,

2004). Consequently, the empirical relationships will be site-, time-, and species-specific,

xxvii

xxvii

Page 28: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

and, therefore, poorly suited for large-scale operational use (Barot and Guyot, 1991;

Houborg et al., 2007).

An alternate approach is to use physically based models that describe the interaction

of radiation inside a canopy based on physical principles and provide an explicit

connection between biophysical variables and canopy reflectance (Weiss et al., 2002;

Combal et al., 2002). The physical models of radiation transfer and interaction in

vegetation canopies are usually categorized into four broad types: (1) radiative transfer

models (Myneni et al., 1989; Knyazikhin et al., 1998a), (2) geometrical optical models

(Li and Strahler, 1986; 1992), (3) hybrid models that incorporate both radiative transfer

as well geometric optics (Welles and Norman, 1991), and (4) Monte-Carlo simulation

models (Ross et al., 1988; Lewis, 1999). For a physically based algorithm, which

involves the inverse problem of estimating LAI and FPAR, a Look-Up-Table (LUT) is

employed, where the simulated reflectances are pre-calculated for a range of conditions

with the model of choice and stored in LUTs. The remotely sensed satellite

measurements are then compared with the LUT entries to find a solution set of LAI. In

the MODIS and MISR approach to LAI retrievals, the algorithm generates a set of

acceptable solutions, i.e., all values of LAI and soil reflectance for which modeled and

observed Spectral Surface Reflectance (SSR) differ by an amount equivalent or less than

the combined uncertainty in model and observations (Fig. 1.1).

xxviii

xxviii

Page 29: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

1.2.3 LAI and FPAR products from different sensors

The first global data set of LAI and other biophysical products derived from

satellite data is the FASIR data set (Fourier Adjusted, Solar zenith angle corrected,

Interpolation of missing data, and Reconstruction of data classified as tropical evergreen

broadleaf forests). This data set is a 1 by 1 degree monthly composite based on AVHRR

data from 1982 to 1990. A limitation of this data set is that it is based on a LAI-NDVI

empirical relationship without considering the sensitivity of this relation to vegetation

type and background effects. Similarly, other long-term LAI/FPAR data sets were

derived from improved versions of the AVHRR NDVI data (PAL and GIMMS NDVI)

(Myneni et al., 1997; Buermann et al., 2002).

With the launch of several moderate and coarse spatial resolution satellite sensors,

such as MODIS, MISR, MERIS, POLDER, VEGETATION, etc., there has been a

significant effort to produce global as well as regional data sets of LAI and FPAR

(Knyazikhin et al., 1998a; Gobron et al., 1999; Chen et al., 2002; Baret et al., 2007;

Plummer et al., 2006). Some of these LAI/FPAR products are given in Table 1 of Weiss

et al., (2007). The earliest of these records (Chen et al., 2002) starts from 1993 to the

present. The MODIS LAI/FPAR product is based on a radiative transfer algorithm and is

the first global scale operational LAI/FPAR product, spanning the period March 2000 to

the present. The Canadian Centre for Remote Sensing is similarly generating a Canada-

wide time series of LAI and FPAR from AVHRR and VEGETATION at 1 km resolution

as 10-day composites using empirical algorithms (Chen et al., 2002). Likewise, the Joint

Research Center is developing time series of FPAR (2002-present) from SeaWiFS,

xxix

xxix

Page 30: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

MERIS, and VEGETATION (Gobron et al., 2006). The CYCLOPES (version 3.1) LAI

product (1998-2003) is a new data set derived from the SPOT/VEGETATION sensor at

10-day frequency over a 1/112o plate-carrée spatial grid (Baret et al., 2007).

1.3 Objectives and structure of this dissertation

The monitoring and modeling of the terrestrial biosphere within the larger context

of climate variability and change studies requires multi-decadal time series of key

biophysical variables characteristic of vegetation structure and functioning (NRC

Decadal Survey, 2007; GCOS 2nd Adequacy Report, 2007). Consequently, there is now a

pressing need to develop methodologies for generating continuous long-term Earth

System Data Records (ESDRs) from remote sensing data collected with different sensors

over the past three decades. LAI and FPAR are well defined and measurable

characteristics of vegetation and are independent of the properties of satellite sensors.

Therefore, one should obtain true values of LAI and FPAR irrespective of the sensor used

(AVHRR or MODIS). A key step in assembling these long-term data sets is establishing

a link between data from earlier sensors (e.g. AVHRR) and present/future sensors (e.g.

MODIS TERRA, NPOESS) such that the derived products are independent of sensor

characteristics and represent the reality on the ground both in absolute values and

variations in time and space (Van Leeuwen et al., 2006). Multi-decadal global data sets of

LAI and FPAR of known accuracy and produced with a physically based algorithm are

currently not available, although several recent attempts have resulted in shorter term

research quality data sets from medium resolution sensor data (Knyazikhin et al., 1998a;

xxx

xxx

Page 31: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Gobron et al., 199; Chen et al., 2002; Yang et al., 2006a; Plummer et al., 2006; Baret et

al., 2007). Thus the motivation for this research was to formulate and demonstrate the

performance of a synergistic approach for LAI and FPAR Earth System Data Record

(ESDR) retrievals from measurements of multiple satellite sensors. To test the robustness

and validity of the physical retrieval algorithm, it is crucial to validate the long-term

vegetation data record directly with field measurements and indirectly with data records

from present sensors. The scarcity of field measurements during the 1980s and 90s

represent a more challenging problem, and hence evaluation of the data record should

also be performed by correlating anomalies in the long-term vegetation data record with

different climate surrogates like temperature and precipitation, thus reproducing well

documented trends in vegetation growth vis-à-vis changes in climate. Therefore, the

objectives of this research are:

(a) Development of a theoretical formulation and its implementation as an algorithm that

will generate true values of LAI and FPAR independent of sensor characteristics;

(b) Production, evaluation and validation of a long-term LAI data set from AVHRR

NDVI for the period July 1981 till December 2006;

(c) A case study to assess vegetation response to changes in precipitation and land use in

the semi-arid regions for the period 1981 to 2006, utilizing the long-term LAI data

record.

xxxi

xxxi

Page 32: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

1.3.1 Theoretical formulation for generating LAI and FPAR

In Chapter 2, a physically based methodology based on the radiative transfer theory

of spectral invariants is developed to derive LAI from AVHRR GIMMS NDVI data.

Achieving consistency in the generated LAI and FPAR fields is of utmost importance due

to differences in spatial and spectral resolution of sensors, uncertainties due to calibration

and atmospheric effects, and input information content. First, a physical basis to account

for adjustments due to differences in sensor spatial resolution and spectral bandwidths is

formulated. Second, the requirements for consistency due to differences in input

information content (spectral surface reflectance from MODIS and NDVI from AVHRR)

are specified. The generic mode of the algorithm ingests NDVI from AVHRR data and

generates a solution set of LAI, whose mean value should correspond to MODIS LAI.

The agreement between mean values and corresponding dispersions satisfies the

consistency requirements. Thirdly, the algorithm is implemented for different vegetation

types through parameterization in order to retrieve consistent LAI fields. The accuracy

and dispersion of the generated LAI is also an important aspect of this research, and are

addressed in both Chapters 2 (implementation phase) and 3 (production phase).

1.3.2 Global LAI from AVHRR NDVI: production, evaluation and validation

The evaluation of a new global monthly AVHRR LAI data set for the period July

1981 to December 2006 is presented in Chapter 3. The implementation of the algorithm

for retrieving LAI from AVHRR NDVI and production of the global data set is first

detailed. The evaluation of the data set is done both through direct comparisons to ground

xxxii

xxxii

Page 33: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

data and indirectly through inter-comparisons with other similar data sets. The indirect

validation included comparisons with existing LAI products (MODIS and CYCLOPES

LAI products for the 2000-2003 period of overlap), at a range of spatial scales, and

correlations with key climate variables in areas where temperature and precipitation limit

plant growth. Indirect validation with climate variables is an important part of the study

as long-term field measurements of LAI are few. One alternate approach to evaluation of

the LAI data set for the earlier time frame (1981-1999) is to assess the magnitudes and

inter-annual variations in the LAI fields with respect to known trends in climate

variables. Finally, in a direct validation approach, the comparison of the LAI dataset with

field measurements of LAI and high resolution field LAI maps obtained via the ORNL

DAAC Mercury system is performed.

1.3.3 Vegetation changes versus climate/land use change in the semi-arid tropics

The availability of a long-term consistent LAI data set permits investigation of

changes in vegetation activity. The semi-arid tropics, stretching across a large number of

developing countries with rapidly increasing population, and characterized by low and

highly variable precipitation, are projected to be impacted by desertification and overall

declines in vegetation productivity by ongoing and future climate changes. In these

regions, reductions in vegetation activity and expansion of desertification are expected to

arise from drier conditions due to continued warming trends accompanied by the

prediction of a reduction in precipitation (IPCC 2007) and low adaptation capacity of the

affected communities (Parry et al., 2007). In spite of these climatic changes, which would

xxxiii

xxxiii

Page 34: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

suggest that tropical dry lands are already becoming drier, satellite observations of

vegetation greenness provide evidence that primary productivity over extensive portions

of the semi-arid tropics has been on the rise (Tucker and Nicholson, 1999; Eklundh and

Olsson, 2003; Hermann et al., 2005). Other driving factors, apart from climatic correlates

like precipitation and temperature, such as changes in land cover and land use, and

fertilization effects due to atmospheric increases in C and N, have been invoked more

loosely to explain vegetation dynamics in these tropical dry lands. Thus, to assist in

projecting the impacts of climate change on ecosystems and societies, it is crucial that the

changes in the drivers of ecosystem dynamics are properly understood. With the goal of

identifying the relative contributions and spatial distribution of climate and socio-

economic and land use change in promoting the greening of the tropical dry lands,

changes in AVHRR LAI in conjunction with changes in climatic and land use data for the

period 1981-2006 are analyzed (Chapter 4). This study focuses on the semi-arid tropics of

the eastern hemisphere, where the largest contiguous dry lands are inhabited by close to

1.7 billion people and spread across 120 countries, most of which are among the poorest

countries in the world and with the lowest human development index. In analyzing the

changes in vegetation greenness, special emphasis is placed in the context of changes in

socio-economic and land use change data for India, where high resolution data are

available at the national scale. This case study of understanding vegetation changes in

semi-arid lands over a 26 year period highlights the utility of the derived LAI data set.

xxxiv

xxxiv

Page 35: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Sensor Independent Algorithm

Compares input SSR to simulated SSR using a merit function

Sensor Dependent LUTCanopy Architecture Radiative

Transfer ProductConfigurable Parameters

InputSpectral Surface

Reflectance (SSR) and

Landcover map

Set of Acceptable Solutions

(intermediate product)

Output

Mean LAI, FPAR and their dispersions

Figure 1.1 Flowchart of the LAI retrieval process. In the MODIS and MISR approach to LAI retrievals, the algorithm first generates a set of acceptable solutions, i.e., all values of LAI and soil reflectance for which modeled and observed Spectral Surface Reflectance (SSR) differ by an amount equivalent or less than the combined uncertainty in model and observations. FPAR is calculated for each acceptable solution. From this set, mean LAI, FPAR and their dispersions are calculated. The algorithm requires a sensor specific Look-up-Table (LUT) to rapidly model the radiative transfer process of complex canopy/soil models to determine the SSR and canopy absorption. The LUT includes (a) Canopy Architecture Radiative Transfer Product containing structural characteristics of the vegetation canopies and radiative transfer parameters and (b) a small set of configurable parameters needed to adjust the LUT for sensor characteristics, data resolution and uncertainties.

xxxv

xxxv

Page 36: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Chapter 2

2. Generating Leaf Area Index Earth System Data Record from

multiple sensors: Theory

2.1 Introduction

The monitoring and modeling of the terrestrial biosphere within the larger context

of climate variability and change studies requires multi-decadal time series of key

variables characteristic of vegetation structure and functioning (NRC Decadal Survey,

2007; GCOS 2nd Adequacy Report, 2007). Consequently, there is now a pressing need to

develop methodologies for generating continuous long-term Earth System Data Records

(ESDRs) from remote sensing data collected with different sensors over the past three

decades. This study focuses on two key biophysical variables, leaf area index (LAI) and

fraction vegetation absorbed photosynthetically active radiation (FPAR), that control the

exchange of energy, mass (e.g. water and CO2) and momentum between the Earth surface

and atmosphere (Dickinson et al., 1986; Potter et al., 1993; Sellers et al., 1996; Tian et

al., 2004; Demarty et al., 2007).

The Advanced Very High Resolution Radiometers (AVHRR) onboard NOAA 7-14

series satellite platforms delivered the first high temporal resolution global time series of

data suitable for vegetation sensing starting from July 1981 (Tucker et al., 2005). The

NASA Moderate Resolution Imaging Spectroradiometer (MODIS) and Multi-angle

Imaging SpectroRadiometer (MISR) onboard Terra and Aqua platforms started providing

xxxvi

xxxvi

Page 37: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

higher quality spectral and angular measurements since February 2000 (Diner et al.,

1999; Justice et al., 2002). These records are expected to be extended by

the planned Visible/Infrared Imager Radiometer Suite (VIIRS)

instrument onboard the NPOESS Preparatory Project (NPP) to be

launched in the near future (Murphy, R. E., 2006). Other long-term

sources of data for vegetation monitoring include the Sea-viewing Wide

Field-of-view Sensor (SeaWiFS), Systeme Pour l'Observation de la Terre (SPOT)

VEGETATION, and ENVISAT Medium Resolution Imaging Spectrometers (MERIS).

The challenges underlying the generation of continuous time series of land products

from data of multiple instruments include both remote sensing science and sensor-related

issues. The scientific challenges include modeling the highly variable radiative properties

of global vegetation, scaling, and atmospheric correction of data. Traditionally, the

Normalized Difference Vegetation Index (NDVI) has been used for long-term global

vegetation monitoring (Myneni et al., 1997). Biophysical parameters (LAI and FPAR)

have been retrieved from NDVI using empirical relationships (Sellers et al., 1996).

However, those relationships are site-, time-, and biome-specific and their use in global

operational production may be limited (Baret and Guyot, 1991; Wang et al., 2004).

Scaling issues (mixture of different vegetation types) introduce an additional bias as

vegetation classes with relatively low pixel fractional coverage are under-represented in

coarse resolution retrievals (Tian et al., 2002; Steltzer and Welker, 2006; Shabanov et al.,

2007). Finally, the retrieval of biophysical parameters require surface reflectances,

xxxvii

xxxvii

Page 38: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

however, a complete atmospheric correction of satellite measurements was not performed

in the past due to limited availability of requisite ancillary data.

The hardware issues include differences in sensor spectral characteristics, spatial

resolution, calibration, measurement geometry and data information content. Differences

in sensor spectral bands (central wavelength and bandwidth) result in differential

sensitivity of the sensor’s spectral response functions (SRF) to the impact of Rayleigh

scattering, ozone, aerosol optical thickness, water vapor content and reflection from the

ground (Vermote and Saleous, 2006; Van Leeuwen et al., 2006). The variation in spatial

resolution involves the impact of sensor-dependent point spread function (PSF), such that

radiometric measurements for a particular pixel are partially mixed with those of adjacent

pixels and re-sampling to the common resolution almost always results in a bias (Tan et

al., 2006). Calibration corrections in post processing of raw data results in varying

sensitivity of satellite image Digital Numbers (DN) to recorded radiation (Vermote and

Saleous, 2006). The calibration issues are further complicated by orbital drift and related

changes in illumination/observation geometry (Gutman et al., 1998). Finally, the

information content of measurements will vary between sensors, and retrieval techniques

should take advantage of available multi-angular, multi-spectral, high spatial or temporal

resolution measurements.

The two prominent NDVI time series data are the Pathfinder AVHRR Land (PAL)

and Global Inventory Monitoring and Modeling Studies (GIMMS) (Tucker et al., 2005).

These data sets cover nearly the entire record (July 1981 to the present) at 8-km spatial

resolution as 15-day temporal composites (PAL is 10-day composite). The data

xxxviii

xxxviii

Page 39: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

processing included calibration, interpolation of missing data, and partial atmospheric

correction with statistical techniques. Several studies reveal significant trends in NDVI

over the Northern high latitudes (Myneni et al., 1997; Zhou et al., 2001); however the

accuracy of the assessment was questioned (Gutman, 1998). The Canadian Center for

Remote Sensing is routinely generating Canada-wide time series of LAI and FPAR from

AVHRR and VEGETATION at 1-km resolution as 10-day composites using empirical

algorithms (Chen et al., 2002). Likewise, the Joint Research Center is currently

developing time series of FPAR from SeaWiFS, MERIS, and VEGETATION (Gobron,

et al., 2006).

Multi-decadal global data sets of LAI and FPAR of known accuracy and produced

with a physically based algorithm are currently not available. Efforts are underway to

perform rigorous physically based calibration and atmospheric correction to achieve

consistency with the reference MODIS NDVI records (Vermote and Saleous, 2006).

These efforts should result in higher quality surface reflectance data ideally suited for

producing LAI and FPAR ESDRs. Thus, the objectives of this research are to formulate

and demonstrate the performance of a synergistic approach for LAI and FPAR ESDR

retrievals from measurements of multiple satellite sensors. The theoretical aspects of the

approach are presented in this study, while results from evaluation of the quality of the

generated data series are presented in Chapter 3. This research is organized as follows.

The criteria for ensuring consistency between retrievals from different sensors are

formulated in Section 2.2. The next section introduces the theoretical basis of a multi-

sensor retrieval algorithm, namely, parameterization of the canopy spectral reflectance

xxxix

xxxix

Page 40: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

using the radiative transfer theory of spectral invariants. Methods for accounting

differences in sensor spatial resolution and spectral bands are presented in Sections 2.4

and 2.5. The following section describes an approach to adjust the retrieval technique for

handling variations in the information content of satellite data. Finally, the concluding

remarks are given in Section 2.7.

2.2 Criteria for ensuring consistency

Generation of ESDRs from observations of multiple instruments requires deriving

an inter-sensor consistent product which also matches well with ground truth

measurements. A one-to-one relationship between remote observations and a land

parameter of interest can be achieved only in the case of error-free measurements

delivering sufficient information content. In practice, the retrieval of LAI and FPAR from

satellite data should be treated as an ill-posed problem; that is, small variations in input

data due to uncertainties in measurements can result in a change in the relationship,

leading not only to non-physically high variations in the retrieved values but also to the

loss of a true solution, since it may not satisfy the altered relationship (Wang et al., 2001;

Combal et al., 2002; Tan et al., 2005). Input data and their uncertainties are, “in general,

the minimal information necessary to construct approximate solution for ill-posed

problems” (p.3, Tikhonov et al., 1995). The inclusion of more measured information

(spectral and/or angular variation) tends to improve the relationship between satellite

observations and the desired parameters. This however not only increases the overall data

information content but also increases their overall uncertainty. The former enhances the

xl

xl

Page 41: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

quality of retrievals while the latter suppresses it. Therefore, the specification of an

optimal combination of data information content and overall uncertainty is a key task to

achieving continuity in the multi-sensor time series of LAI and FPAR products.

In general, the information conveyed by surface reflectances is not sufficient to

retrieve a unique LAI value. For example, different combinations of LAI and soil types

can result in the same value of canopy spectral reflectances; or different spectral

reflectances can correspond to the same LAI value but for different vegetation types

(Diner et al., 2005). A particular observation of surface reflectance is therefore associated

with a set of canopy parameter values and are referred to as the set of acceptable

solutions (Knyazikhin, et al., 1998a). This set of solutions depends on the properties of

measured surface reflectances: absolute values and uncertainties, spectral characteristics,

spatial resolution, and observation geometry. In general, a larger volume and higher

accuracy of the measured information corresponds to a better localized set of solutions.

The solution set “size” can, therefore, be used as a measure of the data information

content. This concept is used to formulate the following requirements for a multi-sensor

algorithm to generate consistent LAI and FPAR retrievals from AVHRR and MODIS

sensors:

(a) The algorithm should generate a set of acceptable solutions given AVHRR

NDVI;

(b) This set should include all acceptable solutions generated by the MODIS

algorithm when given the corresponding AVHRR spectral reflectances;

xli

xli

Page 42: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

(c) The algorithm should also be capable of admitting AVHRR spectral

reflectances, in addition to NDVI, and generate the same set of acceptable

solutions as the MODIS algorithm.

In the above formulation, Terra MODIS LAI and FPAR products serve as the benchmark.

2.3 Parameterization for canopy spectral reflectance

Retrievals of the LAI/FPAR ESDR from multiple sensors require parameterization

of the retrieval algorithm that can be adjusted for the specific features of the Bidirectional

Reflectance Factor (BRF) measurements by a particular sensor (spatial resolution,

bandwidth, calibration, atmospheric correction, information content, etc, cf. Section 2.1).

The radiative transfer theory of canopy spectral invariants provides the required BRF

parameterization via a small set of well-defined measurable variables that specify the

relationship between the spectral response of vegetation canopy bounded below by a non-

reflecting surface to the incident radiation at the leaf and canopy scales (Huang et al.,

2007; Lewis and Disney, 2007; Smolander and Stenberg, 2005).

2.3.1 Canopy spectral invariants

Photons that have entered the vegetation canopy undergo several interactions with

leaves before either being absorbed or exiting the medium through its upper or lower

boundary (Fig. 2.1). Interacting photons can either be scattered or absorbed by a

phytoelement. The probability of a scattering event, or leaf single scattering albedo, ,

depends on the wavelength and is a function of the leaf structural biochemical

xlii

xlii

Page 43: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

constitution. If objects are large compared to the wavelength of the radiation, e.g., leaves,

branches, etc., the photon free path between two successive interactions is independent of

the wavelength. The interaction probabilities for photons in a vegetation media, therefore,

are determined by the structure of the canopy rather than photon frequency or the optical

properties of the canopy. To quantify this feature, Smolander and Stenberg (2005)

introduced the notion of recollision probability, p, defined as the probability that a photon

scattered by a foliage element in the canopy will interact within the canopy again. This

spectrally invariant parameter is a function of canopy structural arrangement only (Huang

et al., 2007; Lewis and Disney, 2007). Scattered photons can escape the vegetation

canopy either through the upper or lower boundary. Their angular distribution at the

upper boundary is given by the directional escape probability, () (Huang et al., 2007).

Given recollision, pm, and escape, m(), probabilities as a function of scattering order, m,

the bidirectional reflectance factor, BRFBS,(), for a vegetation canopy bounded below

by a non-reflecting surface can be expanded in a series of successive orders of scattering

(Huang et al., 2007).

. (2.1)

Here i0 is the probability of initial collision, or canopy interceptance, defined as the

proportion of photons from the incident beam that are intercepted, i.e., collide with

foliage elements for the first time. This parameter gives the proportion of shaded area on

the ground which in turn is directly related to the proportion of the sunlit leaf area.

Canopy interceptance does not depend on the wavelength and is a function of the

direction of the incident beam and canopy structure.

xliii

xliii

Page 44: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

In the general case, the recollision and escape probabilities vary with the scattering

order m. For m=1, the directional escape probability coincides with the bi-directional gap

probability. These probabilities, however, reach plateaus as the number of interactions m

increases. Monte Carlo simulations of the radiation regime in 3D canopies suggest that

the probabilities saturate after 2 to 3 interactions for low to moderate LAI canopies

(Lewis and Disney, 2007) with the recollision probability exhibiting a much faster

convergence (Huang et al., 2007). Neglecting variations in pm with m (i.e., pmconst=p)

and in m() for m>1 (i.e., m()const=2() for m≥2) in Eq. (2.1), one obtains the

first order approximation for the BRFBS, (Huang et al., 2007)

. (2.2)

Here R1()=1()i0 and R2()=2()pi0 are the escape probabilities expressed relative to

the number of incident photons. The accuracy of this first order approximation depends

on the difference between successive approximation to p multiplied by the factor

(Huang et al., 2007).

