snow and ice cover changes in the … degrassi...chapter 8. spectral analysis and...

163
SNOW AND ICE COVER CHANGES IN THE GLACIERS OF THE CORDILLERA REAL (BOLIVIA) 1987-2001: A REMOTE SENSING STUDY by CHRISTIAN DEGRASSI A thesis submitted in partial fulfillment of the requirements for the degree of Masters of Arts in Geography Appalachian State University May 2004

Upload: dophuc

Post on 27-Jun-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

SNOW AND ICE COVER CHANGES IN THE GLACIERS OF THE CORDILLERA REAL (BOLIVIA) 1987-2001:

A REMOTE SENSING STUDY

by

CHRISTIAN DEGRASSI

A thesis submitted in partial fulfillment of the requirements for the degree of

Masters of Arts in Geography

Appalachian State University

May 2004

ii

SNOW AND ICE COVER CHANGES IN THE GLACIERS OF THE CORDILLERA REAL (BOLIVIA) 