Under the above assumption regarding dependence of the recollision probability on

the scattering order, the spectral absorptance, aBS, of the vegetation canopy with non-

reflecting background can be expressed as (Fig. 2.1)

. (2.3)

The corresponding FPAR is a weighted integral of Eq. (2.3) over the PAR spectral region

(Knyazikhin et al., 1998a). According to Smolander and Stenberg (2005) Eq. (2.3)

xliv

xliv

Page 45: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

provides an accurate estimate of canopy spectral absorptance. A detailed analysis of the

approximations given by Eqs. (2.2) and (2.3), their accuracies as well as how the

recollision probability and canopy interceptance can be accurately measured in the field

are discussed in Huang et al. (2007).

2.3.2 Canopy-ground interactions

The three-dimensional radiative transfer problem with arbitrary boundary conditions

can be expressed as a superposition of some basic radiative transfer sub-

problems with purely absorbing boundaries and to which the notion of

spectral invariant can be directly applied (Knyazikhin and Marshak,

2000). These two problems are: (1) the black soil problem, “BS-

problem” specified by the original illumination conditions at the top of

the canopy and a completely absorbing soil at the bottom; (2) the soil

problem, “S-problem” specified by no input energy at the top, but

Lambertian energy sources at the bottom. This decomposition

technique was implemented in the MODIS LAI/FPAR operational

algorithm (Knyazikhin et al., 1998a). According to this approach, the

spectral BRF and canopy spectral absorptance are approximated as

,

(2.4)

. (2.5)

xlv

xlv

Page 46: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

The second term on the right hand side of Eqs. (2.4) and (2.5) describes the contribution

to the BRF and absorptance from multiple interactions between the ground and

vegetation (cf. Appendix A). Here, sur, is an effective ground reflectance, and tBS, is the

transmittance of the vegetation canopy for the BS-problem. Variables rS,, aS,, and JS,()

represent solutions to the “S-problem” (cf. Appendix A). The expansion in successive

order of scattering as given by Eq. (2.1) and illustrated in Fig. 2.1 is also applicable to the

“S-problem” with the only difference that i0 is replaced with i0,S, the proportion of

photons from sources below the canopy that are intercepted (i.e., those that collide with

foliage elements for the first time). A full set of equations describing canopy-ground

interaction is given in Appendix A.

Thus, a small set of well defined measurable variables provide an accurate

parameterization of canopy optical and structural properties required to fully describe the

spectral response of a vegetation canopy to incident solar radiation. This set includes

spectrally varying soil reflectance (sur,), single-scattering albedo (), spectrally

invariant canopy interceptances (i0 and i0,S), recollision probability (p) and the directional

escape probability (1() and 2()) and their hemispherically averaged values.

2.3.3 Generation of structural parameters

The global classification of canopy structural types utilized in the Collection 5

MODIS LAI/FPAR algorithm was adopted in this study (Shabanov et al., 2005; Yang et

al., 2006a). According to this classification, global vegetation is stratified into eight

canopy architectural types or biomes: (1) grasses and cereal crops, (2) shrubs, (3)

xlvi

xlvi

Page 47: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

broadleaf crops, (4) savannas, (5) evergreen broadleaf forests, (6) deciduous broadleaf

forests, (7) evergreen needle leaf forests and (8) deciduous needle leaf forests. The

structural attributes of these biomes are parameterized in terms of variables that the

transport theory admits (Knyazikhin et al., 1998a). The stochastic radiative transfer

equation was used to generate the Collection 5 Look-up-Tables (LUT) - a set of tabulated

BRF values as a function of biome type, LAI, view/illumination geometry, etc.

(Shabanov et al., 2000; Huang et al., 2008). The aim is to generate the spectrally invariant

parameters for which the spectral BRF and absorptance coincide with values of the

Collection 5 MODIS LUTs for all combinations of LUT entries. Given these parameters,

the BRF and absorptance for specific wavelengths can be calculated using Eqs. (2.4)-

(2.5) and (A2 in Appendix A)-(A4 in Appendix A) with varying single scattering albedo

which is used as the tuning parameter to adjust the LUTs for data spatial resolution and

spectral band characteristics (cf. Sections 2.4 and 2.5).

The canopy interceptances can be directly calculated using the stochastic radiative

transfer equation. Figure 2.2 shows i0 and i0,S as a function of LAI for one example

vegetation type (savannas). Isotropic diffuse sources below the canopy are used to

specify the interceptance i0,S. Notably, i0,S is greater than i0, that is, interception is higher

under diffuse illumination conditions (Min, 2005; Gu et al., 2002), which is captured in

the simulations.

The stochastic radiative transfer equation is used to simulate the solutions of the

“BS-problem” (BRFBS, aBS and tBS) and “S-problem” (aS, rS, and tS) as a function of the

single scattering albedo, , for various LAI and sun-view geometries. For given LAI and

xlvii

xlvii

Page 48: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

sun-view geometry, the spectrally invariant parameters are obtained by fitting the

analytical approximations (cf. Eqs. (2.2)-(2.3) and Appendix A) to their simulated

counterparts. The parameters thus obtained are functions of LAI and sun-view geometry.

Figure 2.3 shows the recollision probabilities calculated by fitting solutions of the “BS-

problem” and “S-problem” to their simulated values. As expected, these variables are

close to each other (Huang et al., 2007).

Given spectrally invariant parameters, the reflectances (rBS and rS), transmittances

(tBS and tS), and absorptances (aBS and aS) are calculated for varying LAI and single

scattering albedo values using the spectral invariant approximations and checked for

validity of the energy conservation law, r+t+a=1, for both the “BS-” and “S-” problems

(Fig. 2.4). Finally, BRFs at red and NIR spectral bands corresponding to values of single

scattering albedo of Collection 5 MODIS LUT are calculated as a function of the

effective ground reflectances at red and NIR wavelengths, LAI and sun-view geometry.

The BRF values are compared with corresponding values stored in the Collection 5

MODIS LUTs to ensure consistency (Fig. 2.5).

2.4 Algorithm adjustment for the sensors spatial resolution

The scaling approach in this study is based on the scale dependence of the single

scattering albedo. This variable is defined as the probability that a photon intercepted by

foliage elements in volume V will escape V. The volume V is associated with the scale at

which the single scattering albedo is defined, e.g., single leaf, clump of leaves, tree

crown, patch, or even a satellite pixel. The theory of canopy spectral invariants provides

xlviii

xlviii

Page 49: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

an accurate description of variations in the single scattering albedo with scale V

(Smolander and Stenberg, 2005; Lewis and Disney, 2007). In the present approach, BRF

is an explicit function of the single scattering albedo and thus this theory can be applied

to imbue scale dependence to the algorithm. The aim of this section is to demonstrate

how the canopy spectral invariant relationships can be employed to adjust solutions of the

“BS” and “S” problems for sensor resolutions.

Consider two volumes, V0 and V, representing pixel and tree crown scales. Their

single scattering albedos (V0) and (V) quantify the scattering properties of the pixel

V0 and its constituent objects of volume V. The latter are distributed within V0 in a certain

fashion. It follows from Eq. (2.3) that the pixel single scattering albedo, (V0), can be

estimated as (Smolander and Stenberg, 2005; Lewis and Disney, 2007)

= .

(2.6)

Here i0(V0) is the proportion of photons intercepted by the volume V0 (Fig. 2.1), and

p(VV0) is the recollision probability, defined as the probability that a photon scattered

by volume V (e.g., by a tree crown) resident in the pixel V0 will hit another volume V

(e.g., another tree crown) in the same pixel. Its value is determined by the distribution of

volumes V (e.g., tree crowns) within V0.

Equation (2.6) links canopy spectral behavior at the pixel and tree crown scales.

Indeed, the canopy single scattering albedo (V0) (pixel scale) is an explicit function of

the spectrally varying single scattering albedo (V) at the tree crown scale V and the

xlix

xlix

Page 50: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

spectrally invariant recollision probability p(VV0). The latter is a scaling parameter that

accounts for the cumulative effect of canopy structure from tree crown to pixel scales.

Both (V) and p(VV0) vary with scale V. For example, the single scattering albedo and

the recollision probability associated with needle, shoot, branch, tree crown, etc., are

different (Smolander and Stenberg, 2003). However since the left-hand side of Eq. (2.6)

does not depend on V, the algebraic expression on the right-hand side of this equation

should also be independent of the leaf scale V. Based on this property, Smolander and

Stenberg (2005) specified variation in the leaf single scattering albedo and the recollision

probability with the scale V as follows.

Consider the scale V (e.g., tree crown) which in turn consists of smaller objects (e.g.

clump of leaves) distributed in V. Let and represent the scale of the object and

its single scattering albedo. Equation (2.6) can also be applied to the volume V, i.e.,

. (2.7)

Substitution of this equation into Eq. (2.6) results in the same equation for (V0) with the

only difference that (V) is replaced with and p(VV0) is replaced with a new

recollision probability calculated as

. (2.8)

One can see the probability that a photon scattered by a volume (e.g.,

clump of leaves) will interact within the pixel V0 again follows the Bayes’ formula. The

recollision probability, therefore, is a scaling parameter that accounts for the cumulative

effect of multi-level hierarchical structure in a vegetated pixel.

l

l

Page 51: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Smolander and Stenberg (2003; 2005) demonstrated the validity of the scaling

relationships for needle ( =needle) and shoot (V=shoot) scales. Lewis and Disney

(2007) found that Eqs. (2.6)-(2.8) are applicable to the within leaf ( =a within-leaf

scattering object) and leaf (V=leaf) scales, implying that scaling equations provide a

framework through which structural information can be maintained in a consistent

manner across multiple scales from within-leaf to canopy level scattering. Tian et al.

(2002) used a semi-empirical approach to account for biome mixtures within a coarse

resolution pixel ( =biome type in a patch V).

The scaling properties of the scattering process underlies the following approach for

developing scale-dependent formulation of the radiative transfer process in vegetation

canopies. First, one defines a base scale, V, in a canopy-radiation model, e.g., tree crown,

patch, etc. The structure-dependent coefficients that appear in the radiative transfer

equation are parameterized in terms of the distribution of objects in the volume V within

the pixel V0 and thus are independent of the structure that exists within V. The concepts of

the pair-correlation function (Huang et al., 2008) and biome mixtures (Shabanov et al.,

2007) are used to obtain these coefficients. Second, the single scattering albedo (V) of

the object (which also appears in the equation) is calculated using Eqs. (2.7) and (2.8).

The radiative transfer equation describes the interaction between photons and objects of

the volume V while multiple scattering within V is accounted by the single scattering

albedo (V).

In the parameterization of this approach (cf. Section 2.3), canopy reflectance and

absorptance are explicit functions of structural parameters and single scattering albedo.

li

li

Page 52: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

The accuracy of the approximation depends on

; that is, the smaller the p value, the

more accurate the approximation is (Huang et al., 2007). It follows from Eqs. (2.7) and

(2.8) that the single scattering albedo and the recollision probability are decreasing

functions of V, i.e., and if . In other

words, the smaller the base scale, the more hierarchical levels of the vegetation in a pixel

are involved, and more accurately the contributions of scattering orders should be

accounted to estimate canopy absorptive and reflective properties (Knyazikhin et al.,

1998b). The base scale therefore should be chosen sufficiently large to minimize the

number of hierarchical levels and to achieve a good accuracy in the first order

approximation.

In the present approach, the structural variables are calculated for a homogeneous

pixel of a single vegetation type only. The base scale represents an individual plant (e.g.,

tree in a woody vegetation class) or a group of plants (e.g., in grasses). When the spatial

resolution of the imagery decreases (i.e., the volume V0 increases), the degree of

vegetation mixing within the pixel increases. It means that different structures can exist at

the base scale and consequently more hierarchical levels of the canopy structure may be

present in the imagery. This directly follows from Eq. (2.8), i.e., ≥

if . Assuming varies continuously with the base scale V and the

resolution V0, an increase in the recollision probability due to increase in V0 can be

compensated by an increase in V such that = where and

lii

lii

Page 53: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

. Thus, the structural parameters can be pre-calculated for a fixed base scale. The

spectral BRFBS, (and solutions to the “S” problem) can be adjusted for the resolution by

using the single scattering albedo at a scale . The single scattering albedo

therefore allows us to scale up the simulated BRF to a coarser resolution.

2.5 Algorithm adjustment for the sensors spectral bandwidth

For a given spectral band, the observed BRF is a weighted integral of the spectral

BRF over a spectral interval, i.e., the bandwidth. The weight is the spectral response

function that describes the sensitivity of the sensor to a particular wavelength in the

spectral interval. Both the weight and the interval are sensor specific and vary with the

spectral band. Figure 2.6 shows spectral response functions for red

(580nm680) and NIR (725nm1100) spectral bands for the NOAA

16 AVHRR sensor (WWW1). The corresponding MODIS spectral bands,

620nm670nm and 841nm876, are much narrower and shapes of

the response functions (WWW2) differ from their AVHHR counterparts (Fig. 2.6). The

difference in the spectral band characteristics is a factor that changes spectral signatures

of pixels measured by two sensors. In the parameterization, the structural and radiometric

components of the measured signal are separated. This feature provides a simple way to

adjust the algorithm for sensor band characteristics. Since the solution of the “BS-

problem” is a major source of information about the intrinsic canopy properties, the focus

is given on this component of the signal.

liii

liii

Page 54: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

The measured reflectance, BRFM,, is a weighted integral of Eq. (2.2) over a spectral

interval , i.e.,

. (2.9)

Here is the mean single scattering albedo; f(), =1, is the

spectral response function, and (p) is the ratio function defined as:

. (2.10)

Note that in the integral term of Eq. (2.10) is a convex function with

respect to values of the single scattering albedo. It follows from the Jensen’s inequality

(Gradshteyn and Ryzhik, 1980) for convex functions that the numerator in Eq. (2.10) is

no less than the denominator, and thus, (p)≥1.

Values of and (p) depend on the variation of the single scattering albedo

with wavelength. If the single scattering albedo is constant in the interval ,

then = and (p)=1 and no adjustment is needed. Such a situation is typical for NIR

spectral bands in which the single scattering albedo is almost flat with respect to

wavelength. The single scattering albedo exhibits much stronger variation at wavelengths

between 580nm and 680nm. In this interval, is a decreasing function with a local

minimum at about 680 nm. The averaging of over the red AVHHR spectral band

results in a higher value of than over the narrower spectral interval of the red MODIS

band. This effect tends to increase the measured AVHRR surface reflectances at red

compared to the corresponding MODIS values.

liv

liv

Page 55: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

The variation of causes the ratio (p) to deviate from unity which enhances the

measured reflectance. Figure 2.7 shows (p) for red and NIR spectral bands for AVHRR

and MODIS. The ratio is an increasing function with respect to the recollision

probability. For the NIR spectral band, (p) is very close to unity, with maximum

deviation being less than 0.5%. For the red spectral band, values of the ratio are higher

and can deviate from unity by 8%. However, the overall variation in the ratio does not

exceed 2%. In this example, the ratio was calculated using a typical single scattering

albedo of an individual leaf and thus its values correspond to the leaf scale. Recall that

the single scattering albedo and the recollision probability are decreasing functions of the

base scale. It follows from Eq. (2.10) that the change in the single scattering albedo by a

factor k alters the ratio from (p) to (kp). The adjustment of the reflectance for a

coarser data resolution, therefore, lowers its variation due to decreases in both the single

scattering albedo (k1) and the recollision probability.

To summarize, the problem of accounting for differences in spectral characteristics

between sensors can be reduced to finding band dependent values of the single scattering

albedo that compensate for changes in due to differences in the bandwidths and

deviation of (p) from unity due to variation in , where the latter is dependent on the

base scale. The single scattering albedo therefore is the basic configurable parameter to

adjust the simulated MODIS BRF for the spatial resolution and spectral band

composition of the AVHRR sensor. Its value can be specified by fitting the simulated

BRF to the observed BRF values over different vegetation types during the green peak

lv

lv

Page 56: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

season (Hu et al., 2003; Shabanov et al., 2005). This technique will be demonstrated in

Chapter 3.

2.6 Data information content and observation uncertainty

The difference in information content of MODIS surface reflectance and AVHRR

NDVI can be quantified as follows. The spectral reflectance of a surface can be depicted

as a point in the red-NIR spectral space. The location of the point in the polar coordinate

system is given by the polar angle, = , and the radius

r . Here RED and NIR represent BRF values at red and NIR spectral

bands. Pixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This

line intersects the origin of the spectral plane at an angle . In the case of MODIS, the

surface reflectance data provide both the angle and location on the line, while the

AVHRR NDVI data provide the angle only.

The MODIS LAI/FPAR algorithm exploits the location information by attributing

each point in the spectral space to a specific physical state that is characterized by a

background brightness and LAI. A pixel can have a background ranging from dark to

bright soils, and the LAI can vary over a range for each specific instance of background

brightness. In order to meet consistency requirements formulated in Section 2.2, a

specific mode in the MODIS algorithm (“AVHRR mode”) is implemented, which accepts

the angle, , and a range (rminrrmax) of valid radii as inputs. This range corresponds

to variations of r for given as given by Collection 5 MODIS LUTs. While executing

the algorithm in the AVHRR mode, the following situations are possible (cf. Fig. 2.8):

lvi

lvi

Page 57: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

If rmin=rmax, the set of acceptable solutions coincides with that generated by the

operational MODIS LAI/FPAR algorithm.

If rmin<rmax, the set of acceptable solutions includes as subset standard MODIS

retrievals (Fig. 2.9).

The following merit function is used to select the set of acceptable solutions in the

MODIS-algorithm,

(2.11)

Here and denote values of measured surface reflectances, while NIR and

RED correspond to values of simulated reflectances. The dispersions and

quantify combined model and observations uncertainties at NIR and red spectral bands

and are configurable parameters in our approach (Wang et al., 2001). The dispersions are

represented as, and , where and are the

corresponding relative uncertainties (Wang et al., 2001). The variable characterizing

the proximity of measured surface reflectances to simulated values has a chi-square

distribution with two degrees of freedom. A value of indicates good proximity

between observations and simulations (Wang et al., 2001). All LAI and soil reflectance

values satisfying this criterion constitute the set of acceptable solutions for a particular

MODIS observation ( and ).

In the AVHRR-mode, the criteria is applied to each point on the line (Fig.

2.8), i.e., for , and .

lvii

lvii

Page 58: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Here, is the input simple ratio from AVHRR, and is the radius

obtained from the MODIS LUT. Note that the AVHRR mode does not increase

computation time since in the MODIS mode the inequality is checked for all

combinations of LAI and soil patterns.

Figure 2.10 shows correlation between LAI retrievals using MODIS and AVHHR

modes of the algorithm. In both modes, the algorithm generates similar mean LAI values

given data from the same instrument. The corresponding dispersions, however, can differ

significantly, indicating varying information content of retrievals. Recently, Hu et al.

(2007) compared MODIS and MISR LAI seasonal profiles retrieved from data which

have the same accuracy but different information content (Fig. 4 in the cited paper). The

use of multi-angle and spectral information allows capturing seasonal LAI variations that

are not detected by single-angle views. The impact of the information content on

retrievals with respect to MODIS and AVHHR retrievals will be detailed in Chapter 3.

2.7 Conclusions

This research introduces a physically based approach for generating LAI and FPAR

ESDRs and its application to developing a long time series of these products from

MODIS and AVHRR data (Chapter 3). In general, ESDR algorithms ingesting data from

different instruments should account for differences in spatial resolution, spectral

characteristics, uncertainties due to atmospheric effects and calibration, information

content, etc. The approach to this problem is based on the radiative transfer theory of

spectral invariants. Accordingly, the canopy spectral BRF is parameterized in terms of a

lviii

lviii

Page 59: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

compact set of parameters - spectrally varying soil reflectances, single-scattering albedo,

spectrally invariant canopy interceptance, recollision probability and the directional

escape probability. This approach ensures energy conservation and allows decoupling the

structural and radiometric components of the BRF. According to this theory, the single

scattering albedo accounts for the dependence of BRF on sensor’s spatial resolution and

spectral bandwidth. The parameter characterizing data uncertainty accounts for variation

in the information content of the remote measurements. Thus, the single scattering albedo

and data uncertainty are two key configurable parameters in our algorithm. The algorithm

supports two modes of operation: the MODIS mode (retrievals from BRF) and the

AVHRR mode (retrievals from NDVI). In both cases, the algorithm simulates similar

mean LAI values, if input data from the same instrument are used. The corresponding

dispersions, however, differ significantly, indicating varying input information content

and related uncertainties (MODIS BRF vs. AVHRR NDVI). Overall, the problem of

generating LAI/FPAR is reduced to the problem of finding values of data uncertainty and

single scattering albedo for which: a) the consistency requirements for retrievals from

MODIS and AVHRR are met; b) the difference between MODIS and AVHRR

LAI/FPAR is minimized; c) the probability of retrieving LAI/FPAR is maximized. The

implementation of this algorithm and evaluation of the derived product will be detailed in

the next chapter.

lix

lix

Page 60: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 2.1 Schematic plot of the photon-canopy interactions. Photons incident on the vegetation canopy will be intercepted by phytoelements with probability i0. This probability of initial collision, or canopy interceptance, does not depend on wavelength and is a function of the direction of incident beam and canopy structure. The intercepted photons will be scattered by the foliage elements with probability , and, in turn, will either interact again or escape the canopy with probabilities p and , respectively. Given pm and m as a function of scattering order m, the probability that photons from the

incident beam will escape the vegetation after m interactions is 0121 )( ippp mm

m .

The probability of absorption after m interactions is 0121 )()1( ippp mm

. The proportion of absorbed or exiting photons is equal to the sum of corresponding probabilities for scattering order m.

lx

lx

Page 61: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

1

Figure 2.2 Canopy interceptances i0 and i0,S as a function of LAI. The latter is the proportion of photons from isotropic sources below the canopy that is intercepted by the vegetation. Calculations were performed for a vegetation canopy consisting of identical cylindrical “trees” uniformly distributed in the canopy layer bounded from below by both a non-reflecting (black soil problem) and reflecting (soil problem) surface. The canopy structure is parameterized in terms of the leaf area index of an individual tree, Lo, ground cover, g, crown height, H, and crown diameter D. The LAI varies with the ground cover as LAI=gLo. The stochastic radiative transfer equation was used to derive canopy spectral interaction coefficient i() for both reflecting and non-reflecting surfaces. The interceptances i0 and i0,S are obtained by fitting the spectral invariant approximation to i(). The crown diameter, height and plant LAI are set to 1(in relative units), 2 and 10, respectively. The solar zenith angle and azimuth of the incident beam are 30o and 0o. The view zenith angle is nadir.

Page 62: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

2

Figure 2.3 Recollision probabilities calculated by fitting equations for the black soil (BRFBS, aBS and tBS) and “S” (aS, rS, and tS) problems to their simulated values. Here pBS

and pS are specified by fitting aBS and aS; pt, pJ and pS by fitting tBS, JS and tS; pr and prs by fitting BRFBS and rS. Calculations were performed for the 3D vegetation canopy described in Fig. 2.2. The crown diameter, height and plant LAI are set to 1(in relative units), 2 and 10, respectively. The solar zenith angle and azimuth of the incident beam are 30o and 0o. The view zenith angle is nadir.

Page 63: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

3

Figure 2.4 Values of the energy conservation relationships rBS+tBS+aBS as a function of single scattering albedo and LAI. Calculations were performed for the 3D vegetation canopy described in Fig. 2.2. Parameters rBS, tBS and aBS are obtained from the spectral invariant approximation to the reflectance, transmittance and absorption values as derived from the stochastic radiative transfer simulations.

Page 64: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

4

0 0.05 0.1 0.15 0.2 0.25 0.3 0.350

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Simulated BRF at Red

Sim

ulat

ed B

RF

at N

IR

MODIS-LUTModel-LUT

soil line

nir =0.84, red =0.14

Figure 2.5 The spectral invariant approximation to the BRF superimposed on the MODIS LUTs entries for the BRF. The effective ground reflectance patterns for the MODIS LUT are restricted to dark and intermediate brightnesses for illustration purpose, while the spectral invariant simulation includes backgrounds ranging from dark to bright soils. The red-NIR spectral space is displayed for the broadleaf forest vegetation class, characterized by the single scattering albedo’s at the NIR(ωnir=0.84) and red(ωred=0.14) bands. Calculations were performed for the 3D vegetation canopy described in Fig. 2.2. The solar zenith angle and azimuth of the incident beam are 30o and 0o. The view zenith angle is nadir.

Page 65: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

5

Figure 2.6 Relative spectral response function in the red and NIR spectral intervals for the NOAA AVHRR-16 and MODIS sensors.

Page 66: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

6

Figure 2.7 The ratio (p) for red and NIR spectral bands for AVHRR and MODIS. Spectral response functions shown in Fig. 2.6 for AVHRR-16 and MODIS and a typical single scattering albedo of an individual leaf are used to calculate the ratio (“std” in the figure refers to standard deviation of (p)).

0 0.1 0.2 0.3 0.40

0.1

0.2

0.3

0.4

Surface Reflectance at Red

Surf

ace

Ref

lect

ance

at N

IR

= tan( SM* ) = arctan SM

Page 67: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

7

Figure 2.8 Reflectance of vegetated surface in the red-NIR spectral plane. The cross symbols mark the spectral space of the MODIS LAI/FPAR LUTs for a range of simulated LAI and soil background brightnesses. The line with circles intersects the origin at an angle defining the Simple Ratio (SM). This line depicts different possible combinations of red and near-infrared reflectances corresponding to different LAI values and soil spectral reflectance patterns. The ellipse represents the inequality criterion for which the solution set is obtained.

Page 68: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

8

0 1 2 3 4 5 6 70

20

40

60

80

100

LAI

Tota

l Num

ber o

f Ret

rieva

ls

AVHRR modeMODIS modemean LAIAVHRR modemean LAIMODIS mode

Figure 2.9 Distribution of acceptable LAI values corresponding to the full range of possible values of the radius (squares) and to a specific value of the radius (triangles). The mean LAI values and their dispersions are taken as the LAI retrievals and their uncertainties. The AVHHR and MODIS modes were applied to the MODIS surface reflectance using the MODIS LUT.

Page 69: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

9

0 1 2 3 4 5 60

2

4

6

MODIS mode LAI

AVH

HR

mod

e LA

I For Black Soil

0 1 2 3 4 5 60

2

4

6

MODIS mode LAI

AVH

HR

mod

e LA

I For Medium Soil

0 1 2 3 4 5 60

2

4

6

MODIS mode LAI

AVH

HR

mod

e LA

I For Bright Soil

Figure 2.10 Correlation between LAI values retrieved using the MODIS (horizontal axis) and AVHHR (vertical axis) modes of the algorithm for dark (sur,RED=sur,RED=0.05), medium (sur,RED=sur,RED=0.16) and bright (sur,RED=sur,RED=0.26) backgrounds. Surface reflectances shown in Fig. 2.8 that correspond to selected backgrounds and simple ratio (SM) calculated from the surface reflectances were used as input. The relative uncertainties at red and NIR spectral bands were set to 0.3 and 0.15.

Page 70: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

10

Chapter 3

3. Generating Leaf Area Index Earth System Data Record from

multiple sensors: Implementation, analysis and validation

3.1 Introduction

Long-term global vegetation monitoring requires temporally and spatially

consistent data sets of vegetation biophysical variables characteristic of vegetation

structure and functioning like the Leaf Area Index (LAI) and Fraction of

Photosynthetically Active Radiation (FPAR). Such data sets are useful in many

applications ranging from ecosystem monitoring to modeling of the exchange of energy,

mass (e.g. water and CO2), and momentum between the Earth’s surface and atmosphere

(Dickinson et al., 1986; Sellers et al., 1996; Tian et al., 2004; Demarty et al., 2007). A

key step in assembling these long-term data sets is establishing a link between data from

earlier sensors (e.g. AVHRR) and present/future sensors (e.g. MODIS TERRA,

NPOESS) such that the derived products are independent of sensor characteristics and

represent the reality on the ground both in absolute value and variations in time and space

(Van Leeuwen et al., 2006). Multi-decadal globally validated data sets of LAI and FPAR

produced with a physically based algorithm and of known accuracy are currently not

available, although several recent attempts have resulted in shorter term research quality

data sets from medium resolution sensor data (Knyazikhin et al., 1998; Gobron et al.,

1999; Chen et al., 2002; Yang et al., 2006a; Plummer et al., 2006; Baret et al., 2007).

Page 71: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

11

In Chapter 2, a physically based approach was presented for deriving LAI and

FPAR products from AVHRR data that are of comparable quality to the MODIS

products. The theoretical approach is based on the radiative transfer theory of canopy

spectral invariants which facilitates parameterization of the canopy spectral bidirectional

reflectance factor (BRF). The methodology permits decoupling of the structural and

radiometric components and is applicable to any optical sensor. However, it requires a set

of sensor-specific values of configurable parameters, namely the single scattering albedo

and data uncertainty, in order to maintain consistency in the derived products.

The objective of this chapter is to present a comprehensive evaluation of a new

global monthly LAI data set derived from the AVHRR NDVI for the period July 1981 to

December 2006 with the algorithm presented in Chapter 2. The outline of this chapter is

as follows. The implementation of the algorithm and production of the data set are first

detailed. The data set was evaluated both by direct comparisons to ground data and

indirectly through inter-comparisons with similar data sets. This indirect validation

included comparisons with MODIS and CYCLOPES LAI products at a range of spatial

scales, and correlations with key climate variables in areas where temperature and

precipitation limit plant growth. The data set was also analyzed to reproduce spatio-

temporal trends and inter-annual variations in vegetation activity that have been reported

previously in the literature. Direct validation included comparisons to field data from

several campaigns conducted as part of the Land Product Validation Subgroup (LPV) of

the Committee Earth Observing Satellite (CEOS) (Justice et al., 2000; Morisette et al.,

Page 72: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

12

2006). In the final section, conclusions from the production and evaluation exercises are

presented.

3.2 Production of the LAI dataset

This section provides a schematic of the algorithm implementation as well as a brief

description of the input datasets and LAI retrieval mechanism. The MODIS Collection 5

(C5) LAI product over the period of overlap between the two sensors (2000 to 2002) is

taken as a benchmark in this study. Descriptions related to satellite input data and land

cover classification map are provided in section 3.2.1. A step-by-step implementation of

the theoretical framework (Chapter 2) is given in section 3.2.2 and 3.2.3. The LAI

product is described in section 3.2.4.

3.2.1 Input satellite data and land cover classification map

The 15-day maximum value AVHRR NDVI composites (Holben, B., 1986) from

the NASA GIMMS group for the period July 1981 to December 2006 are used as the

input data in this study (Tucker et al., 2005). The data are at 8 km spatial resolution in a

geographic latitude-longitude projection and have been corrected for loss of calibration,

view and solar zenith angle variations, contamination from volcanic aerosols, and other

effects not related to vegetation change (Tucker et al., 2005; Pinzon et al., 2005). The

maximum value compositing diminishes the atmospheric effects such as sensitivity of the

AVHRR wide spectral bands to the presence of water vapor, ozone, etc (Brown et al.,

2006) as well as minimizes residual atmospheric and cloud contamination (Holben, B.,

Page 73: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

13

1986). The average of the two 15-day maximum value composites was used to generate

the monthly 8 km LAI product.

The latest version (Collection 5) of the MODIS LAI data set was used as a

benchmark in the production of the AVHRR LAI data set (Shabanov et al., 2005; Yang et

al., 2006b). Monthly 1 km MODIS data set was aggregated to 8 km spatial resolution. In

the aggregation process, the LAI value of the 8 km pixel is calculated as the mean over 1

km LAI values of high quality only. The high quality retrievals refer to LAI values

generated by the C5 MODIS LAI/FPAR algorithm (Yang et al., 2006b). Further, the

quality flag value for the 8 km pixel is evaluated as the percentage of the corresponding

sixty four 1 km pixels of high quality. This 8 km MODIS data set was then re-projected

to the geographic latitude-longitude projection from its native ISIN projection. For the

inter-comparison of AVHRR and MODIS LAI (Section 3.3), only 8 km MODIS LAI

pixels with quality flag values greater than 90% were used. Similarly, only main RT

algorithm AVHRR retrievals (Section 3.2.3) were used when inter-comparing the data

sets.

The land cover map (or biome classification map) is a key ancillary input to the LAI

retrieval process. The Collection 5 MODIS LAI/FPAR operational algorithm references a

1 km eight biome classification map consisting of the following classes: (1) grasses and

cereal crops, (2) shrubs, (3) broadleaf crops, (4) savannas, (5) evergreen broadleaf

forests, (6) deciduous broadleaf forests, (7) evergreen needleleaf forests, (8) deciduous

needleleaf forests (Yang et al., 2006a). This map is used in the retrieval process but at 8

Page 74: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

14

km resolution by retaining the most frequently occurring land cover amongst the sixty

four 1 km pixels (Tian et al., 2002).

3.2.2 Achieving consistency with the Terra MODIS LAI products

The AVHRR mode of the proposed algorithm accepts inputs of surface reflectances

as projections of the simple ratio (SM = near-infrared/red reflectances) line onto the red-

NIR spectral axes, i.e., the red and NIR reflectance are proportional to cos and sin,

respectively, where = (Section 2.6 in Chapter 2).

Errors in cos and sin are determined by errors in the NDVI data. The consistency

condition of the multi-sensor LAI algorithm (Section 2.2 in Chapter 2) ensures that the

difference between mean LAI values from the AVHRR (using NDVI as input) and

MODIS (using spectral reflectance as input) modes of the algorithm is minimized.

Specifically, the adjustment procedure was reduced to finding values of sensor-specific

input and model uncertainty and single scattering albedo for which

a) The consistency conditions are met;

b) The retrieval index (RI) is maximized;

c) The difference (RMSE) between AVHRR and MODIS LAI is minimized.

The retrieval index is the ratio of the number of pixels for which the algorithm retrieves a

value of LAI to the total number of processed pixels,

(3.1)

Page 75: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

15

The RI is a function of uncertainties in modeled and observed reflectances (or, the simple

ratio) and the total number of spectral bands used. In general, the RI increases with

increasing values of uncertainties (Figs. 3.1a and 3.1b) but at the same time, higher

values of uncertainty refer to poor quality input data, and thus, poor quality retrievals.

The RMSE is defined as the root mean square error between the output AVHRR

LAI image and the corresponding MODIS LAI image,

. (3.2)

Here LAIAVHRR represents LAI values generated by the AVHRR mode of the algorithm

given the relative uncertainty and the single scattering albedo , and LAIMODIS denotes

the aggregated MODIS LAIs. The summation is performed over all 8 km vegetated

pixels (k) in a given image. The RMSE is a function of the relative uncertainty, single

scattering albedo, biome type and the image size. The following procedure was

implemented to achieve efficient production of AVHRR LAI on a global scale.

The configurable parameters, relative uncertainty ( ) and single scattering albedo

() at the red, are varied to minimize RMSE and maximize RI. Since uncertainties in the

NIR reflectance are minimally impacted by the differences in spectral bandwidth and data

resolution (Miura et al., 2000; Van Leeuwen et al., 2006; Section 2.5 in Chapter 2), their

values are taken from the corresponding MODIS LUT and kept constant in the tuning

procedure. This process is illustrated in Fig. 3.1. The RI increases with increasing values

of . A value of of 30 % generally yields a RI value of 90% or more.

values greater than 30% are usually not considered as they represent poor quality input

Page 76: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

16

data and thus unreliable LAI retrievals. Also, for a preset value of , the RMSE is

minimum for a specific value of red (red = 0.25 in Fig. 3.1a and red = 0.22 in Fig. 3.1b).

Thus, for each vegetation type, AVHRR LUT can be created consisting of the spectral

BRF values calculated with optimum values of the single scattering albedo. Optimal

values of the relative uncertainties are then used to specify the merit function (see Eq.

2.11 in Chapter 2).

An extensive analyses of the above procedure for all months of the year 2001,

biome-by-biome, is performed in order to examine variations in the configurable

parameters with biome type and month. The results suggest that the optimal values for

the configurable parameters exhibit a non-negligible variation with respect to the biome

type and a weak sensitivity to the month. For each biome type, mean values over months

are taken as instrument specific configurable parameters and used to generate global

AVHRR LAI time series.

3.2.3 Global LAI production

Consider an input pixel from the AVHRR NDVI map representing a certain

vegetation class. The algorithm queries the corresponding AVHRR LUT for this class

and calculates the mean value and standard deviation (dispersion) of LAI from the

retrieved solution set (Chapter 2). A successful retrieval is classified as a main algorithm

retrieval. If the retrieval is unsuccessful, a backup algorithm similar to the MODIS

approach is adopted, where the input simple ratio will be used to calculate a LAI value

Page 77: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

17

based on NDVI-LAI empirical relations (Yang et al., 2006a). A schematic representation

of the algorithm implementation is shown in Fig. 3.2.

3.2.4 LAI data dissemination

The data set is stored as monthly LAI values together with their standard deviation

and quality control flags at 8 km resolution in a geographic latitude-longitude projection

for the period July 1981 to December 2006. The accuracy of the AVHRR LAI product is

comparable to the MODIS LAI (cf. later sections), but with increased dispersion, which

reflects the lower information content of AVHRR input (NDVI) compared to the MODIS

input (spectral reflectances).

3.3 Comparison with other satellite LAI products

3.3.1 Assessment of AVHRR LAI for the tuning year

A global scale assessment of the AVHRR LAI data set was performed for the

overlapping year 2001 for which MODIS LAI was used as a reference for tuning the

AVHRR algorithm. The quality of the AVHRR product depends on the quality of input

NDVI and land cover data. The input land cover map is the same for both MODIS and

AVHRR, therefore the input NDVI/reflectance data determine the LAI product quality. A

pixel data is considered reliable only if the MODIS quality flag for the corresponding

pixel corresponds to a best quality retrieval and simultaneously has an AVHRR radiative

transfer (RT) algorithm retrieval. For each biome, all such pixels are selected for analysis

Page 78: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

18

over the course of a year. A month by month spatial difference, defined as delta LAI

(LAI) for each such pixel over a particular biome type is calculated as

(3.3)

Here ib, jb are pixel coordinates for a specific biome “b” and “t” is the month. Table 3.1

shows the accuracy ( = mean value of ) and precision ( = standard

deviation of ) of AVHRR LAI with respect to MODIS LAI for different

biomes and four different months. The last row in Table 3.1 shows calculated from

annual maximum LAI values, which reflect the worst possible case.

The herbaceous biomes (grasses/cereal crops, shrubs, broadleaf crops, savannas)

show a in the range of zero to 0.25 LAI in all the four months, while the woody

biomes (broadleaf and needleleaf forests) show a from nearly zero to 0.42. AVHRR

LAI underestimates MODIS LAI especially in the savannas and needleleaf forests. The

accuracy of annual maximum LAI is usually within 0.6 LAI for most of the vegetation

classes. The precision is always within 0.65 standard deviation units. In deciduous

broadleaf forests, peak annual AVHRR LAI overestimates its MODIS equivalent by

almost 0.61 LAI. This is due to uncertainty in input NDVI and differences in the time of

the year at which the peak LAI value exists in these products. These larger differences

suggest that LAI retrievals from NDVI (low information content) will poorly capture the

seasonality in comparison to LAI retrievals from surface reflectances (higher information

content). The input information content and uncertainties control whether the retrievals

are over- or under-estimates, which in this case is not systematic as shown in Fig. 3.3.

Page 79: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

19

Pixels with greater than 0.6 LAI difference constitute only 3.3 % of the total global

vegetated pixels. Overall, the difference values indicate spatio-temporal consistency

between the AVHRR and MODIS LAI data sets as well as acceptable levels of accuracy

and precision, suggesting that the tuning process has been successfully implemented.

3.3.2 Comparison with MODIS C5 LAI data

The inter-comparison of AVHRR and MODIS data sets for a three year period

(2000 to 2002) is presented here. The choice of these years is dictated by the availability

of latest version of MODIS products (Collection 5, or C5).

3.3.2.1 Global scale comparison

Figure 3.4a shows a comparison between the mean monthly AVHRR and MODIS

LAI for each of the 8 vegetation classes. The AVHRR values explain 97.5% of the

variability in MODIS values and on average will be in error in their estimation by 0.18

LAI. This comparison indicates qualitative agreement between the two data sets;

however, global averaging over all pixels of a particular vegetation class masks the

underlying variability in LAI amongst those pixels. Therefore, an inter-comparison at

regional and local scales was performed, as reported below.

3.3.2.2 Regional scale comparison

Grasslands and evergreen needleleaf forests were considered for a regional scale

inter-comparison exercise. Mean seasonal values – DJF (December to February), MAM

(March to May), JJA (June to August), and SON (September to November) - for a

homogeneous region of 100 by 100 evergreen needleleaf forest pixels in Western Russia

Page 80: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

20

were calculated for the period 2000 to 2002 and shown in Fig. 3.4b. The R2 (0.895) and

RMSE (0.25) indicate good agreement between the two data sets. The dispersion does not

exceed 0.5 LAI. Figure 3.4c is a similar plot for three homogeneous grassland sites,

100x100 pixels each, located in the Great Plains of the USA, the Sahel and Central Asia.

Again, the R2 (0.84) and RMSE (0.12) indicate a satisfactory agreement between the two

data sets. Some scatter at LAI values greater than 1.25 is seen which is within the

accuracy limit of 0.5 LAI.

3.3.2.3 Local scale comparison

The June to August mean LAI of a homogeneous patch of 50 by 50 evergreen

needleleaf forest pixels in western Russia was used for local scale inter-comparison

between the two data sets. Figure 3.4d shows the mean and dispersion of AVHRR LAI as

a function of MODIS LAI. The regression relation in Fig. 3.4d has R2 of 0.94 and RMSE

of 0.32 which indicate good correspondence between the two data sets, although some

dispersion can be seen at LAI values greater than 3.

3.3.3 Comparison with CYCLOPES LAI product

The CYCLOPES LAI product (version 3.1) was derived from data from the

SPOT/VEGETATION sensor over a 1/112o resolution plate-carrée spatial grid and 10-

day frequency (Baret et al., 2007). CYLOPES products from a select BELMANIP

benchmark network of sites (Baret et al., 2006) are utilized here (Table B1, Appendix B).

The land surface type of each BELMANIP site is defined using the ECOCLIMAP

classification (Masson et al., 2003), which classifies the land into seven main categories.

Page 81: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

21

This inter-comparison was done for four vegetation types – needleleaf forests, grasslands,

savannas and croplands. For each vegetation type, we chose representative sites, each 50

x 50 km2 (about 7 x 7 AVHRR pixels) in area, and calculated the mean monthly LAI

values for the period 2001 to 2003 using simultaneously available AVHRR and

CYCLOPES pixels that cover the same areas (Fig. 3.5). Mean values accumulated over

areas composed of 20 or more AVHRR pixels were used in our inter-comparison

analyses.

In grasslands and needleleaf forests, the AVHRR and CYCLOPES LAI values are

close to 1:1 line – slopes are 0.85 and 0.88, respectively and corresponding offsets are

0.17 and -0.05 (Figs. 3.5a and 3.5b). They explain 80% and 68% of the variability in

CYCLOPES LAI and will be in error by 0.32 and 0.47 LAI on average. The correlation

between the two products is also very strong in the case of croplands (R2=0.77) and

savannas (R2=0.85); they differ by 0.34 LAI (croplands) and 0.24 LAI (savannas) (Figs.

3.5c and 3.5d). The deviation of AVHRR-CYCLOPES relationships from the 1:1 line is

larger compared to grasslands and needleleaf forests (Figs. 3.5a and 3.5b). The deviations

can partly be explained by a narrow dynamic range of LAI values which is comparable to

variation due to observation errors (Wang et al., 2001; Tan et al., 2005; Huang et al.,

2006). In this example, mean (standard deviation) values of CYCLOPES LAI over

croplands and savannas are 0.87 (0.5) and 0.52 (0.5), respectively. Corresponding values

of the AVHRR LAI are 0.73 (0.2) and 0.56 (0.32). Although there is variation in the

comparability between the two data sets not only across vegetation types but also across

Page 82: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

22

sites with in a vegetation class, the disagreement does not exceed the disagreement

between AVHHR LAI and ground truth data (Section 3.5).

3.4 Comparison with climate variables

In the absence of long-term field LAI measurements spanning a couple of decades,

another way to evaluate the data set is to compare temporal LAI trends with well

documented trends in vegetation growth and their relationship to temperature and

precipitation anomalies in areas where these climatic variables limit plant growth

(Buermann et al., 2003). The ability of the LAI data set to track vegetation changes due

to surface temperature variations in the northern latitudes and precipitation changes in the

semi-arid regions is evaluated in sections 3.4.1 and 3.4.2. The well correlated modes of

co-variability between temperature, precipitation and LAI are isolated, and the

relationship to large-scale circulation anomalies associated with the El Niño-Southern

Oscillation (ENSO) and Arctic Oscillation (AO) are assessed in section 3.4.3.

3.4.1 LAI variation with surface temperature in the northern latitudes

The northern latitudes, 40o to 70o N, have witnessed a persistent increase in growing

season vegetation greenness related to unprecedented surface warming during the period

1981 to 1999 (Myneni et al., 1997; Zhou et al., 2001; Slayback et al., 2003). This

greening was observed in Eurasia and less prominently in North America (Zhou et al.,

2001). In fact, a decline in greenness was observed in parts of Alaska, boreal Canada and

northeastern Eurasia (Barber et al., 2000; Goetz et al., 2005). The slightly longer LAI

Page 83: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

23

data set thus facilitates a re-assessment of these changes. The spatial trends (in %) in

growing season, April to October, LAI for the region 40o to 70o N, are calculated for the

periods 1982 to 1999 and 1982 to 2006. The greening trend (Fig. 3.6a) is evident in

Eurasia, Northern Alaska, Canada and parts of North America, for the period 1982 to

1999. When this analysis is extended to 2006 (Fig. 3.6b), it is found that large contiguous

areas in North America, Northern Eurasia and Southern Alaska show a decreasing trend

in growing season LAI. This browning trend, especially in the boreal forests of Southern

Alaska, Canada and in the interior forests of Russia has also been reported in recent

studies (Angert et al., 2005; Goetz et al., 2005).

The spatial (40o-70o N) and growing season (April to October) averages of

standardized anomalies (anomalies normalized by their standard deviation) of LAI,

NDVI and surface temperature (Hansen et al., 1999) are shown in Fig. 3.7 for tundra and

needleleaf forests, separately for North America and Eurasia. The anomaly of a given

variable is defined as the difference between its growing season (April to October) mean

at a given year and the growing season mean over the 1982 to 2006 time interval. The

standardized anomalies of LAI and NDVI track each other very well (Table 3.2), that is,

the long-term trends in the LAI product are not an artifact of the data set or the algorithm.

The results indicate that vegetation activity significantly correlates with trends in

surface temperature in the Eurasian and North American tundra over the entire period of

the record (Table 3.2). This is consistent with reports of persistent greening in the tundra

and evidence of shrub expansion in northern Alaska and the pan-Arctic (Goetz et al.,

2005; Tape et al., 2006).

Page 84: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

24

A decreasing trend in vegetation greenness is observed after 1996-97 period despite

a continuing warming trend in the North American needleleaf forests. The regression

model of LAI vs. surface temperature and time is statistically significant at the 10% level

for the period 1982 to 1999 but is statistically insignificant for the period 1982 to 2006

(Table 3.2). Similar patterns are observed in the Eurasian needleleaf forests also. These

results imply a decreasing trend in vegetation activity possibly due to warming induced

drought stress as has been suggested previously (Barber et al., 2000; Wilmking et al.,

2004; Lapenis et al., 2005; Bunn et al., 2006). There also have been reports of declining

growth and health of white spruce trees in Alaska, upsurge in insect disturbance in

southern Alaska, and increase in fire frequency and severity in Alaska, Canada and

Siberia during the past 6 to 7 years of consistent warming (Soja et al., 2007). These

changes buttress the need for continued monitoring of vegetation activity in these

northerly regions in the face of unprecedented climatic changes.

3.4.2 LAI variation with precipitation in the semi-arid tropics

The availability of water critically limits plant growth in the semi-arid tropical

regions of the world, especially in grasslands where precipitation received in the wet

months is the primary driver of plant growth (Nemani et al., 2003; Hickler et al., 2005;

Prince et al., 2007). This relationship provides a basis for evaluating the LAI product by

examining the correlation between LAI and precipitation (Mitchell et al., 2003; Huffman

et al., 2007). NDVI is also included in this analysis to argue that correlations observed

Page 85: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

25

between LAI and precipitation are not an artifact of the LAI algorithm if they are seen in

the NDVI data as well.

For the purposes of this analysis, the semi-arid regions in the tropics and subtropics

are defined as those with peak annual NDVI values in the range 0.12 to 0.55. These

regions approximately correspond to areas with annual total rainfall less than 700 mm.

Four regions were selected in the Eastern hemisphere for this study - Sahel, Southern

Africa, Southeast Asia and Australia. The Sahelian region consisted of Senegal,

Mauritania, Mali, Burkina Faso, Niger, Nigeria, Chad and Sudan; the Southern African

region of Botswana, South Africa and Namibia; and the Southeast Asian region of

Afghanistan, Pakistan and India.

A highly significant correlation between variations in standardized anomalies of

precipitation and annual peak vegetation greenness (LAI or NDVI) is seen in Australia,

Sahel and Southern Africa (Fig. 3.8; Table 3.3). A slightly less stronger correlation

between these variables is observed in Southeast Asia. An increasing trend in

precipitation and greenness is also observed in these semi-arid regions during the period

of this study (1981 to 2006), which is in agreement with several recent reports on

greening and increased precipitation in the Sahelian region (Herrmann et al., 2005;

Hickler et al., 2005; Seaquist et al., 2006). The greenness increase in Southeast Asia

(especially India) is not supported by enhanced precipitation and is therefore likely due to

other factors such as irrigation and fertilizer use, but this needs to be further investigated.

The strength of these correlations imbues confidence in the interannual variations

embedded in the derived LAI product.

Page 86: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

26

3.4.3 Canonical correlation analysis

The correlations observed between LAI and temperature in the northerly regions

and between LAI and precipitation in the semi-arid areas raise a question about the

mechanistic basis for these relations. It has been reported previously that large scale

circulation anomalies, such as the El Niño-Southern Oscillation (ENSO) and Arctic

Oscillation (AO), explain similar correlations, but at the hemispheric scale (Buermann et

al., 2003). The canonical correlation analysis (CCA) is ideally suited for this purpose as it

seeks to estimate dominant and independent modes of co-variability between two sets of

spatio-temporal variables (Barnett and Preisendorfer, 1987; Bjornsson and Venegas,

1997). The variables are linearly transformed into two new sets of uncorrelated variables

called canonical variates, which explain the co-variability between the two original

variables, in a descending order. Thus, most of the co-variability is captured by the first 2

to 3 canonical variates.

For the canonical correlation analysis in the North, each year is denoted as a

variable (1982 to 2006, that is, 25 variables in total) and each pixel as an observation (the

total number of observations is the number of vegetated pixels in the latitudinal zone 45o

N and 65o N). The two sets of variables for CCA are the spring time (March to May) LAI

and surface temperature anomalies at 1o resolution (Buermann et al., 2003). The

anomalies were normalized by their respective standard deviation. Each of the set of 25

(time) variables was transformed to Principal Components (PC’s) using singular value

decomposition. In each case, only the first six PC’s were retained as they explain a large

fraction of the variance in the input set of variables. In CCA, each canonical variate is a

Page 87: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

27

time-series which accounts for a certain fraction of the co-variability between the

variables (PC’s). In this analysis, the first two canonical variates derived from each set of

six PC’s explained about 50% of the co-variability between the two sets of variables.

The September to November (SON) NINO3 index (WWW3) is used to represent

ENSO because the sea surface temperature anomalies then approach peak values during

an ENSO cycle (Dai et al., 1997). Figure 3.9a shows that the correlation between SON

NINO3 index and first canonical variate related to LAI is very low (r = 0.1). The same is

true for the correlation between SON NINO3 index and the first canonical variate related

to temperature anomalies. This is in contrast to a strong correlation reported in Buermann

et al. (2003) for the period 1982 to 1998. This decline in correlation may be due to weak

ENSO activity and/or changes in teleconnection patterns since the 1998-2000 period

(WWW4). The correlation between the Arctic Oscillation (AO) index and the second

canonical variates of both LAI and temperature is reasonably strong (0.45 and 0.61,

respectively; Fig. 3.9b), consistent with the strong correlations reported by Buermann et

al. (2003) for the period 1982 to 1998. Thus, the AO seems to continue to be a prominent

driver of surface temperature (Thompson and Wallace, 1998) and plant growth variability

in the northern latitudes.

CCA was also performed on standardized anomalies of annual maximum LAI and

precipitation for the semi-arid regions of 40oN to 40oS latitudinal zone (cf. Section 3.4.2).

The first two canonical variates explained about 50% of the co-variability between annual

peak LAI and precipitation anomalies. A reasonable correlation is seen between the

September to November NINO3 index and the first canonical variates of LAI and

Page 88: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

28

precipitation (0.33 and 0.32, respectively; Fig. 3.9c), consistent with several previous

reports of ENSO influence on interannual variability in tropical and sub-tropical

precipitation (Ropelewski and Halpert, 1987; Dai and Wigley, 2000). The correlation

between the second canonical variates and the AO index is weak (Fig. 3.9d) which is not

surprising as the AO is not known to be a driver of precipitation and thus plant growth

variability in these regions.

In summary, the strong ENSO driven linked variations between northern vegetation

greenness and surface temperature observed during the 1980s and 90s have weakened

since 2000. The AO influence however continues to be strong. In the tropical and sub-

tropical regions, the ENSO influence on linked variations between semi-arid vegetation

greenness and precipitation continues to be apparent. These results further add confidence

in the LAI data set as we are not only able to reproduce previously reported results but

also update them.

3.5 Validation with ground data

The validation of coarse resolution satellite products with ground measurements is a

complicated task for several reasons - scaling of plot level measurements to sensor

resolution, geo-location uncertainties, limited temporal and spatial sampling of ground

data, field instrument calibration, sampling errors, etc. (Buermann et al., 2002; Tan et al.,

2005; Huang et al., 2006; Yang et al., 2006a ; Weiss et al., 2007). In an indirect

validation approach, such as inter-comparison of satellite products, the effects of co-

registration and differences in spatial resolution can be minimized if the satellite products

Page 89: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

29

are compared at a coarser resolution (Weiss et al., 2007; Tarnavsky et al., 2008). In a

direct validation such as with field measurements, spatial re-sampling is not always

feasible because of inadequate sampling on the ground. The MODIS experience of LAI

validation will be used to evaluate the AVHRR LAI data set. This work was performed

within the purview of the Land Product Validation (LPV) LAI subgroup of the

Committee Earth Observing Satellites’ Working Group on Calibration and Validation

(CEOS WGCV). Currently, a large number of US and international investigators

participate in this activity by sharing field measured LAI/FPAR data as well as high

resolution maps though the ORNL DAAC Mercury system (Justice et al., 2000; Morisette

et al., 2006). This large and growing data base provides the most up-to-date and

comprehensive information needed for validation of LAI products. The scarcity of field

LAI measurements during the 1980s and 90s represents a more challenging problem.

Nevertheless, we attempted to utilize the available field data from multiple campaigns

(Table B2, Appendix B) and high resolution LAI maps (Table 3.4) to validate the

AVHRR LAI product.

3.5.1 Validation with field/plot-level observations

Most field measurements are typically several plot level samples within a small

homogeneous region representing a certain vegetation type. An ideal field sampling of

LAI must adequately represent its spatial distribution and cover the natural dynamic

range within each major land cover type at the site (Yang et al., 2006b). On the other

hand, the satellite retrievals cover an entire region of interest, but at a coarse scale. If the

Page 90: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

30

sampling in both cases is adequate, the LAI distributions from field measurements and

satellite retrievals should ideally converge to the true intrinsic distribution of the

vegetation class in a given region at a given time (Buermann et al., 2002). In this study,

the sites were selected with adequate plot level measurements over homogeneous patches

of a vegetation type. A pixel-by-pixel comparison between AVHRR LAI and reference

field LAI values is not feasible for at least three reasons – first, the actual spatial location

of the corresponding pixels in the two LAI maps may not match because of geolocation

uncertainties and pixel-shift errors due to the point spread function (Tan et al., 2006);

second, the AVHRR LAI algorithm generates a mean LAI value from all possible

solutions corresponding to possible variation in input due to observation and model

uncertainties (Knyazikhin et al., 1998a). Therefore, the retrieved LAI value for a single 8

km pixel can differ from its measured counterpart, but the mean LAI of multiple pixels

over a homogeneous patch may be valid (Wang et al., 2004); third, the spatial sampling

for a particular field site can aggregate to an area which can be less than or equal to a

single 8 x 8 km2 AVHRR LAI pixel.

Mean LAI values from the 44 field measurements (28 sites) listed in Table B2 in

Appendix B were used in this analysis. Monthly AVHRR LAI values from nearby pixels

with the same vegetation type were averaged. This averaging window, centered on the

field site, was typically about 2 by 2 or 4 by 4 pixels depending on the aerial extent of the

field site. The results of comparison for all the six major vegetation classes indicate that

the AVHRR product underestimates field LAI by about 7% for LAI greater than 1 (Fig.

3.10a). Additionally, the temporal profile of LAI was compared over three sites, Konza

Page 91: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

31

(grasslands), Mongu (savannas), and Harvard forest (broadleaf deciduous forest), for

which there are more than one field value over the course of one or more years. Figure

3.10b shows that the seasonal dynamics of in situ LAI over the Mongu site are well

captured by the AVHRR LAI product, but are underestimates by 0.01-0.4 LAI. Similarly,

the AVHRR LAI seems to capture values of in situ LAI at the other two sites (Figs. 3.10a

and 3.10c), but here the field samples are limited to a single measurement per year.

3.5.2 Validation with fine resolution LAI maps

Another approach to validation involves generation of fine resolution LAI maps

from ground measurements and high-resolution satellite imagery such as ETM+, SPOT,

ASTER, etc. using the so-called transfer function (Yang et al., 2006b). The transfer

function could be empirical methods (Chen et al., 2002; Weiss et al., 2002), physical

models (Tan et al., 2005), or hybrid approaches (Weiss et al., 2002). Currently, several

fine resolution maps are being disseminated via the ORNL DAAC Mercury system

(Morisette et al., 2006). Maps from ten sites representing large homogeneous patches of

distinct land cover types were used in this analysis (Table 3.4). The typical aerial extent

of these maps is between 50 km2 to 100 km2, except for the Canada Center for Remote

Sensing sites (576 km2 ~ 70000 km2). Leaf area index distributions derived from blocks

of 5 by 5 to 10 by 10 eight km AVHRR pixels were compared to LAI distributions from

the fine resolution maps (Fig. 3.11). The AVHRR blocks were larger (30 x 30) for the

Manitoba Black Spruce forest site at Thompson and the needleleaf forest site at Watson

Lake.

Page 92: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

32

Table 3.4 shows the mean values () and corresponding dispersions () for the

distributions displayed in Fig. 3.11. The AVHRR LAI and fine resolution LAI

distributions generally compare well, although there is a consistent underestimation at

most sites possibly related to scale and land cover heterogeneity (Tian et al., 2002). The

distributions agree remarkably well for the Canadian sites, with of

0.46 LAI for the Kejimikujik conifer site and 0.14 LAI for the Manitoba Black Spruce

forest at Thompson. The distributions also compare well for the Alpilles (grasses/cereal

crops) and Ruokolahti (needleleaf forest) sites, with the mean values differing by about

0.14 and 0.36 LAI, respectively. As for the Konza, Harvard forest and Larose sites, the

mean values differ by about 0.11 to 0.5 LAI. However, the AVHRR values for the

broadleaf site at Tapajos site in the Amazon underestimate the actual LAI values

significantly, possibly due to poor NDVI quality as a result of persistent cloud cover and

water vapor contamination. In general, the AVHRR LAI values compare well to field

observations at select sites representative of grasses, crops, broadleaf forests, and

needleleaf forests.

3.6 Conclusions

In this chapter, the evaluation of a new global monthly AVHRR LAI data set for the

period July 1981 to December 2006, derived from AVHRR NDVI is presented. The

production of long term LAI data sets involves a host of inter-sensor related issues like

differences in spatial resolution, spectral characteristics, uncertainties due to atmospheric

effects and calibration, information content, etc. In Chapter 2, a physically based

Page 93: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

33

algorithm for the retrieval of LAI from AVHRR NDVI was introduced. The theoretical

approach is based on the radiative transfer theory of spectral invariants and describes in

detail the physical constraints as well as the conditions required to generate LAI fields of

quality comparable to MODIS LAI products.

The theme of this chapter is establishing the validity and accuracy of the derived

LAI product. The evaluation of the data set is done both through direct comparisons to

ground data and indirectly through inter-comparisons with similar data sets. This

included comparisons with existing LAI products (MODIS and CYCLOPES LAI

products for the 2000 to 2003 period of overlap) at a range of spatial scales, and

correlations with key climate variables in areas where temperature and precipitation limit

plant growth. There is an overall qualitative agreement between the AVHRR and MODIS

data sets at scales ranging from global to regional to pixel. At the global scale, the

AVHRR values explain 97.5% of the variability in the MODIS product and will be in

error in their estimation by 0.18 LAI, on average. The regional and pixel-scale inter-

comparison suggests an average error of less than 0.3 LAI. Comparison with

CYCLOPES LAI indicates satisfactory agreement in most of the biomes with the RMSE

below 0.5 LAI. The data set was also analyzed to reproduce well-documented spatio-

temporal trends and inter-annual variations in vegetation activity in the northern latitudes,

where temperature limits plant growth, and semi-arid areas, where precipitation limits

plant growth. Additionally, to assess the mechanistic basis behind the observed

correlations between LAI and temperature in the northern latitudes and LAI and

precipitation in the semi-arid tropics, a multivariate data-reduction technique (canonical

Page 94: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

34

correlation analysis) was used to isolate well correlated modes of spatio-temporal

variability between LAI and the climate variables. The isolated modes suggest El Niño-

Southern Oscillation and Arctic Oscillation as key drivers of linked interannual variations

in vegetation greenness and precipitation in the semi-arid regions and, vegetation

greenness and surface temperature in the northern latitudes, respectively.

Finally, the LAI data were compared to field measurements and high-resolution

LAI maps from a host of sites. The comparison with plot scale measurements over biome

specific homogeneous patches indicates a 7% underestimation in the AVHRR LAI when

all major vegetation types are considered. The error in mean values obtained from

distributions of AVHRR LAI and high-resolution field LAI maps for different biomes is

within 0.6 LAI with the exception of a broadleaf evergreen forest site. These validation

exercises though limited by the amount of field data, and thus less than comprehensive,

nevertheless indicate satisfactory comparability between the LAI product and field

measurements. In summary, the inter-comparison with other short-term LAI data sets,

evaluation of long-term trends with known variations in climate variables, and validation

with field measurements together build confidence in the utility of this new 26 year LAI

record for long-term vegetation monitoring and modeling studies.

Page 95: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

35

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0

0.5

1.0

1.5

2.0

2.5

Single Scattering Albedo (red)

RM

SE (L

AI

Diff

eren

ce)

0.98

0.99

1.00

Ret

rieva

l Ind

ex (R

I)

= 0.1

= 0.15

= 0.2

= 0.25

= 0.3

= 0.35

= 0.4

0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Single Scattering Albedo (red)

RM

SE (L

AI D

iffer

ence

)

0.1 0.15 0.2 0.25 0.30

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

Ret

rieva

l Ind

ex

=0.35

=0.3

=0.25

=0.2

=0.15

Figure 3.1 The difference (RMSE) between MODIS and AVHRR LAI values (vertical axis on the left side and solid lines) and the Retrieval Index (vertical axis on the right side and dashed lines) as a function of the relative uncertainty =RED/RED and single scattering albedo at red spectral band. Optimal values of the single scattering albedo and the uncertainty should minimize RMSE and maximize the retrieval index. Simultaneously available AVHRR NDVI and MODIS C5 LAI data sets over grasses and cereal crops (panel (a), MODIS tile h10v05) and broadleaf deciduous forests (panel (b), MODIS tile h12v04) for the year 2001 are used in this example.

(a)

(b)

Page 96: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

36

Figu

re 3

.2 F

low

char

t sho

win

g gl

obal

LA

I pro

duct

ion.

Page 97: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

37

Figure 3.3 Average annual LAI difference (AVHRR minus MODIS) for the year 2001. The grey areas represent changes within 0.6 LAI units (96.7 % of the total vegetated pixels). Areas with black color represent water bodies and white color denotes non-vegetated pixels or snow. Land pixels with LAI differences greater than 0.6 represent only 3.3 % of the total vegetated pixels globally.

Page 98: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

38

Figure 3.4 Panel (a) shows comparison between MODIS and AVHRR LAI for the year 2001 (blue color) and 2002 (red color) for different vegetation classes. The LAI values are globally averaged values for the respective vegetation pixels. A similar comparison between MODIS and AVHRR LAI for years 2000-2002 but for different seasons on homogeneous patches of evergreen needle leaf forests in Western Russia (Panel b) and grasslands in the USA, Sahel and Central Asia (Panel c). Panel (d) shows a comparison between MODIS and AVHRR LAI values averaged from June to August, 2001, over a homogeneous patch of evergreen needle-leaf forests in Western Russia. For each MODIS LAI bin, the corresponding AVHRR LAI values were averaged (red box) to reduce scatter. The error bars represent the dispersion of the AVHRR LAI values within each bin.

(a) (b)

(c) (d)

Page 99: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

39

Figure 3.5 Panels (a) to (d) shows comparison between CYCLOPES and AVHRR LAI values for years 2001 to 2003 and different vegetation classes over the BELMANIP site database. The site numbers in the legend are given in Table B1 of Appendix B.

(d)(c)

(a) (b)

Page 100: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

40

Figure 3.6 Trends in AVHRR LAI for the growing season, April to October, for the region 40o N to 70o N, for the periods 1982 to 1999 (panel (a)) and 1982 to 2006 (panel (b)). For each 8 km AVHRR LAI pixel, the April to October mean LAI was regressed on time (years). The slope obtained from this regression, which if statistically significant based on the t-statistic at or lower than 10% level, was converted to a percent trend by multiplying by the number of years times 100 and dividing by the mean April to October AVHRR LAI of 1982.

(a)

(b)

Page 101: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

41

Figure 3.7 Standardized April to October anomalies of AVHRR LAI (green), AVHRR NDVI (blue), and GISS Temperature (red dashed line) for Eurasian and North American needleleaf forests (panels (c) and (d)) and tundra (panels (a) and (b)) from 1982 to 2006.

(a) (b)

(c) (d)

Page 102: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

42

Figure 3.8 Standardized anomalies of annual peak AVHRR LAI (green line), annual peak AVHRR NDVI (blue line) and annual peak (three wettest month CRU+TRMM) precipitation (red dashed line) for the semi-arid regions (panels (a)-(d)) from 1981 to 2006.

(a) (b)

(c) (d)

Page 103: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

43

Figure 3.9 Correlation between standardized time series of the first canonical factor (CF-1, panels (a) and (c)) and second canonical factor (CF-2, panels (b) and (d)) with NINO3 and AO indices in the northern and tropical/subtropical regions.

(a) (b)

(c) (d)

Page 104: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

44

Figure 3.10 Panel (a) shows a comparison of AVHRR LAI with field measurements for the six major vegetation classes. Altogether 44 field data values were used (Table B2 of Appendix B). The AVHRR LAI product is an underestimate by about 7%. Panels (b)-(d) show the temporal profile of AVHRR LAI (blue line) and corresponding field values (red squares) for different vegetation classes. The vertical bars (red and blue) represent the standard deviation associated with the data.

(a)(b)

(c)

(d)

Page 105: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

45

Figure 3.11 Histograms from fine resolution LAI maps (blue color) and AVHRR LAI (red color) over different sites. Information about the sites is given in Table 3.4 together with values of accuracy and precision inferred from these distributions.

Page 106: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

46

Page 107: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

47

Page 108: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

48

Page 109: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

49

Page 110: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

50

Page 111: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

51

Page 112: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

52

Page 113: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Chapter 4

4. Climate and Land Use Impacts on Vegetation in the Semi-arid

Tropics: A Case Study with the LAI Data Set

4.1 Introduction

The semi-arid tropics are projected to be among the most hard hit areas by ongoing

and future climate changes (Parry et al., 2007; Adeel et el., 2007). In these regions,

reductions in vegetation productivity and expansion of desertification are expected to

arise from drier conditions due to continued warming trends accompanied by a reduction

in precipitation (IPCC 2007) and low adaptation capacity of the affected communities

(Parry et al., 2007). Over the past recent decades, tropical dry lands have shown changes

in climate that have resulted in average temperature increases in the range of 0.2-2 oC

(IPCC 2007) and modest but less homogeneous increases in precipitation (Gu et al.,

2007; Zhang et al., 2007), more marked over ocean than over land.

In spite of these climatic changes, which would suggest that tropical dry lands are

already becoming drier, satellite observations of vegetation greenness provide evidence

that, similar to other parts of the globe, and also over extensive portions of the semi-arid

tropics regions, primary productivity has been on the rise (Tucker and Nicholson, 1999;

Eklundh and Olsson, 2003; Pandya et al, 2004; Hermann et al., 2005). The availability of

globally consistent climate datasets has led to investigating quantitatively the climatic

Page 114: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

correlates of the global greening trends (Myneni et al., 1997; Kawabata et al., 2001;

Nemani et al., 2003; Cao et al., 2004). Other driving factors, such as changes in land

cover and land use (Xiao and Moody, 2005), and fertilization effects due to atmospheric

increases in C and N (Ichii et al., 2002), have been invoked more loosely to explain the

remaining portion of the trend and a quantitative analysis of the influence of these factors

on vegetation dynamics is still missing. However, to assist in projecting the impacts of

climate change on ecosystems and societies, it is crucial that the changes in the drivers of

ecosystem dynamics are properly understood.

With the goal of identifying the relative contributions and spatial distribution of

climate, socio-economic and land use change in promoting the greening of the tropical

dry lands, changes in leaf area index (LAI) (derived from GIMMS AVHRR Normalized

Difference Vegetation Index (NDVI)) in conjunction with changes in climatic and land

use data for the period 1981-2006 are analyzed in this chapter. The study focuses on the

semi-arid tropics of the eastern hemisphere, where the largest contiguous dry lands are

inhabited by close to 1.7 billion people (Table 4.1) and spread across 120 countries, most

of which are among the poorest countries in the world and with the lowest human

development index. First, the greening of the semi-arid tropics is compared with changes

in precipitation across all the countries of the eastern hemisphere semi-arid tropics.

Second, the changes in vegetation greenness in the context of changes in socio-economic

and land use change data with a particular focus on India are analyzed, where high-

resolution data are available at the national scale.

Page 115: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

4.2 Data and methods

4.2.1 Normalized Difference Vegetation Index

The maximum value AVHRR NDVI composites (Holben, B., 1986) from the

NASA GIMMS group at the Laboratory for Terrestrial Physics for the period July 1981

to December 2006 is used in this study (Tucker et al., 2005). The GIMMS AVHRR data

are at 8 km spatial resolution and 15-day frequency with a geographic latitude-longitude

projection scheme. This data has been corrected for loss of calibration, view and solar

zenith angle variations, contamination from volcanic aerosols, and other effects not

related to vegetation change (Tucker et al., 2005; Pinzon et al., 2005). The maximum

value compositing usually diminishes the effect of atmospheric artifacts like sensitivity of

the AVHRR wide spectral bands to the presence of water vapor, ozone, etc (Brown et al.,

2006) as well as minimizes differences in the spectral properties, radiometric resolution,

residual atmospheric effects, and, most important, minimizes cloud effects (Holben et al.,

1986). The 8 km GIMMS NDVI has been shown to represent with reasonable precision

interannual variation of photosynthetic activity in global arid and semi-arid areas (Lostch

et al., 2003; Anyamba et al., 2005).

4.2.2 Leaf Area Index

The long term global leaf area index (LAI) data set described in Chapters 2 and 3 is

used in this study. The data span the period July 1981 to December 2006 at monthly

interval. The LAI fields were derived from the maximum NDVI value global composites

Page 116: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

of the AVHRR GIMMS data set, utilizing a physically based algorithm as described in

Chapter 2. Analysis of this LAI dataset (Chapter 3) indicates that temporal trends in the

LAI fields correspond well with existing trends in vegetation growth vis-à-vis

temperature anomalies in the northern latitudes and rainfall anomalies in the semi-arid

tropics (Section 3.4.1 and Section 3.4.2 in Chapter 3). Additionally, well correlated

modes of variability between temperature, precipitation and greenness (LAI) were

isolated and their relationship to large-scale circulation anomalies associated with the El

Niño-Southern Oscillation (ENSO) and the Arctic Oscillation (AO) were assessed

(Section 3.4.3 in Chapter 3). The consistent pattern among the independently derived data

sets further confirm the reliability of the LAI data set in tracking inter-annual vegetation

dynamics with respect to fluctuations in climatic factors.

4.2.3 Precipitation

A 26-yr data set of gridded precipitation was compiled for this study. The data

consists of 0.5o resolution monthly Climate Research Unit (CRU 2.0) data set (Mitchell et

al., 2003), for the period 1981-2000, and, 0.25o resolution monthly Tropical Rainfall

Measuring Mission (TRMM-3B43 v.6) (Huffman, et al., 2007) data set, for the period

2001-2006. The CRU data set has a spatial coverage of -180o(W) to +180o(E) longitudes

and +90o(N) to -90o(S) latitudes. The TRMM data set has a spatial coverage of -180o(W)

to +180o(E) longitudes and +50o(N) to -50o(S) latitudes. The TRMM data set was

degraded to 0.5o to make it spatially compatible with the CRU data set and create a

continuous monthly precipitation record spanning 26 years i.e. 1981-2006.

Page 117: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

4.2.4 Land cover

The Collection 4 (C4) land cover (biome classification map) data (MOD12Q1) and

Collection 3 (C3) Vegetation Continuous Fields (VCF) product derived from MODIS are

utilized in this study. The 1 km land cover type 1 data layer consists of fourteen classes as

referenced by the IGBP classification scheme (Friedl et al., 2002). This land cover map

was re-projected to the geographic latitude-longitude projection from its native ISIN

projection and spatially degraded to a 8 km map by retaining the most frequently

occurring land cover amongst the sixty four 1 km pixels.

The VCF data contains the proportional estimates for vegetation cover types

representing woody vegetation, herbaceous vegetation, and bare ground. The VCF data

are at 500 m spatial resolution in a geographic latitiude-longitude projection and are

available as continental grids (Hansen et al., 2003). The continental grids were mosaicked

to a global grid and the re-sampling to 8 km was performed in a similar manner as the

land cover data.

4.2.4 Ancillary data

National time series on crop production, fertilizer use intensity, and irrigation were

derived from the FaoStat database (http://faostat.fao.org/). Time series on population

growth were extracted from the World Resource Institute Earthtrends database

(http://earthtrends.wri.org/). Sub-national data on irrigated areas for India were obtained

from Narayanamoorthy (2002) and from the Web Based Land Use Statistics Information

Page 118: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

System of the Directorate of Economics and Statistics; Ministry of Agriculture of the

Government of India, New Delhi (http://dacnet.nic.in/lus/).

4.2.5 Methods

Annual maximum NDVI climatology was calculated to define the tropical dry lands

and mask out the remaining pixels within 40oN and 40oS latitudinal band (Fig. 4.1).

Pixels corresponding to tropical dry lands were selected as those with annual maximum

NDVI climatology values in the range of 0.12-0.55, which generally correspond to areas

with climatological precipitation of 700 mm/yr. The most frequent land cover types are

shrublands, grasslands, croplands and, to a lesser extent, savannas (Fig. 4.2a). The

frequency of herbaceous vegetation cover is also dominant in these areas (Fig. 4.2b).

The LAI of semi-arid vegetation tends to fluctuate during the year depending on the

vagaries of rainfall. Long term average LAI values may be expected to be more stable,

unless major shifts in precipitation, land use practices, or a combination of both, affect an

ecosystem. To characterize the spatial distribution of such persistent changes in dry land

vegetation greenness and precipitation over the data record of period, the percentage

change in decadal means of annual maximum LAI and precipitation were calculated.

Here, annual maximum precipitation is defined as the total precipitation of the three

wettest months during a year.

For representative countries in the study area, anomalies of annual maximum LAI

and precipitation were correlated and compared with country-wide decadal changes in

total food production, irrigation area, fertilizer use and macroeconomic indicators. The

Page 119: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

countries included in this study did not have border changes during the 26 year period,

had a considerable portion (at least 40 %) of their surface area within the tropical dry

lands as defined above, and comprised at least 50 half-degree pixels to perform

meaningful comparison with the precipitation data set. Additionally, areas with increased

NDVI and the corresponding changes in net irrigated area were correlated at the state

level to test the hypothesis that changes in land use due to expansion in irrigated areas

have been a major driver of vegetation greenness in India.

4.3 Results and Discussion

4.3.1 Climate-driven increases in LAI

Notable increases in annual maximum LAI is observed between the decade 1981-

1990 and 1995-2006 in over 70% of the tropical dry lands of the eastern hemisphere (Fig.

4.3), encompassing Turkey, large portions of the Middle Eastern countries, the Sahel,

Horn of Africa and Southern African countries, most of tropical Asia and portions of

Australia. About 29% of the area reports decline in photosynthetic activity, principally

distributed in eastern and southern Australia, South-west China, along the Namibian

Desert, and other portions of the coast of Western Africa up to the Iberian Peninsula.

In general, areas that have greened up within the semi-arid tropics show increases in

precipitation over the two decades (Fig. 4.4). Increases in decadal precipitation are

particularly marked along the Sudano-Sahelian semi-arid tropics, the Horn of Africa, the

Middle East and Western Australia. More modest increases are found in most other

Page 120: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

regions and are consistent with findings of increases in tropical land precipitation (Gu et

al., 2007; Zhang et al., 2007), which could be a consequence of the recent warming trends

(Wentz et al., 2007). Reduction in precipitation during the last decade of the order of 20-

40% appear in Egypt, southern Ethiopia and northern Kenya, and especially over

Pakistan, Afghanistan and eastern Australia.

To investigate whether the observed increases in photosynthetic capacity of the

tropical dry lands of the Eastern Hemisphere are connected with local changes in

precipitation amounts, detrended anomalies of annual maximum LAI were correlated

with detrended anomalies of precipitation for the three wettest months in each major

country included in the study area (Fig. 4.5, Table 4.1). For the countries of the Sahel,

Sudan, Somalia, South Africa, Botswana and Australia, significant (p<0.05) positive

linear correlations between the two variables are observed, supporting the hypothesis that

changes in climate that brought increased rainfall especially since the early 1990’s over

most of the subtropical semi-arid countries have helped promoting plant growth in these

dryland regions. Trends in the Palmer Drought Severity Index (Dai et al., 2004), which

integrates the atmospheric moisture with the evaporative demand of the vegetation, also

point to an increased moisture availability of the tropical dry lands, and therefore

enhanced vegetation growth. A recovery of total annual precipitation to the pre-1960’s

levels and consequent greening trends over the Sahel have been particularly well

described by Tucker and Nicholson (1999) and Eklundh and Olsson (2003). Precipitation

in Turkey, Saudi Arabia and India has been modestly on the rise; however, it does not

correlate with the greening. Finally, LAI has markedly increased even in countries

Page 121: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

(Afghanistan, Egypt, Oman, Pakistan) where precipitation has declined in the recent

decades, suggesting that land use change has been more important than climate in

promoting vegetation growth in these regions.

4.3.2 Land use change-driven increases in LAI

All of the major countries in the study area, with the exception of Somalia, reported

an increase in total food production over the two decades of our study period (FAOstat,

2007), indicating that land management and land use changes may have contributed to

the observed greening, especially where this is not supported by the changes in

precipitation. Along with increased precipitation, changes in land use such as transition

from rainfed to irrigated agriculture and increases in consumption of mineral fertilizers,

and probably other, less widely documented, improvements in agricultural practices and

natural resource management (Tappan and McGahuey, 2007; Reij et al., 2005; Niemejer

and Mazzucato, 2002) also likely to have contributed to the increases in photosynthetic

activity recorded by the satellite data.

The role of land use changes, helped by even modest changes in climate, in

promoting large scale increases in plant growth is particularly evident in India, where

52% of the country’s land area is devoted to croplands (FAOStat 2007). While monthly

average temperatures have been on the rise in India, monthly precipitation trends point to

a modest redistribution of the monsoonal precipitation (Fig. 4.6) and no significant

increasing trend in total precipitation has been detected (Goswami et al., 2006; Fig. 4.5).

Yet, 80% of the semi-arid dry lands of India display significant increases in decadal LAI.

Page 122: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

The analysis of the 1981-2006 trend in monthly LAI (Fig. 4.7) reveals that the largest

increases in vegetation growth have occurred during the months of January and February,

which correspond to the peak of the rabi cropping season. The rabi cropping season

starts at the end of the summer monsoon (November) and extends through the following

spring (February-May). Water for the rabi crops is supplied by the less abundant north-

east (winter) monsoon, by the moisture accumulated from the south west (summer

monsoon) during the kharif season, or, increasingly, by irrigation. Irrigation, beyond

making possible the cultivation of non-rainfed crops during the rabi season (i.e., a second

rice crop), also supplements crop water requirements during the kharif season, when

monsoon rains are delayed. It is therefore suggested that land use changes have been the

principal driver of the signal of enhanced plant growth detected from satellite over this

predominantly water-limited country.

To test this hypothesis, the surface area within each major Indian state that has a

statistically significant trend in LAI during the period 1981-2006 is regressed against the

corresponding change in net irrigated areas reported by the Directorate Of Economics &

Statistics of the Indian Ministry Of Agriculture (Naraynamoorthy, 2002) for the same

period (Fig. 4.8a). The regression indicates that the remarkable expansion in irrigated

areas that took place in the country throughout the 1980s until the mid 1990s explains

(R2=0.32, pvalue<0.05) the enhancement in photosynthetic activity observed over large

portions of India. Figure 4.8b includes all pixels within each major Indian state with a

positive trend in LAI for the period 1981-2006, thus suggesting the fact that different

Page 123: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

regions may not have a consistent increase in vegetation over a longer time scale, but

may show significant changes in vegetation even for smaller time scales.

Noteworthy are the states of Madhya Pradesh and Rajasthan where decadal scale

changes in LAI and changes in net irrigated area have been significantly higher than other

states (Fig. 4.8 and Fig. 4.9). Large scale increases in decadal LAI are seen in Mandsaur,

Jhalawar, Ujjain, Shajapur, Ratlam and Kota districts (Fig. 4.9). In particular, the semi-

arid region of Mandsaur district is spread over an area of 5554 Km2 with approximately

~1600 inhabited villages and water for irrigation is sustained through several macro-level

watersheds that are spread over an area of around 15500 hectares across the Sitamau and

Mandsaur blocks (http://fes.org.in/includeAll.php?pId=Mi0yNi0z). These districts are

under the coverage of the Chambal Valley Project, which facilitates large scale

establishment of dams for providing hydro-electric power and water for irrigation and

agricultural practices.

Further enhancements to vegetation productivity can be explained by the increase in

fertilizer use intensity that accompanies the resource-demanding high yield varieties of

crops which tend to substitute traditional cultivars once the supply of water is ensured

through irrigation (Fig. 4.10b) (Bhattaray and Naraynamoorthy, 2003;

http://dacnet.nic.in/ ). As displayed in Figure 4.10a, currently in most Indian States, more

than half of the cropped area is irrigated and increasingly the water used for irrigation is

derived from groundwater sources, representing up to 80% of all water sources used for

irrigation in some states, such as in Punjab and Uttar Pradesh (Naraynamoorthy, 2002).

Access to microcredit and to heavily subsidized electric power have favored the

Page 124: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

expansion of private wells to irrigate fields distant from major irrigation infrastructures

(Shah, 2005), particularly benefiting the Indian agriculture by increasing crop yields and

total food production, and, thus contributing to alleviating rural poverty.

These greening trends, however, are not expected to continue to be as strong over

these regions in the future. Over India, the annual LAI has already slowed down, and it is

reflected in a flattening of growth in total food production (FAOStat, 2007). The reasons

for this slowdown are complex. Since the mid 1990’s, a number of basins are

increasingly suffering from groundwater overexploitation and are at risk of salinization

(http://cgwb.gov.in/gw_profiles) and commodity prices have been declining due to

globalization (Narayanamoorthy, 2007), reducing the farmers potential investments in

production. While these factors can be reversed through increased irrigation efficiency

and proper policy, the current trends in climate, if here to stay, will further dampen the

trends.

4.4 Conclusions

Widespread portions of the tropical dry lands of the densely populated and

developing eastern hemisphere have shown marked trends in vegetation productivity over

the period 1981-2006 as measured by the 26 years of monthly LAI data. For most of the

countries belonging to Africa’s dry lands and Australia, an increase in peak annual

precipitation can be identified as the principal driver of the greening. In countries such as

Turkey, Saudi Arabia and India, the analysis suggests that over the study period, although

climate has been favorable to vegetation growth, the main driver of greening trend have

Page 125: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

been land cover changes associated with the expansion of irrigation. However, predicting

whether the greening trends in India and in other countries of the semi-arid tropics

undergoing rapid socio-economic change will continue, flatten or decline will likely

depend from a variety of climatic, land use management and policy factors. Projecting

how future climate changes will impact the tropical dry lands is a difficult task as it will

depend on biophysical feedbacks and on societal and economic behaviors caused by

adaptation but also forces of globalization. Establishing what causes the changes in

vegetation dynamics is the first step to improve our skill in such projections and

assessments.

Page 126: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.1 Panel (a) shows colormap of peak annual NDVI climatology. Peak annual NDVI climatology was calculated by first estimating the 26-yr (1981-2006) mean of monthly NDVI (monthly NDVI climatology) and then selecting the maximum value (per pixel, from 12 monthly climatological NDVI values). A spatial mask was applied on the colormap based on peak annual NDVI climatology values in the range of 0.12-0.55. Panel (b) shows the spatial frequency distribution (number of pixels) of peak annual NDVI climatology values in the range 0.12-0.55.

(a)

(b)

Page 127: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.2 Percentage distribution of IGBP land cover classes (Panel (a)) and frequency distribution of bare (red), herbaceous (blue) and tree (black) cover from MODIS VCF map - expressed as percentage of total number of pixels (Panel (b)) for peak annual NDVI climatology range of 0.12-0.55.

(a)

(b)

Page 128: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.3 Colormap of percentage change in mean peak annual LAI between decade 1 (1981-1990) and decade 2 (1995-2006). For each year in a decade, the maximum/peak LAI was selected (per pixel from 12 LAI values). The mean peak LAI was calculated for each decade. Finally, the percentage change was calculated as [100*(mean peak LAI decade2 – mean peak LAI decade1)/(mean peak LAI decade1)]. A spatial mask was applied on the colormap based on peak annual NDVI climatology values in the range of 0.12-0.55 (all values outside this range appear in gray – masked out).

Page 129: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.4 Colormap of percentage change in mean peak annual precipitation (mm/yr) between decade 1 (1981-1990) and decade 2 (1995-2006). The peak precipitation for each year was calculated by summing the precipitation in three wettest months. The mean peak annual precipitation was calculated for each decade. Finally, the percentage change was calculated as [100*(mean peak annual precipitation decade2 – mean peak annual precipitation decade1)/(mean peak annual precipitation decade1)]. A spatial mask is applied on the colormap based on peak NDVI climatology values in the range of 0.12-0.55 (all values outside this range appear in gray – masked out).

Page 130: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.5 Plots of peak annual LAI anomaly and peak annual rainfall anomaly for 8 representative countries in the study region – Australia, Botswana, India, Mali, Niger, Saudi Arabia, South Africa and Turkey. Peak annual LAI anomaly was calculated as (peak annual LAI – peak annual LAI climatology). Peak annual precipitation anomaly was calculated as (peak annual precipitation – peak annual precipitation climatology). Peak annual precipitation climatology is the 26-yr (1981-2006) mean of peak annual precipitation. For each country the mean of all anomaly pixels in the peak annual NDVI climatology range of 0.12-0.55 was calculated and plotted for each year.

Page 131: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.6 Monthly average temperature and monthly precipitation trend for India in the 1980’s and 1990’s.

Page 132: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.7 Long term mean monthly LAI (red line) and percent trend in monthly LAI (green line) for India over the time period 1981-2006. The long term mean monthly LAI was calculated by averaging the maximum monthly LAI of each pixel over the time period 1982-2006. The spatially averaged mean monthly LAI is then plotted for each month. The monthly trend of LAI is calculated as the slope of a linear regression fitted through the spatially averaged maximum LAI of each month as a function of the time period 1982-2006. The percent trend is then calculated as: LAI trend (%) = [slope * 26/1982 monthly maximum LAI]*100. Only pixels falling within the peak annual NDVI climatology mask of 0.12-0.55 were considered in the calculation.

Page 133: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.8 Panel (a) shows cross plot between area with increased LAI (pixels with statistically significant (pvalue<0.05) positive LAI trend during the period 1981-2006) within each major Indian state and change in net irrigated area over the same period. Panel (b) is a similar plot as panel (a) but with all pixels with positive LAI trend during the same period.

(a)

(b)

Page 134: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.9 Percentage change in mean peak annual LAI as in Fig. 4.3 for the semi-arid districts of Mandsaur (State: Madhya Pradesh), Kota (State: Rajasthan), Jhalawar (State: Rajasthan) and Ujjain (State: Madhya Pradesh) in India. Decadal scale change in LAI shows a percent increase of more than 50% in these districts. The black boundaries are state boundaries, while the white boundaries depict district level partition.

Page 135: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Figure 4.10 The percentage of cropped area that is irrigated (blue bar) and percentage of irrigated land utilizing groundwater (green bar) for each of the major Indian states (Panel (a)). Panel (b) tabulates the fertilizer consumption (kg/Ha) and fraction irrigated sown area (in %) for different states (year 2002-2003) and shows the corresponding regression relation (orange squares represents states).

(a)

(b)

(b)

Page 136: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects
Page 137: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Chapter 5

5. Concluding remarks

The monitoring and modeling of the terrestrial biosphere within the larger context

of climate variability and change studies requires multi-decadal time series of key

biophysical variables characteristic of vegetation structure and functioning (NRC

Decadal Survey, 2007; GCOS 2nd Adequacy Report, 2007). Prior research (cf. Sections

1.1.1.1 and 1.1.1.2) has shown the relevance of vegetation in studies of land use and

cover change, terrestrial productivity and modeling of biogeochemical cycles and climate

(cf. Section 1.1.2). Consequently, there is now a pressing need to develop methodologies

for generating long-term Earth System Data Records (ESDRs) from remote sensing data

collected with different sensors over the past three decades.

LAI and FPAR are well defined and measurable characteristics of vegetation and

are independent of the properties of satellite sensors, although, due to differences in

sensor characteristics and methodological issues, it is a challenging task to build a

seamless consistent long term data record of these biophysical variables from multiple

sensors. A key step in assembling these long-term data records is establishing a link

between data from earlier sensors (e.g. AVHRR) and present/future sensors (e.g. MODIS

TERRA, NPOESS) such that the derived products are independent of sensor

characteristics and represent the reality on the ground both in absolute values and

variations in time and space (Van Leeuwen et al., 2006). Multi-decadal global data sets of

LAI and FPAR of known accuracy and produced with a physically based algorithm are

Page 138: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

currently not available, although several recent attempts have resulted in shorter term

research quality data sets from medium resolution sensor data (Knyazikhin et al., 1998a;

Gobron et al., 1999; Chen et al., 2002; Yang et al., 2006a; Plummer et al., 2006; Baret et

al., 2007). Thus the objective for this research was to formulate and demonstrate the

performance of a synergistic approach for retrieving LAI and FPAR from measurements

of multiple sensors.

Three specific themes are addressed in this dissertation. The first is regarding the

challenges underlying the generation of continuous time series of biophysical variables

from multiple sensors, that addresses methodological issues in remote sensing science

and differences in sensor properties. In general, ESDR (Earth System Data Record)

algorithms ingesting data from different instruments should account for differences in

spatial resolution, spectral characteristics, uncertainties due to atmospheric effects and

calibration, information content, etc. The theoretical approach of our algorithm is based

on the radiative transfer theory of spectral invariants. Accordingly, the canopy spectral

Bidirectional Reflectance Factor (BRF) is parameterized in terms of a compact set of

parameters – spectrally varying soil reflectances, single scattering albedo, spectrally

invariant canopy interceptance, recollision probability and the directional escape

probability. This approach ensures energy conservation and allows decoupling the

structural and radiometric components of the BRF.

According to the formulated theory, the single scattering albedo accounts for the

dependence of BRF on sensor’s spatial resolution and spectral bandwidth. Another

parameter that accounts for variation in the information content of the remote

Page 139: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

measurements is the data uncertainty. Both single scattering albedo and data uncertainty

are two key configurable parameters in the multi-sensor retrieval algorithm. The

algorithm supports two modes of operation: the MODIS mode (retrievals from BRF) and

the AVHRR mode (retrievals from NDVI). In both cases, the algorithm simulates similar

mean LAI values with differences in corresponding dispersions indicating varying input

information content (MODIS BRF vs. AVHRR NDVI) and related uncertainties. Overall,

the problem of generating LAI/FPAR is reduced to the problem of finding values of data

uncertainty and single scattering albedo for which: (a) the consistency requirements (cf.

Section 2.2) for retrievals from MODIS and AVHRR are met; (b) the difference between

MODIS and AVHRR LAI/FPAR values is minimized; (c) the probability of retrieving

LAI/FPAR is minimized.

The second theme of this dissertation is implementation of the physically based

multi-sensor algorithm for retrieving global fields of LAI from GIMMS AVHRR NDVI

for the period July 1981 to December 2006, and subsequent evaluation of the new global

monthly AVHRR LAI data set for its validity and accuracy. The evaluation of the data set

is done both through direct comparisons to ground data and indirectly through inter-

comparisons with similar data sets. This included comparisons with existing LAI

products (MODIS and CYCLOPES LAI products for the 2000 to 2003 period of overlap)

at a range of spatial scales, and correlations with key climate variables in areas where

temperature and precipitation limit plant growth. Overall, there is good agreement

between the AVHRR and MODIS data sets at scales ranging from global to regional to

pixel. At the global scale, the AVHRR values explain 97.5% of the variability in the

Page 140: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

MODIS product and will be in error in their estimation by 0.18 LAI, on average. The

regional and pixel-scale inter-comparison suggests an average error of less than 0.3 LAI.

Comparison with CYCLOPES LAI indicates satisfactory agreement in most of the

biomes with RMSE values less than 0.5 LAI.

The data set was also analyzed to reproduce well-documented spatio-temporal

trends and interannual variations in vegetation activity in the northern latitudes where

temperature limits plant growth, and in semi-arid areas where precipitation limits plant

growth. Additionally, to assess the mechanistic basis underlying the observed correlations

between LAI and temperature in the northern latitudes and LAI and precipitation in the

semi-arid tropics, a multivariate data-reduction technique (canonical correlation analysis)

was used to isolate well correlated modes of spatio-temporal variability between LAI and

the climate variables. The isolated modes suggest El Niño-Southern Oscillation and

Arctic Oscillation as key drivers of linked interannual variations in vegetation greenness

and precipitation in the semi-arid regions and, vegetation greenness and surface

temperature in the northern latitudes, respectively.

The derived LAI data were compared to field measurements and high-resolution

LAI maps from a host of field sites representative of all the major vegetation classes. The

comparison with plot scale measurements over biome specific homogeneous patches

indicates a 7% underestimation in the AVHRR LAI when all major vegetation types are

considered. The error in mean values obtained from distributions of AVHRR LAI and

high-resolution field LAI maps for different biomes is within 0.6 LAI, with the exception

of a broadleaf evergreen forest site. These validation exercises though limited by the

Page 141: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

amount of field data, and thus less than comprehensive, nevertheless indicate satisfactory

comparability between the LAI product and field measurements. In summary, the inter-

comparison with other short-term LAI data sets, evaluation of long term trends with

known variations in climate variables, and validation with field measurements together

build confidence in the utility of this new 26 year LAI record for long term vegetation

monitoring and modeling studies.

The third and final theme of this dissertation is a case study of interpreting climate

and land use impacts on vegetation in the semi-arid tropics utilizing the long term LAI

data set. Results indicate that widespread portions of the tropical dry lands of the densely

populated and developing eastern hemisphere have experienced marked trends in

vegetation productivity over the period 1981-2006 as measured by the 26 years of

monthly LAI data. Notable increases in annual maximum LAI are observed between the

decade 1981-1990 and 1995-2006 in over 70% of the study area, encompassing Turkey,

large portions of the Middle Eastern countries, the Sahel, Horn of Africa and Southern

African countries, most of tropical Asia and portions of Australia. About 29% of the area

shows decline in photosynthetic activity, principally distributed in eastern and southern

Australia, along the Namibian Desert, and other portions of the coast of Western Africa

up to the Iberian Peninsula. For most of the countries belonging to Africa’s dry lands and

Australia, an increase in peak annual precipitation can be identified as a the principal

driver of greening.

In countries such as Turkey, Saudi Arabia and India, the analysis suggests that over

the study period, although climate has been favorable for vegetation growth, the main

Page 142: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

driver of greening trend has been land cover changes associated with the expansion of

irrigation. About 80% of the semi-arid dry lands of India display significant increases in

decadal LAI, despite no significant trend in total precipitation. The analysis of the 1981-

2006 trend in monthly LAI reveals that the largest increases in vegetation growth have

occurred during the months of January and February, which correspond to the peak of the

rabi cropping season (predominantly irrigation driven). Further analysis suggest that land

use changes have been the principal driver of enhanced plant growth in this

predominantly water-limited country. However, predicting whether the greening trends in

India and in other countries of the semi-arid tropics undergoing rapid socio-economic

change will continue, flatten or decline will likely depend on a variety of climatic, land

use management and policy factors. Projecting how future changes will impact the

tropical dry lands is a difficult task as it will depend on biophysical feedbacks and on

societal and economic behaviors caused by adaptation but also forces of globalization.

Establishing what causes the changes in vegetation dynamics is the first step to improve

our skill in such projections and assessments.

The results presented in this dissertation are the first comprehensive analysis of a

physically derived multi-decadal global data set of LAI. They imbue confidence in

utilizing this seamless, consistent product for large scale terrestrial-biosphere models and

for monitoring global vegetation dynamics. Despite the robustness of the methodological

approach and accuracy of the derived product, there are inevitably certain limitations.

First, data measurement uncertainties from different sensors can significantly impact the

retrieval of a biophysical product. This requires better calibration and atmospheric

Page 143: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

correction algorithms, along with solar and view angle corrections for surface reflectance.

Secondly, global retrievals of biophysical products utilize land cover classification maps

which set the basis for identifying the spatial heterogeneity of biome distribution.

Classification inaccuracies are a critical factor, especially for those cases where regions

show dramatic changes in land cover dynamics (for e.g. changes from herbaceous to

woody biomes). The validity of the LAI product during the 1980’s and 90’s represents a

more challenging problem in terms of land cover dynamics as the present algorithm relies

on a single land cover map for the entire period. Thirdly, the retrieval of LAI is an ill

posed problem and an optimal combination of data information content and overall

uncertainty is key to achieving continuity in multi-sensor times series of LAI products.

The dispersion of the retrieved LAI solution set can be minimized provided the input data

has more spectral information. In general, a larger volume and higher accuracy of

measured information corresponds to a better localized set of solutions (LAI values close

to ground truth). Finally, direct validation of coarse resolution products with ground

measurements is a complicated task due to reasons pertaining to scaling of plot level

measurements to sensor resolution, geo-location uncertainties, limited temporal and

spatial sampling of ground data, field instrument calibration, sampling errors, etc

(Buermann et al., 2002; Tan et al., 2005; Huang et al., 2006; Yang et al., 2006a ; Weiss et

al., 2007). The accuracy of the direct validation exercise is a function of area

homogeneity as comparing field level measurements with larger pixels of a satellite

product is a valid exercise only if performed over regions with same land cover type.

In the future, this research could continue along the following directions:

Page 144: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

a) The proposed approach for retrieving the LAI fields can be applied to AVHRR

surface reflectance data. Efforts are underway to perform rigorous physically based

calibration and atmospheric correction to achieve consistency with the MODIS surface

reflectance data (Vermote and Saleous, 2006). Once the data are available from 1981

onwards, the algorithm can be applied to generate LAI fields with better retrieval

accuracy and precision (higher information content; cf. Section 2.6).

b) The scale dependency is a critical issue in retrieving LAI across multiple sensors.

The scaling methodology described in this research can be seen as a benchmark for

retrieving LAI fields at any given resolution for any given sensor. The single scattering

albedo depends on the scale at which it is defined. For e.g., the single scattering albedo of

a needle, shoot, branch, tree crown, etc., are different. The theory of canopy spectral

invariants will thus provide a framework through which structural information can be

maintained in a self-consistent manner across multiple scales (cf. Section 2.4). In a

similar way, the single scattering albedo will also account for the sensitivity in the

spectral bandwidths (cf. Section 2.5). This algorithm can thus be applied to retrieve LAI

at finer resolutions (e.g. Landsat), thus allowing to better capture the spatial heterogeneity

of leaf dynamics. In the future, to ensure data continuity of LAI, surface reflectances

from VIIRS onboard NPOESS should be analyzed to maintain product consistency with

the AVHRR and MODIS data.

c) Discrepancies between field measurements and satellite observations also arise

due to the scaling problem. The understanding of scale dependency from this work will

facilitate an improved validation scheme to better compare coarse resolution retrievals

Page 145: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

with field measurements (cf. Section 3.5), as well as explain the physics behind inter-

comparing data of different resolutions and from multiple sensors (cf. Section 3.3.3).

d) The consistent long term data record of LAI and FPAR can be utilized to produce

a long term GPP/NPP time series based on the MODIS NPP logic (Nemani et al., 2003).

NPP is the source of most food, fiber and fuel, and changes in NPP integrate climatic,

ecological, geochemical, and human influences on the biosphere (Nemani et al., 2003).

NPP algorithm inputs vegetation parameters (Land cover type, LAI, FPAR) and daily

climate data (incident solar radiation (IPAR), minimum and average air temperatures and

humidity). NPP estimates are sensitive to uncertainties in input LAI/FPAR (e.g.

differences in LAI from multiple sensors) and hence the product from this research will

hopefully improve future NPP estimates.

e) The multiyear global LAI data set will be a significant input to different climate

models (cf. Section 1.1.2) for investigating the response of ecosystems to changes in

climate, carbon cycle, land cover and land-use. An improvement over the long term data

set would be to further create a consistent dynamic vegetation layer or an improved

phenology record covering the AVHRR, MODIS and NPOESS eras. The algorithm

proposed in this research also accounts for generating consistent surface reflectances

across multiple sensors, thus extending the scope of this study to create consistent

vegetation indices such as the enhanced vegetation index (EVI). EVI has improved

sensitivity in high biomass regions and improved vegetation monitoring through a de-

coupling of the canopy background signal and a reduction in atmosphere influences

(Huete et al., 2002). Overall, long term global data sets of leaf area index and phenology

Page 146: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

with a monthly temporal resolution will be an indispensable input to coupled climate-

vegetation-land-surface models to quantify global land cover change and terrestrial

productivity in the context of climate change, land use change and anthropogenic

influences.

f) Finally, following the case study in Chapter 4, research can be extended to

develop a deterministic model for anticipating changes in crop productivity and/or

vegetation greenness due to continued warming over the semi-arid tropical regions (as

projected in the IPCC 2007), especially in countries like India and China and in the

Sahel. A convincing stride would be to explore the further sustainability of the greening

trend as observed in a developing and highly populated country like India, where the

greening due to land use change is dominant and countries in the Sahel, where

precipitation induced greening is significant. Due to over exploitation of groundwater for

irrigation, changes in policies subsidizing crop inputs and subsequent projections in

future warming trends, it will be a challenging food security scenario for a large number

of developing countries with rapidly increasing population in the semi-arid tropics.

Page 147: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Appendices

Appendix A: Analytical expression of the “S” problem

The second term on the right hand side of Eqs. (4) and (5) describes the

contribution of multiple interactions between the ground and the canopy to the total

canopy BRF and absorptance. Let the downward flux at the surface level be tBS, in the

case of a black surface. The incoming flux after interacting with the ground will act as the

initial source at the surface. The reflected radiation flux (tBS, . sur,) will interact with the

canopy further and return to the surface (tBS, . sur, . rS,), where sur, and rS, are the

hemispherically integrated ground and canopy reflectance, respectively. Let JS,() be the

radiance from the Lambertian surface for a unit radiant exitance. Taking into account that

the radiant exitance at the first interaction is (tBS, . sur,), the corresponding radiance from

the surface can be expressed as (JS,(). tBS, . sur,.). The total radiance, S, can be

expressed as the sum of successive orders of scattering,

S= tBS, sur, JS,()+2sur, rS, tBS, JS,()+3

sur, r2S, tBS, JS,()+ …

… +nsur, rn-1

S, tBS, JS,()

where n is the order of scattering. The above geometric series can be rewritten in the

closed form

. (A1)

The spectral invariant approximations for tBS, and JS,() are

Page 148: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

. (A2)

. (A3)

where, the term t0 is the zero order direct transmittance, T1()= τ1()i0 and T2()=

τ2()pti0. τ1 and τ2 are probabilities that the scattered photons can escape the lower

boundary of the canopy. J0, J1() and J2() are analogous to t0, T1() and T2(). The

term tBS, is the sum of the two components. The first is the zero order or uncollided

transmittance, t0=1 – i0, which is defined as the probability that a photon in the incident

flux will arrive at the bottom of canopy without suffering a collision. The second

component represents transmittance of the diffuse radiation, i.e., the probability that a

photon will exit the vegetation canopy through the lower boundary after one or more

interactions. The expression for diffuse transmittance is obtained by hemispherically

averaging Eq. (1) over downward directions. The expression for spectral reflectance, rS, ,

is obtained by hemispherically integrating Eq. (1), formulated for i0,S instead of i0, over

the upper hemisphere,

(A4)

Page 149: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Appendix B: Ancillary tables

Table B1. BELMANIP sites used for CYCLOPES LAI and AVHRR LAI inter-comparison

Site name (Country) Site ID

Lat Lon Biome Type

ARM/CART Shilder (USA) 32 36.93oN 96.86oW GrassesKonza (USA) 36 39.08oN 96.57oW GrassesLarzac (France) 44 43.93oN 3.12oE GrassesBarrax (Spain) 35 39.07oN 2.10oW Broadleaf cropsAlpilles (France) 43 43.80oN 4.74oE Broadleaf cropsPlan-de Dieu (France) 45 44.20oN 4.95oE Broadleaf cropsFundulea (Romania) 47 44.40oN 26.58oE Broadleaf cropsTshane (Botswana) 7 24.00oS 21.83oE SavannasOkwa (Botswana) 8 22.40oS 21.71oE SavannasMongu (Zambia) 11 15.44oS 23.25oE SavannasBurkina, Ghana (Burkina Faso)

385 10.86oN 3.07oW Savannas

Kejimikujik (Canada) 48 44.35oN 65.19oW Needle leaf forestsThompson, Manitoba (Canada)

81 56.05oN 98.15oW Needle leaf forests

NOBS-BOREAS NSA (Canada)

82 53.66oN 105.32oW Needle leaf forests

Jarvselja (Estonia) 85 58.30oN 27.26oE Needle leaf forestsRuokolahti (Finland) 88 61.52oN 28.71oE Needle leaf forestsHirsinkanjas (Finland) 89 62.64oN 27.01oE Needle leaf forestsFlakaliden (Sweden) 90 64.11oN 19.47oE Needle leaf forestsRovaniemi (Finland) 91 66.45oN 25.34oE Needle leaf forests

Page 150: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Table. B2. Summary of MODIS LAI Field Campaigns used for Validation with AVHRR LAI

Site(Country)

Lat/Lon Biome Type Date LAI

Bondville, Illinois (AGRO, USA)

40.007oN/88.292oW

Broadleaf Crops Aug 2000 3.60

Fundulea(Romania)

44.410oN/26.570oE

Broadleaf Crops (Mar, May) 2001Jun 2002(May, Jun) 2003

1.071, 1.8781.3091.063, 1.10

Barrax(Spain)

39.060oS/2.100oW

Broadleaf Crops Jul 2003 0.965

Alpilles(France)

43.810oN/4.750oE

Grasses/Cereal Crops

Mar 2001Jul 2002

0.9281.054

Haouz(Morocco)

31.660oN/7.600oW

Shrubs Mar 2003 1.20

Turco(Bolivia)

18.240oS/68.200oW

Shrubs Apr 2003 0.10

Konza Prairie(USA)

39.080oN/96.570oW

Grasses Jun 2000 1.96

Dahra (Senegal)

15.350oN/15.480oW

Grasses/ Savannas

Aug 2001Aug 2002

2.000.40

Pandamatenga(Botswana)

18.650oS/25.500oE

Savannas Mar 2000 1.24

Maun(Botswana)

19.920oS/23.600oE

Savannas Mar 2000 1.52

Mongu(Zambia)

15.440oS/23.253oE

Savannas Apr 2000Sep 2000

1.900.80

TessekreNorthTessekreSouth(Senegal)

15.810oN/15.070oW

Shrubs

Shrubs

Aug 2002

Aug 2002

0.35

0.30

Tshane(Botswana)

24.160oS/21.893oE

Savannas Mar 2000 0.78

Okwa(Botswana)

22.400oS/21.713oE

Savannas Mar 2000 1.28

Hirsikangas(Finland)

62.520oN/27.030oE

Needle Leaf forests

Aug 2003Jun 2005

2.5481.419

Ruokolahti(Finland)

61.320oN/28.430oE

Needle Leaf forests

Jun 2000 2.06

Harvard Forest (HARV, USA)

42.530oN/72.173oW

Deciduous Broadleaf forests

Jul 2000Jul 2001

5.085.50

Wisconsin(USA)

45.800oN/90.080oW

Deciduous Broadleaf forests

May 2002Jul 2002

1.705.70

Concepcion(Chile)

37.467oS/ Evergreen Jan 2003 3.096

Page 151: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

73.470oE Broadleaf forestsDemmin(Germany)

53.892oN/13.207oE

Broadleaf Crops Jun 2004 2.632

Jarvselja(Estonia)

58.292oN/27.260oE

Needleleaf forests

Jul 2000Jun 2001Jun 2002

2.9252.754.201

Laprida(Argentina)

36.990oS/60.552oW

Grasses Nov 2001Oct 2002

4.1241.923

Larose(Canada)

45.380oN/75.217oW

Needleleaf forests

Aug 2003 3.581

Nezer(France)

44.567oN/1.038oW

Needle Leaf forests

Jul 2000Apr 2001Apr 2002

1.4431.4351.331

Plan-de-Dieu(France)

44.198oN/4.948oE

Broadleaf Crops Jul 2004 0.469

Rovaniemi(Finland)

66.455oN/25.351oE

Needleleaf forests

Jul 2004Jun 2005

1.2481.401

Sud_Ouest(France)

43.506oN/1.237oE

Broadleaf Crops Jul 2002 1.228

Wankama(Niger)

13.644oN/2.635oE

Grasses Jun 2005 0.081

The references regarding further description of site characteristics are provided in Table I of Yang et al., 2006a and Table 3.4 of Chapter 3. Detailed methodologies and documentation of field campaigns can be obtained from http://mercury.ornl.gov/ornldaac/ and http://lpvs.gsfc.nasa.gov/lai_intercomp.php.

Page 152: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

List of Journal Abbreviations

Adv. Geophys. Advances in Geophysics

Agric. For. Meteorol. Agricultural Forest Meteorology

Agric. Sys. Agricultural Systems

Bull. Am. Meteorol. Soc. Bulletin of American Meteorology Society

Climate Dyn. Climate Dynamics

Curr. Sci Current Science

Earth Inter. Earth Interactions

Econ. Polit. Week. Economic Political Weekly

Geophys. Res. Lett. Geophysical Research Letters

Global Biogeochem. Cy. Global Biogeochemical Cycles

Global Change Biol. Global Change Biology

Global Environ. Ch. Global Environmental Change

Global Planet. Ch. Global and Planetary Change

IEEE Trans. Geosci. Remote Sens. IEEE Transactions on Geoscience and Remote

Sensing

Int. J. Biometeorol. International Journal of Biometeorology

Int. J. Climatol. International Journal of Climatology

Int. J. Remote Sens. International Journal of Remote Sensing

J. Appl. Meteorol. Journal of Applied Meteorology

J. Arid Environ. Journal of Arid Environments

Page 153: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

J. Atm. Sci. Journal of Atmospheric Sciences

J. Biogeogr. Journal of Biogeography

J. Climate Journal of Climate

J. Geophys. Res. Journal of Geophysical Research

J. Hydrometeorol. Journal of Hydrometeorology

J. Quant. Spec. Rad. Trans. Journal of Quantitative Spectroscopy &

Radiative Transfer

Mon. Weather Rev. Monthly Weather Review

Proc. Natl. Acad. Sci. U.S.A. Proceedings of the National Academy of

Sciences of the United States of America

Qtrly. J. Royal Meteorol. Soc. Quaterly Journal of the Royal Meteorological

Society.

Remote Sens. Environ. Remote Sensing of Environment

Remote Sens. Rev. Remote Sensing Reviews

Tree Physiol. Tree Physiology

Trends in Ecol. & Evol. Trends in Ecology and Evolution

Water Res. J. Water Resources Journal

Water Sci. Tech. Water Science Technology

Page 154: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Bibliography

WWW Sites

WWW1: NOAA KLM User’s Guide, Appendix D.2.

http://www2.ncdc.noaa.gov/docs/klm/html/d/app-d2.htm#fd2-1

WWW2: MODIS Characterization Support Team (MCST), L1B Relative Spectral

Response LUT.

http://www.mcst.ssai.biz/mcstweb/index.html

WWW3: Monthly Atmospheric and SST Indices, NOAA/NWS/NCEP Climate Prediction

Center.

http://www.cpc.ncep.noaa.gov/data/indices/wksst.for

WWW4: Climate Diagnostics Bulletin, NOAA/NWS/NCEP Climate Prediction Center,

December, 2007.

http://www.cpc.ncep.noaa.gov/products/CDB/Tropics/figt5.shtml

Adeel, Z. et al. (2007). Overcoming one of the greatest environmental challenges of our times: Re-thinking policies to cope with desertification. Policy brief. United Nations, pp. 46.

Angert, A., Biraud, S., Bonfils, C., Henning, C. C., Buermann, W., Pinzon, J., Tucker, C. J., & Fung, I. (2005). Drier summers cancel out the CO2 uptake enhancement induced by warmer springs. Proc. Natl. Acad. Sci.

U.S.A., 102, 10823-10827.Anyamba, A., & Tucker, C. J. (2005). Analysis of Sahelian vegetation dynamics using

NOAA-AVHRR NDVI data from 1981-2003, J. Arid Environ., 63(3), 569-614.

Asrar, G., Fuchs, M., Kanemasu, E. T., & Hatfield, J. L. (1984). Estimating absorbed

photosynthetic radiation and leaf-area index from spectral reflectance in wheat.

Agronomie, 76, 300-306.

Page 155: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Barber, V. A., Juday, G. P., & Finney, B. P. (2000). Reduced growth of Alaskan white spruce in the twentieth century from temperature-induced drought stress. Nature, 405, 668–673.

Baret, F., & Guyot, G. (1991). Potentials and limits of vegetation indices for LAI and

APAR assessment. Remote Sens. Environ., 35, 161-173.

Baret, F., Morissette, J., Fernandes, R., Champeaux, J. L., Myneni, R. B., Chen, J.,

Plummer, S., Weiss, M., Bacour, C., Garrigues, S., & Nickeson, J. (2006). Evaluation

of the representativeness of networks of sites for the global validation and inter-

comparison of land biophysical products. Proposition of the CEOS-BELMANIP,

IEEE Trans. Geosci. Remote Sens., 42(7), 1794-1803.

Baret, F., Hagolle, O., Geiger, B., Bicheron, P., Miras, B., Huc, M., Berthelot, B., Niño,

F., Weiss, M., Samain, O., Roujean, J. L., & Leroy, M. (2007). LAI, fAPAR, and

fCover CYCLOPES global products derived from VEGETATION Part 1: Principles

of the algorithm. Remote Sens. Environ., 110, 275-286.

Barnett, T. P., & Preisendorfer, R. (1987). Origins and levels of monthly and seasonal

forecast skill for United States surface air temperatures determined by canonical

correlation analysis, Mon. Weather Rev., 115, 1825– 1850.

Bhattaray, M., & Naraynamoorthy, A. (2003). Impact of irrigation on rural poverty in

India: an aggregate panel-data analysis. Water Pol., 5, 443-458.

Bjornsson, H., & Venegas, S. A. (1997). A manual for EOF and SVD analysis of climate

data. McGill University, CCGCR Report No. 97-1, Montreal, Quebec, pp. 52.

Bonan, G. B., Pollard, D., & Thompson, S. L. (1992). Effects of boreal forest vegetation

on global climate. Nature, 359, 716-718.

Bonan, G. B. (1998). The land surface climatology of the NCAR Land Surface Model

coupled to the NCAR Community Climate Model. J. Climate, 11, 1307-1326.

Bonan, G. B., Levis, S., Sitch, S., Vertenstein, M., & Oleson, K. W. (2003). A dynamic

global vegetation model for use with climate models: concepts and description of

simulated vegetation dynamics. Global Change Biol., 9, 1543-1566.

Page 156: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Brown, M. E., Pinzon, J. E., Morisette, J. T., Didan, K., & Tucker, C. J. (2006).

Evaluation of the consistency of long term NDVI time series derived from AVHRR,

SPOT-Vegetation, SeaWIFS, MODIS, and Landsat ETM+. IEEE Trans. Geosci.

Remote Sens., 44(7), 1787-1793.

Buermann, W., Dong, J. R., Zeng, X. B., Myneni, R. B., & Dickinson, R. E. (2001).

Evaluation of the utility of satellite-based vegetation leaf area index data for climate

simulations. J. Climate, 14, 3536-3550.

Buermann, W., Wang, Y., Dong, J., Zhou, L., Zeng, X., Dickinson, R. E., Potter, C. S., &

Myneni, R. B. (2002). Analysis of a multiyear global vegetation leaf area index data

set. J. Geophys. Res., 107, 1-15.

Buermann, W., Anderson, B., Tucker, C. J., Dickinson, R. E., Lucht, W., Potter, C. S., &

Myneni, R. B. (2003). Interannual covariability in northern hemisphere air

temperatures and greenness associated with El Niño-Southern oscillation and the

Arctic oscillation. J. Geophys. Res., 108, 1-16.

Bunn, A. G., & Goetz, S. J. (2006). Trends in satellite-observed circumpolar photosynthetic activity from 1982 to 2003: The influence of seasonality, cover type, and vegetation density. Earth Inter., 10(12), 1-19.

Cao, M., Prince, S. D., Small, J. & Goetz, S. J. (2004). Remotely sensed interannual

variations and trends in terrestrial net primary productivity 1981-2000. Ecosystems, 7,

233-242.

Carlson, T. N., & Ripley, D. A. (1997). On the relation between NDVI, fractional

vegetation cover, and leaf area index. Remote Sens. Environ., 62, 241-252.

Charney, J. G. (1975). Dynamics of deserts and drought in the Sahel. Qtrly. J. Royal

Meteorol. Soc. , 101, 193-202.

Chen, J. M., & Cihlar, J. (1996). Retrieving leaf area index of boreal conifer forests using

landsat TM images. Remote Sens. Environ., 55, 153-162.

Chen, J. M., Rich, P. M., Gower, S. T., Norman, J. M., & Plummer, S. (1997). Leaf area

index of boreal forests: Theory, techniques, and measurements. J. Geophys. Res.-

Page 157: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Atmospheres, 102, 29429-29443.

Chen, J. M., Pavlic, G., Brown, L., Cihlar, J., Leblanc S. G., White, H. P., Hall, R. J.,

Peddle, D. R., King, D. J., Trofymow, J. A., Swift, E., Van der Sanden, J., Pellikka, P.

K. E. (2002). Derivation and validation of Canada-wide coarse resolution leaf area

index maps using high-resolution satellite imagery and ground measurements. Remote

Sens. Environ., 80, 165-184.

Churkina, G., & Running, S. W. (1998). Contrasting climatic controls on the estimated

productivity of global terrestrial biomes. Ecosystems, 1, 206-215.

Cleland, E. E., Chuine, I., Menzel, A., Mooney, H. A., & Schwartz, M. D. (2007).

Shifting plant phenology in response to global change. Trends in Ecol. & Evol., 22,

357-365.

Cohen, W. B., Maiersperger, T. K., Turner, D. P., Ritts, W. D., Pflugmacher, D.,

Kennedy, R. E., Kirschbaum, A., Running, S. W., Costa, M., & Gower, S. T. (2006).

MODIS land cover and LAI collection 4 product quality across nine sites in the

western hemisphere. IEEE Trans. Geosci. Remote Sens., 44(7), 1843-1857.

Colombo, R., Bellingeri, D., Fasolini, D., & Marino, C. M. (2003). Retrieval of leaf area

index in different vegetation types using high resolution satellite data. Remote Sens.

Environ., 86, 120-131.

Combal, B., Baret, F., Weiss, M., Trubuil, A., Mace, D., Pragnere, A., Myneni, R. B. &

Knyazikhin, Y. (2002). Retrieval of canopy biophysical variables from bi-directional

reflectance using prior information to solve the ill-posed problem. Remote Sens.

Environ., 84(1), 1-15.

Dai, A., Fung, I. Y., & Del Genio, A. D. (1997). Surface observed global land

precipitation variations during 1900-88. J. Climate, 10, 2943-2962.

Dai, A., & Wigley, T. M. L. (2000). Global patterns of ENSO-induced precipitation,

Geophys. Res. Lett., 27, 1283-1286.

Dai, A., Trenberth, K. E., & Qian, T. (2004). A global data set of Palmer Drought

Severity Index for 1870-2002: Relationship with soil moisture and effects of surface

warming. J. Hydrometeorol., 5, 1117-1130.

Page 158: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Demarty, J., Chevallier, F., Friend, A. D., Viovy, N., Piao, S., & Ciais, P. (2007).

Assimilation of global MODIS leaf area index retrievals within a terrestrial biosphere

model. Geophys. Res. Lett., 34, L15402, doi:10.1029/2007GL030014.

Dickinson, R. E. (1983). land surface processes and climate surface albedos and energy-

balance. Adv. Geophys., 25, 305-353.

Dickinson, R. E., Hendersen-Sellers, A., Kennedy, P. J., & Wilson, M. F. (1986).

Biosphere-Atmosphere Transfer Scheme (BATS) for the NCAR CCM, NCAR Res.,

Boulder, CO, NCAR/TN-275-STR.

Dickinson, R. E., Shaikh, M., Bryant, R., & Graumlich, L. (1998). Interactive canopies

for a climate model. J. Climate, 11, 2823-2836.

Diner, D. J., Asner, G. P., Davies, R., Knyazikhin, Y., Muller, J. P., Nolin, A. W., Pinty,

B., Schaaf, C. B., & Stroeve, J. (1999). New directions in earth observing: Scientific

applications of multiangle remote sensing. Bull. Am. Meteorol. Soc., 80(11), 2209-

2228.

Diner, D. J., Braswell, B. H., Davies, R., Gobron, N., Hu., J., Jin, Y., Kahn, R. A.,

Knyazikhin, Y., Loeb, N., Muller, J. P., Nolin, A. W., Pinty, B., Schaaf, C. B., Seiz,

G., & Stroeve, J. (2005). The value of multiangle measurements for retrieving

structurally and radiatively consistent properties of clouds, aerosols, and surfaces.

Remote Sens. Environ., 97, 495-518.

Douville, H., & Royer, J. F. (1996). Influence of the temperate and boreal forests on the

Northern Hemisphere climate in the Meteo-France climate model. Climate Dyn., 13,

57-74.

Eklundh, L., & Olsson, L. (2003). Vegetation index trends for the African Sahel 1982-

1999. Geophys. Res. Lett., 30(1430), doi:10.1029/2002GL016772.

FAOStat 2007, Food and Agriculture Organization (2007). FAO Statistical Databases:

Agriculture, Fisheries, Forestry, Nutrition (Food and Agriculture Organization,

Rome).

Fassnacht, K. S., Gower, S. T., MacKenzie, M. D., Nordheim, E. V., & Lillesand, T. M.

(1997). Estimating the leaf area index of North Central Wisconsin forests using the

Page 159: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Landsat Thematic Mapper. Remote Sens. Environ., 61, 229-245.

Fernandes, R. A., Abduelgasim, A., Sylvain, L., Khurshid, S. K., & Butson, C. (2005).

Leaf area index maps at 30-m Resolution, selected sites, Canada. Data set. Available

on-line [http://daac.ornl.gov/] from Oak Ridge National Laboratory Distributed

Active Archive Center, Oak Ridge, Tennesse, U.S.A.

Field, C., & Mooney, H. A. (1983). Leaf age and seasonal effects on light, water, and

nitrogen use efficiency in a california shrub. Oecologia, 56, 348-355.

Fitzjarrald, D. R., Acevedo, O. C., & Moore, K. E. (2001). Climatic consequences of leaf

presence in the eastern United States. J. Climate, 14, 598-614.

Foley, J. A., Prentice, I. C., Ramankutty, N., Levis, S., Pollard, D., Sitch, S., & Haxeltine,

A. (1996). An integrated biosphere model of land surface processes, terrestrial carbon

balance, and vegetation dynamics. Global Biogeochem. Cy., 10, 603-628.

Friedl, M. A., McIver, D. K., Hodges, J. C. F., Zhang, X. Y., Muchoney, D., Strahler, A.

H., Woodcock, C. E., Gopal, S., Schneider, A., & Cooper, A. (2002), Global land

cover mapping from MODIS: Algorithms and early results, Remote Sens. Environ.,

183, 287-302.

GCOS 2nd Adequacy Report (2003). The Second Report on the Adequacy of the Global

Observing Systems for Climate in Support of the UNFCCC (2003).

Gobron, N., Pinty, B., Verstraete, M., & Govaerts, Y. (1999). MERIS Global Vegetation

Index (MGVI): Description and preliminary application. Int. J. Remote Sens., 20,

1917-1927.

Gobron, N., Pinty, B., Aussedat, O., Chen, J. M., Cohen, W. B., Fensholt, R., Gond, V.,

Huemmrich, K. F., Lavergne, T., Melin, F., Privette, J. L., Sandholt, I., Taberner, M.,

Turner, D. P., Verstraete, M. M., & Widlowski, J. (2006). Evaluation of fraction of

absorbed photosynthetically active radiation products for different canopy radiation

transfer regimes: Methodology and results using Joint Research Center products

derived from SeaWiFS against ground-based estimations. J. Geophys. Res., 111,

D13110, doi:10.1029/2005JD006511.

Page 160: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Goetz, S. J., Bunn, A. G., G. J. Fiske, & Houghton, R. A. (2005). Satellite-observed photosynthetic trends across boreal North America associated with climate and fire disturbance. Proc. Natl. Acad. Sci.

U.S.A., 102, 13521-13525.Goswami, B. N., Venugopal, V., Sengupta, D., Madhusoodanan, M. S., & Xavier, P. K.

(2006). Increasing trend of extreme rain events over India in a warming environment.

Science, 314, 1442-1445.

Gradshteyn, I. S. & Ryzhik, I. M. (1980). Table of Integrals, Series, and Products. New

York: Academic Press, pp. 1160.

Gu, L., Baldocchi, D., Verma, S. B., Black, T. A., Vesala, T., Falge, E. M., & Dowty, P.

R. (2002). Advantages of diffuse radiation for terrestrial ecosystem productivity. J.

Geophys. Res., 107, D6, 10.1029/2001JD001242.

Gu, G., Adler, R., Huffman, G. & Curtis, S. (2007). Tropical rainfall variability on

interannual-to-interdecadal/longer-time scales derived from the GPCP monthly

product, J. Climate, 20, 4033-4046.

Gutman, G. G. (1998). On the use of long-term global data of land reflectances and

vegetation indices derived from the advanced very high resolution radiometer. J.

Geophys. Res., 104, 6241-6256.

Hansen, J., Ruedy, R., Glascoe, J., & Sato, M. (1999). GISS analysis of surface

temperature change. J. Geophys. Res., 104, 30997-31022.

Hansen, M., DeFries, R. S., Townshend, J. R. G., Carroll, M., Dimiceli, C., & Sohlberg,

R. A. (2003). Global Percent Tree Cover at a Spatial Resolution of 500 Meters: First

Results of the MODIS Vegetation Continuous Fields Algorithm. Earth Inter., 7(10),

1-15.

Hartmann, D. L. (1994). Global Physical Climatology. pp. 411.

Herrmann, S. M., Anyamba A., & Tucker C. J. (2005). Recent trends in vegetation

dynamics in the African Sahel and their relationship to climate. Global Environ. Ch.,

15, 394-404.

Page 161: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Hickler, T., Eklundh, L., Seaquist, J. W., Smith, B., Ardö, J., Olsson, L., Sykes, M. T., &

Sjöström, M. (2005). Precipitation controls Sahel greening trend. Geophys. Res. Lett.,

32, 1-4.

Hogg, E. H., Price, D. T., & Black, T. A. (2000). Postulated feedbacks of deciduous

forest phenology on seasonal climate patterns in the western Canadian interior. J.

Climate, 13, 4229-4243.

Holben, B. (1986). Characteristics of maximum-value composite images from temporal

AVHRR data. Int. J. Remote Sens., 7, 1417-1434.

Houborg, R., Soegaard, H., & Boegh, E. (2007). Combining vegetation index and model

inversion methods for the extraction of key vegetation biophysical parameters using

Terra and Aqua MODIS reflectance data. Remote Sens. Environ., 106, 39-58.

Hu, J., Tan, B., Shabanov, N. V., Crean, K. A., Martonchik, J. V., Diner, D. J.,

Knyazikhin, Y., & Myneni, R. B. (2003). Performance of the MISR LAI and FPAR

Algorithm: A Case Study In Africa. Remote Sens. Environ., 88, 324-340.

Hu, J., Su, Y., Tan, B., Huang, D., Yang, W., Bull, M. A., Martonchik, J. V., Diner, D.J.,

Knyazikhin, Y., & Myneni, R. B. (2007). Analysis of the MISR LAI/FPAR product

for spatial and temporal coverage, accuracy and consistency. Remote Sens. Environ.,

107, 334-347.

Huang, D., Yang, W., Tan, B., Rautiainen, M., Zhang, P., Hu, J., Shabanov, N. V.,

Linder, S., Knyazikhin, Y., & Myneni, R. B. (2006). The importance of measurement

errors for deriving accurate reference leaf area index maps for validation of moderate-

resolution satellite LAI products. IEEE Trans. Geosci. Remote Sens., 44(7), 1866-

1871.

Huang, D., Knyazikhin, Y., Dickinson R. E., Rautiainen, M., Stenberg, P., Disney, M.,

Lewis, P., Cescatti, A., Tian, Y., Verhoef, W., Martonchik, J. V., & Myneni, R. B.

(2007). Canopy spectral invariants for remote sensing and model applications.

Remote Sens. Environ., 106, 106-122.

Huang, D., Knyazikhin, Y., Yang, W., Deering, D. W., Stenberg, P., Shabanov, N. V.,

Tan, B., & Myneni, R. B. (2008). Stochastic transport theory for investigating the

Page 162: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

three-dimensional canopy structure from space measurements. Remote Sens.

Environ., 112, 106-122.

Huete, A. R., Didan, K., Shimabukuro, Y. E., Ratana, P., Saleska, S. R., Hutyra, L. R.,

Yang, W., Nemani, R. R., & Myneni, R. (2006). Amazon rainforests green-up with

sunlight in dry season. Geophys. Res. Lett., 33, doi: 10.1029/2005gl025583.

Huffman, G.J., Adler, R. F., Bolvin, D. T., Gu, G., Nelkin, E. J., Bowman, K. P., Hong,

Y., Stocker, E. F., & Wolff, D. B. (2007). The TRMM multi-satellite precipitation

analysis: quasi-global, multi-year, combined-sensor precipitation estimates at fine

scale. J. Hydrometeorol., 8(1), 38-55.

Ichii, K., Kawabata, A., & Yamaguchi, Y.  (2002). Global correlation analysis for NDVI

and climatic variables and NDVI trends: 1982-1990. Int. J. Remote Sens., 23, 3873-

3878.

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. [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M.

Tignor and H. L. Miller (eds.)] (p. 996): Cambridge University Press, Cambridge,

United Kingdom and New York, NY, USA.

Jinjun, J. (1995). A climate-vegetation interaction model: simulating physical and

biological processes at the surface. J. Biogeogr., 22, 445-451.

Justice, C. O., Vermote, E., Townshend, J. R. G., Defries, R., Roy, D. P., Hall, D. K.,

Salomonson, V. V., Privette, J. L., Riggs, G., Strahler, A., Lucht, W., Myneni, R. B.,

Knyazikhin, Y., Running, S. W., Nemani, R. R., Wan, Z. M., Huete, A. R., van

Leeuwen, W., Wolfe, R. E., Giglio, L., Muller, J. P., Lewis, P., & Barnsley, M. J.

(1998). The Moderate Resolution Imaging Spectroradiometer (MODIS): Land remote

sensing for global change research. IEEE Trans. Geosci. Remote Sens., 36, 1228-

1249.

Justice, C., Belward, A., Morisette, J., Lewis, P., Privette, J., & Baret, F. (2000).

Developments in the ‘validation’ of satellite sensor products for the study of the land

surface. Int. J. Remote Sens., 21, 3383-3390.

Page 163: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Justice, C. O., Townshend, J. R. G., Vermote, E. F., Masuoka, E., Wolfe, R. E., Saleous,

N., Roy, D. P., & Morisette, J. T. (2002). An overview of MODIS Land data

processing and product status. Remote Sens. Environ., 83, 3-15.

Kawabata, A., Ichii, K., & Yamaguchi, Y.  (2001). Global monitoring of international

changes in vegetation activities using NDVI and its relationship to temperature and

precipitation. Int. J. Remote Sens., 22, 1377-1382.

Knorr, W., & Kattge, J. (2005). Inversion of terrestrial ecosystem model parameter values

against eddy covariance measurements by Monte Carlo sampling. Global Change

Biol., 11, 1333-1351.

Knyazikhin, Y., Martonchik, J. V., Myneni, R. B., Diner, D. J., & Running, S. W.

(1998a). Synergistic algorithm for estimating vegetation canopy leaf area index and

fraction of absorbed photosynthetically active radiation from MODIS and MISR data.

J. Geophys. Res., 103, 32257–32274.

Knyazikhin, Y., Kranigk, J., Myneni, R. B., Panfyorov, O., & Gravenhorst, G. (1998b).

Influence of small-scale structure on radiative transfer and photosynthesis in

vegetation cover. J. Geophys. Res., 103(D6), 6133-6144.

Knyazikhin, Y., & Marshak, A. (2000). Mathematical aspects of BRDF modeling:

Adjoint problem and Green’s function. Remote Sens. Rev., 18, 263-280.

Lambers, H., Chapin, F. S. I., & Pons, T. L. (2000). Plant Physiological Ecology.

Springer.

Lapenis, A., Shvidenko, A., Shepaschenko, D., Nilsson, S., & Aiyyer, A. (2005). Acclimation of Russian forests to recent changes in climate. Global Change Biol., 11, 2090-2102.

Lewis, P. (1999). Three-dimensional plant modelling for remote sensing simulation

studies using the Botanical Plant Modelling System. Agronomie, 19, 185-210.

Lewis, P., & Disney, M. (2007). Spectral invariants and scattering across multiple scales

from within-leaf to canopy. Remote Sens. Environ., 109, 196-206.

Li, X. W., & Strahler, A. H. (1986). Geometric-optical bidirectional reflectance modeling

of a conifer forest canopy. IEEE Trans. Geosci. Remote Sens., 24, 906-919.

Page 164: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Li, X. W., & Strahler, A. H. (1992). Geometric-optical bidirectional reflectance modeling

of the discrete crown vegetation canopy - effect of crown shape and mutual

shadowing. IEEE Trans. Geosci. Remote Sens., 30, 276-292.

Lotsch, A., Friedl, M. A., Anderson, B. T. & Tucker, C. J. (2003). Coupled vegetation-

precipitation variability observed from satellite and climate records. Geophys. Res.

Lett., 30(14). pp. 1774, doi:10.1029/2003GL017506.

Lu, L. X., & Shuttleworth, W. J. (2002). Incorporating NDVI-derived LAI into the

climate version of RAMS and its impact on regional climate. J. Hydrometeorol., 3,

347-362.

Malhi, Y., Roberts, J. T., Betts, R. A., Killeen, T. J., Li, W. H., & Nobre, C. A. (2008).

Climate change, deforestation, and the fate of the Amazon. Science, 319, 169-172.

Masson, V., Champeaux, J. L., Chauvin, F., Meriguer, C., & Lacaze, R. (2003). A global

database of land surface parameters at 1km resolution in meteorological and climate

models. J. Climate, 16, 1261-1282.

Melillo, J. M., McGuire, A. D., Kicklighter, D. W., Moore, B., Vorosmarty, C. J., &

Schloss, A. L. (1993). global climate-change and terrestrial net primary production.

Nature, 363, 234-240.

Min, Q. (2005). Impacts of aerosols and clouds on forest-atmosphere carbon exchange. J.

Geophys. Res., 110, D06203, doi:10.1029/2004JD004858.

Mintz, Y. (1984). The sensitivity of numerically simulated climates to land-surface

boundary conditions: Cambridge University Press.

Mitchell, T. D., Carter, T.R., Jones, P.D., Hulme, M., & New, M. (2003). A

comprehensive set of high-resolution grids of monthly climate for Europe and the

globe: the observed record (1901-2000) and 16 scenarios (2001-2100). J. Climate

(submitted).

Miura, T., Huete, A. R., & Yoshioka, H. (2000). Evaluation of sensor calibration

uncertainties on vegetation indices for MODIS. IEEE Trans. Geosci. Remote Sens.,

38(3), 1399-1409.

Page 165: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Monteith, J. L., & Unsworth, M. H. (1990). Principles of Environmental Physics: Edward

Arnold, London, pp. 291.

Morisette, J. T., Baret, F., Privette, J. L., Myneni, R. B., Nickeson, J. E., Garrigues, S., et

al. (2006). Validation of global moderate resolution LAI products: a framework

proposed within the CEOS Land Product Validation subgroup. IEEE Trans. Geosci.

Remote Sens., 44, 1804-1817.

Murphy, R. E. (2006). The NPOESS preparatory project. Earth Science Satellite Remote

Sensing, Springer Berlin Heidelberg, 182-198, doi:10.1007/978-3-540-37293-6.

Myneni, R. B., Ross, J., & Asrar, G. (1989). A review on the theory of photon transport

in leaf canopies. Agric. Forest Meteorol., 45, 1-153.

Myneni, R. B., & Williams, D. L. (1994). On the relationship between FPAR and NDVI.

Remote Sens. Environ., 49, 200-211.

Myneni, R. B., Keeling, C. D., Tucker, C. J., Asrar, G., & Nemani, R. R. (1997).

Increased plant growth in the northern high latitudes from 1981-1991. Nature, 386,

698-701.

Myneni, R. B., Hoffman, S., Knyazikhin, Y., Privette, J. L., Glassy, J., Tian, Y., Wang,

Y., Song, X., Zhang, Y., Smith, G. R., Lotsch, A., Friedl, M., Morisette, J. T., Votava,

P., Nemani, R. R., & Running, S. W. (2002). Global products of vegetation leaf area

and fraction absorbed PAR from year one of MODIS data. Remote Sens. Environ., 83,

214-231.

Myneni, R. B., Yang, W., Nemani, R. R., Huete, A. R., Dickinson, R. E., Knyazikhin, Y.,

Didan, K., Fu, R., Juarez, R. I. N., Saatchi, S. S., Hashimoto, H., Ichii, K., Shabanov,

N. V., Tan, B., Ratana, P., Privette, J. L., Morisette, J. T., Vermote, E. F., Roy, D. P.,

Wolfe, R. E., Friedl, M. A., Running, S. W., Votava, P., El-Saleous, N., Devadiga, S.,

Su, Y., & Salomonson, V. V. (2007). Large seasonal swings in leaf area of Amazon

rainforests. Proc. Natl. Acad. Sci. U.S.A., 104, 4820-4823.

Naraynamoorthy, A. (2002). Indian irrigation: Five decades of development. Water Res.

J., 212, 1-29.

Page 166: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Naraynamoorthy, A. (2007). Deceleration in agricultural growth. Econ. Polit. Week.,

42(25), 2375-2379.

Nemani, R. R., Keeling, C. D., Hashimoto, H., Jolly, W. M., Piper, S. C., Tucker, C. J.,

Myneni, R. B., & Running, S. W. (2003). Climate-driven increases in global

terrestrial net primary production from 1982 to 1999. Science, 300, 1560-1563.

Niemeijer, D., & Mazzucato, V. (2002). Soil degradation in the west African Sahel: How

serious is it? Environment, 44(2), 20-31.

NRC Decadal Survey (2007). Committee on Earth Science and Applications from Space:

A Community Assessment and Strategy for the Future. National Research Council,

Earth Science and Applications from Space: National Imperatives for the Next

Decade and Beyond, ISBN: 0-309-66900-6.

Pandya, M. R., Singh, R. P. & Dadhwal, V. K. (2004). A signal of increased vegetation

activity of India from 1981 to 2001 observed using satellite-derived fraction of

absorbed photosynthetically active radiation. Curr. Sci., 87, 1122-1126.

Panel on Climate Change Feedbacks, C. R. C., National Research Council (2003).

Terrestrial Hydrology and Vegetation Feedbacks. Understanding Climate Change

Feedbacks.

Parry, M. L., Canziani, O. F., Palutikof, J. P., van der Linden, P. J. & Hanson, C. E. Eds.

(2007). Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of

Working Group II to the FourthAssessment Report of the Intergovernmental Panel on

Climate Change, Cambridge University Press, Cambridge, UK, pp. 1000.

Pielke, R. A., Avissar, R., Raupach, M., Dolman, A. J., Zeng, X. B., & Denning, A. S.

(1998). Interactions between the atmosphere and terrestrial ecosystems: influence on

weather and climate. Global Change Biol., 4, 461-475.

Pinzon, J. E., Brown, M. E., & Tucker, C. J. (2005). Satellite time series correction of

orbital drift artifacts using empirical mode decomposition. In N. Huang (ed.), Hilbert-

Huang Transform: introduction and applications, chapter 10, part II. applications.

Pitman, A. J. (2003). The evolution of, and revolution in, land surface schemes designed

for climate models. Int. J. Climatol., 23, 479-510.

Page 167: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Plummer, S., Arino, O., Simon, W., & Steffen, W. (2006). Establishing an Earth

observation product service for the terrestrial carbon community: the

GLOBCARBON initiative: Mitigation and Adaptation Strategies for Global Change,

11, 97-111.

Potter, C. S., Randerson, J. T., Field, C. B., Matson, P. A., Vitousek, P. M., Mooney, H.

A., & Klooster, S. A. (1993). Terrestrial ecosystem production: A process model

based on global satellite and surface data. Global Biogeochem. Cy., 7(4), 811-841.

Prince, S. D., Wessels, K. J., Tucker, C. J., & Nicholson, S. E. (2007). Desertification in

the Sahel: a reinterpretation of a reinterpretation. Global Change Biol., 13, 1308-

1313.

Reij, C., Tappan, G. & Belemvire, A. (2005). Changing land management practices and

vegetation on the Central Plateau of Burkina Faso (1968-2002). J. Arid Environ., 63,

642-659.

Ropelewski, C. F., & Halpert, M. S. (1987). Global and regional scale precipitation

pattern associated with El Nino/Southern Oscillation, Mon. Weather Rev., 115, 1606-

1626.

Ross, J. K., & Marshak, A. L. (1988). Calculation of canopy bidirectional reflectance

using the monte-carlo method. Remote Sens. Environ., 24, 213-225.

Running, S. W., Peterson, D. L., Spanner, M. A., & Teuber, K. B. (1986). Remote-

sensing of coniferous forest leaf-area. Ecology, 67, 273-276.

Running, S. W., & Gower, S. T. (1991). Forest-BGC, a general-model of forest

ecosystem processes for regional applications .2. dynamic carbon allocation and

nitrogen budgets. Tree Physiol., 9, 147-160.

Sankaran, M., Hanan, N. P., Scholes, R. J., Ratnam, J., Augustine, D. J., Cade, B. S.,

Gignoux, J., Higgins, S. I., Le Roux, X., Ludwig, F., Ardo, J., Banyikwa, F., Bronn,

A., Bucini, G., Caylor, K. K., Coughenour, M. B., Diouf, A., Ekaya, W., Feral, C. J.,

February, E. C., Frost, P. G. H., Hiernaux, P., Hrabar, H., Metzger, K. L., Prins, H. H.

T., Ringrose, S., Sea, W., Tews, J., Worden, J., & Zambatis, N. (2005). Determinants

of woody cover in African savannas. Nature, 438, 846-849.

Page 168: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Scholze, M., Knorr, W., Arnell, N. W., & Prentice, I. C. (2006). A climate-change risk

analysis for world ecosystems. Proc. Natl. Acad. Sci. U.S.A., 103, 13116-13120.

Schwartz, M. D., & Karl, T. R. (1990). Spring phenology - natures experiment to detect

the effect of green-up on surface maximum temperatures. Mon. Wea. Rev., 118, 883-

890.

Schwartz, M. D. (1992). Phenology and springtime surface-layer change. Mon. Wea.

Rev., 120, 2570-2578.

Schwartz, M. D. (1996). Examining the spring discontinuity in daily temperature ranges.

J. Climate, 9, 803-808.

Schwartz, M. D. (1999). Advancing to full bloom: planning phenological research for the

21st century. Int. J. Biometeorol., 42, 113-118.

Seaquist, J. W., Olsson, L., Ardo, J., & Eklundh, L. (2006). Broad-scale increase in NPP

quantified for the African Sahel, 1982-1999. Int. J. Remote Sens., 27, 5115-5122.

Sellers, P. J., Mintz, Y., Sud, Y. C., & Dalcher, A. (1986). A simple biosphere model

(sib) for use within general-circulation models. Journal of the Atm. Sci., 43, 505-531.

Sellers, P., Hall, F., Margolis, H., Kelly, B., Baldocchi, D., Denhartog, G., Cihlar, J.,

Ryan, M. G., Goodison, B., Crill, P., Ranson, K. J., Lettenmaier, D., & Wickland, D.

E. (1995). The boreal ecosystem-atmosphere study (BOREAS) - an overview and

early results from the 1994 field year. Bull. Am. Meteorol. Soc., 76, 1549-1577.

Sellers, P. J., Randall, D. A., Collatz, G. J., Berry, J. A., Field, C. B., Dazlich, D. A.,

Zhang, C., Collelo, G. D., & Bounoua, L. (1996). A revised land surface

parameterization (SiB2) for atmosphere GCMs. Part II : The generation of global

fields of terrestrial biophysical parameters from satellite data. J. Climate, 9, 706-737.

Sellers, P. J., Dickinson, R. E., Randall, D. A., Betts, A. K., Hall, F. G., Berry, J. A.,

Collatz, G. J., Denning, A. S., Mooney, H. A., Nobre, C. A., Sato, N., Field, C. B., &

Henderson, A. (1997). Modeling the exchanges of energy, water, and carbon between

continents and the atmosphere. Science, 275, 502-509.

Shabanov, N. V., Knyazikhin, Y., Baret, F., & Myneni, R. B. (2000). Stochastic

modeling of radiation regime in discontinuous vegetation canopy. Remote Sens.

Page 169: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Environ., 74, 125-144.

Shabanov, N. V., Huang, D., Yang, W., Tan, B., Knyazikhin, Y., Myneni, R. B., Ahl, D.

E., Gower, S. T., Huete, A. R., Eduardo, L., Aragão, O. C., & Shimabukuro, Y. E.

(2005). Optimization of the MODIS LAI and FPAR algorithm performance over

broadleaf forests. IEEE Trans. Geosci. Remote Sens., 43, 1855-1865.

Shabanov, N. V, Huang, D., Knyazikhin, Y., Dikinson, R. E., & Myneni, R. B. (2007).

Stochastic radiative transfer model for mixture of discontinuous vegetation canopies.

J. Quant. Spec. & Rad. Trans., 107, 236-262.

Shah, T. (2005). Groundwater and human development: challenges and opportunities in

livelihoods and environment. Water Sci. Tech., 51, 27-37.

Slayback, D. A., Pinzon, J. E., Los, S. O., & Tucker, C. J. (2003). Northern hemisphere

photosynthetic trends 1982-1999. Global Change Biol., 9, 1-15.

Smolander, S., & Stenberg, P. (2003). A method to account for shoot scale clumping in

coniferous canopy reflectance models. Remote Sens. Environ., 88, 363-373.

Smolander, S., & Stenberg, P. (2005). Simple parameterizations for the radiation budget

of uniform broadleaved and coniferous canopies. Remote Sens. Environ., 94, 355-363.

Soja, A. J., Tchebakova, N. M., French, N. H. F., Flannigan, M. D., Shugart, H. H.,

Stocks, B. J., Sukhinin, A. I., Parfenova, E. I., Chapin III, F. S., & Stackhouse Jr., P.

W. (2007). Climate-induced boreal forest change: Predictions versus current

observations. Global Planet. Ch., 56, 274-296.

Steltzer, H., & Welker, J. M. (2006). Modeling the effect of photosynthetic vegetation

properties on the NDVI-LAI relationship. Ecology, 87(11), 2765-2772

Stige, L. C., Stave, J., Chan, K. S., Ciannelli, L., Pettorelli, N., Glantz, M., Herren, H. R.,

& Stenseth, N. C. (2006). The effect of climate variation on agro-pastoral production

in Africa. Proc. Natl. Acad. Sci. U.S.A., 103, 3049-3053.

Sud, Y. C., Shukla, J., & Mintz, Y. (1988). Influence of land surface-roughness on

atmospheric circulation and precipitation - a sensitivity study with a general-

circulation model. J. Appl. Meteorol., 27, 1036-1054.

Page 170: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Tan, B., Hu, J., Zhang, P., Huang, D., Shabanov, N. V., Weiss, M., Knyazikhin, Y., &

Myneni, R. B. (2005). Validation of MODIS LAI product in croplands of Alpilles,

France. J. Geophys. Res., 110, D01107, doi:10.1029/2004JD004860.

Tan, B., Woodcock, C. E., Hu, J., Zhang, P., Ozdogan, M., Huang, D., Yang, W.,

Knyazikhin, Y., Myneni, R. B. (2006). The impact of gridding artifacts on the local

spatial properties of MODIS data: Implications for validation, compositing, and band-

to-band registration across resolutions. Remote Sens. Environ., 105, 98-114.

Tappan, G., & McGahuey, M. (2007). Tracking environmental dynamics and agricultural

intensification in southern Mali. Agri. Sys., 94, 38-51.

Tape, K., Sturm, M., & Racine, C. (2006). The evidence for shrub expansion in northern Alaska and the pan-Arctic. Global Change Biol., 12, 686–702.

Tarnavsky, E. V., Garrigues, S., & Brown, M. E. (2008). Multiscale geostatistical

analysis of AVHRR, SPOT-VGT, and MODIS global NDVI products. Remote Sens.

Environ., 112, 535-549.

Thompson, D. W. J., & Wallace, J. M. (1998). The Arctic oscillation signature in the

wintertime geopotential height and temperature fields. Geophys. Res. Lett., 25, 1297-

1300.

Tian, Y., Wang, Y., Zhang, Y., Knyazikhin, Y., Bogaert, J., & Myneni, R. B. (2002).

Radiative transfer based scaling of LAI/FPAR retrievals from reflectance data of

different resolutions. Remote Sens. Environ., 84, 143-159.

Tian, Y., Dickinson, R. E., Zhou, L., Zeng, X., Dai, Y., Myneni, R. B., Knyazikhin, Y.,

Zhang, X., Friedl, M., Yu, H., Wu, W., & Shaikh, M. (2004). Comparison of seasonal

and spatial variations of LAI/FPAR from MODIS and Common Land Model. J.

Geophys. Res., 109, D01103, doi:10.1029/2003JD003777.

Tikhonov, A. N., Goncharsky, A. V., Stepanov, V. V., & Yagola, A. G. (1995).

Numerical Methods for the Solution of Ill-Posed Problems, Kluwer Academic

Publisher, Boston, pp. 3.

Page 171: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Tucker, C. J., Fung, I. Y., Keeling, C. D., & Gammon, R. H. (1986). Relationship

between atmospheric CO2 variations and a satellite-derived vegetation index. Nature,

319, 195-199.

Tucker, C. J., & Nicholson, S. E. (1999). Variations in the size of the Sahara Desert from

1980 to 1997. Ambio, 28, 587-591.

Tucker, C. J., Pinzon, J. E., Brown, M. E., Slayback, D. A., Pak, E. W., Mahoney, R.,

Vermote, E. F., & El Saleous, N. (2005). An extended AVHRR 8-km NDVI dataset

compatible with MODIS and SPOT vegetation NDVI data. Int. J. Remote Sens., 26,

4485-4498.

Van Leeuwen., W. J. D., Orr, B. J., Marsh, S. E., & Herrmann, S. M. (2006). Multi-

sensor NDVI data continuity: Uncertainties and implications for vegetation

monitoring applications. Remote Sens. Environ., 100, 67-81.

Vermote, E. F., & Saleous, N. Z. (2006). Calibration of NOAA16 AVHRR over a desert

site using MODIS data. Remote Sens. Environ., 105, 214-220.

Wang, Q., Adiku, S., Tenhunen, J., & Granier, A. (2004). On the relationship of NDVI

with leaf area index in a deciduous forest site. Remote Sens. Environ., 94, 244-255.

Wang, Y., Tian, Y., Zhang, Y., El-Saleous, N., Knyazikhin, Y., Vermote, E., & Myneni,

R. B. (2001). Investigation of product accuracy as a function of input and model

uncertainties: Case study with SeaWiFS and MODIS LAI/FPAR Algorithm. Remote

Sens. Environ., 78, 296-311.

Weiss, M., Baret, F., Leroy, M., Hautecoeur, O., Bacour, C., Prevot, L., & Bruguier, N.

(2002). Validation of neural net techniques to estimate canopy biophysical variables

from remote sensing data. Agronomie, 22, 547-553.

Weiss, M., Baret, F., Garrigues, S., & Lacaze, R. (2007). LAI and fAPAR CYCLOPES

global products derived from VEGETATION. Part 2: validation and comparison with

MODIS collection 4 products. Remote Sens. Environ., 110, 317-333.

Welles, J. M., & Norman, J. M. (1991). Instrument for indirect measurement of canopy

architecture. Agronomie, 83, 818-825.

Page 172: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Wentz, F. J., Ricciardulli, L., Hilburn, K., & Mears, C. (2007). How much more rain will global warming bring? Science, 317, 233-235.

Wilmking, M., Juday, G. P., Barber, V. A., & Zald, H. S. J. (2004). Recent climate warming forces contrasting growth responses of white spruce at treeline in Alaska through temperature thresholds. Global Change Biol., 10, 1724-1736.

Xiao, J., & Moody, A. (2005). Geographical distribution of global greening trends and

their climatic correlates: 1982-1998. Int. J. Remote Sens., 11, 2371-2390.

Yang, W., Shabanov, N. V., Huang, D., Wang, W., Dickinson, R. E., Nemani, R. R.,

Knyazikhin, Y., & Myneni, R. B. (2006a). Analysis of leaf area index product from

combination of MODIS and Aqua data, Remote Sens. Environ., 104, 297-312.

Yang, W., Tan, B., Huang, D., Rautiainen, M., Shabanov, N. V., Wang, Y., Privette, J. L., Huemmrich, K. F., Fensholt, R., Sandholt, I., Weiss, M., Ahl, D. E., Gower, S. T., Nemani, R. R., Knyazikhin, Y., & Myneni, R. B. (2006b). MODIS leaf area index Products: From validation to algorithm improvement, IEEE Trans. Geosci. Remote Sens., 44, 1885-1898.

Zhang, X., Zwiers, F. W., Hegerl, G. C., Lambert, F. H., Gillett, N. P., Solomon, S., Stott,

P. A., & Nozawa, T. (2007). Detection of human influence on twentieth-century

precipitation trends. Nature, 448, 461-465.

Zhou, L., Tucker, C. J., Kaufmann, R. K., Slayback, D., Shabanov, N. V., & Myneni, R.

B. (2001). Variations in northern vegetation activity inferred from satellite data of

vegetation index during 1981 to 1999. J. Geophys. Res., 106, 20069-20083.

Page 173: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

Curriculum Vitae

Sangram Ganguly

Department of Geography and Environment, Boston University675 Commonwealth Avenue, Boston, MA 02215

Phone: (617) 353-8846 (office), (617)319-6249 (cell)Fax: (617) 353-8399

E-mail: [email protected] page: http://www.sangramganguly.com

EDUCATION

Ph.D. in Geography, major in Remote Sensing, 05/2008 expectedDepartment of Geography and Environment, Boston UniversityDissertation: Generating LAI Earth System Data Record from Multiple Sensors.Advisor: Ranga B. Myneni, ProfessorM.Sc in Exploration Geophysics, 05/2004Department of Exploration Geophysics, Indian Institute of Technology, Kharagpur, India.Thesis: Studies on Texture Analysis and Image Understanding systems for Remotely Sensed scenes and their applications to Geological Studies.Advisors: A. K. Ray (Professor) and A. Chakrabarti (Professor).B.Sc (Honors) in Exploration Geophysics, 05/2004Department of Exploration Geophysics, Indian Institute of Technology, Kharagpur, India.

SPECIALIZATION

a) Remote sensing of vegetation and modeling 3-D radiative transfer processes in vegetation canopies.b) Advanced signal processing and image processing techniques for multi-dimensional and multi-temporal analysis of satellite imagery. c) Modeling of large-scale processes of climate-vegetation interaction. Statistical time series analysis and spectral analysis of long term vegetation and climate data records.d) Image classification, compression and recognition using soft computing techniques (Fuzzy, Neural nets, genetic algorithm) and wavelet analysis. e) InSAR and SAR change detection algorithms. f) Programming expertise: C/C++, Fortran, MATLAB, IDL, HTML, DHTML. Technical software package tools: ENVI, Arcmap, RATS, IPW. Operating system/platforms: Windows, Linux, HP Unix, Sun Solaris workstation, Mac.

Page 174: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

RESEARCH EXPERIENCE

Doctoral Research Assistant, Department of Geography and Environment, Boston University, Boston, MA, USA, 2004 – 2008.- Developing physical algorithms (Canopy spectral invariants and stochastic radiative transfer models) to retrieve vegetation structural parameters;- Generating long term data records of vegetation biophysical products (LAI, FPAR) from measurements of multiple sensors (AVHRR and MODIS); - Analyzing and evaluating global trends in vegetation activity over the period ,1981-present, and correlating the trends with observational changes in climate surrogates;- Studying the recent climate changes and their impacts on large-scale vegetation dynamics over the north American/ Eurasian grasslands and boreal forests and the semi-arid regions of Africa, Australia and Southern Asia;- Studying impacts of recent climate and land use change on vegetation in the Indian sub-continent.

Masters Research Assistant, Department of Exploration Geophysics, Indian Institute of Technology, Kharagpur, India, 2002-2004.- Studies on Texture Analysis and Image Understanding systems for Remotely Sensed scenes and their applications to Geological Studies.

Research Internee, Cooperative Research Centre for Sensor Signal and Information Processing (CSSIP) and the Department of Electrical and Electronics Engineering, University of Adelaide, Australia, May-July, 2003.- Repeat Pass SAR Change Detection and Image Processing techniques.

Research Internee, Department of Information Engineering, University of Pisa, Italy, May-July, 2002.- A Study of the Alboran Sea mesoscale system using the empirical orthogonal function decomposition of satellite SST data and interpreting the sea wave direction propagation and its effect by other data processing techniques.

ACADEMIC SERVICE

- Reviewer of Remote Sensing of Environment- Reviewer of IEEE Geoscience and Remote Sensing Letters- Reviewer of IEEE Transactions in Geoscience and Remote Sensing- Reviewer of Climate Change - Reviewer of Climate Research

Page 175: Chapter 2 - Boston Universitycliveg.bu.edu/download/thdis/sganguly.PHD.doc · Web viewPixels with the same NDVI are located on a straight line (red line in Fig. 2.8). This line intersects

PUBLICATIONS

Ganguly, S., Schull, M., Samanta, A., Shabanov, N. V., Milesi, C., Nemani, R. R., Knyazikhin, Y., and Myneni, R. B. (2008). Generating LAI Earth system data record from multiple sensors. Part 1: Theory. Remote Sensing of Environment (submitted).

Ganguly, S., Samanta, A., Schull, M., Shabanov, N. V., Milesi, C., Nemani, R. R., Knyazikhin, Y., and Myneni, R. B. (2008). Generating LAI Earth system data record from multiple sensors. Part 2: Implementation, analysis and validation. Remote Sensing of Environment (submitted).

Schull, M.A., Ganguly, S., Samanta, A., Huang, D., Jenkins, J. P., Shabanov, N. V., Chiu, J. C., Marshak, A., Blair, J. B., Myneni, R. B., Kynazikhin, Y. (2007). Physical interpretation of the correlation between multi-angle spectral data and canopy height, Geophysical research letters, 34, doi:10. 1029/2007GL031143.

PRESENTATIONS AT SELECT MEETINGS

Ganguly, S., Huang, D., Yang, W., Shabanov, N. V., Knyazikhin, Y., and Myneni, R. B., Global data set of LAI and FPAR from AVHRR and MODIS data, Poster Presentation at 2005 AGU Fall Meeting, 2005, San Francisco, CA, USA.

Ganguly, S., Schull, M., Samanta, A., Shabanov, N. V., Huang, D., Knyazikhin, Y., and Myneni, R. B., Physically Based Algorithm for Generating Long Term AVHRR LAI and FPAR Data Records by Joining GIMMS AVHRR and MODIS, Proceedings of the International Geoscience and Remote Sensing Symposium (IEEE IGARSS), 2007, Barcelona, Spain.

Ganguly, S., Schull, M., Samanta, A., Myneni, R. B., and Knyazikhin, Y. (2008). Retrieving canopy structure across multiple sensors. NASA VEG3D & BIOMASS Workshop, March 3-5, 2008, Charlottesville, VA.

Ganguly, S., Repeat pass SAR change detection using wavelet decomposition techniques and adaptive eigen mode reconstruction, Society for Petroleum Geophysicists and Society for Exploration Geophysics (SEG) international conference, 2004, Hyderabad, India (Awarded the Gold medal for the best technical poster).

Ganguly, S., Samanta, A., Ray, A. K., and Chakraborty, A., 2D Fuzzy Gabor filters for Texture classification, ISFUMIP National Conference On Fuzzy Logic and its Application in Technology and Management FLATeM - 2004, Vinod Gupta School of Management, IIT Kharagpur, 2004, Kharagpur, India.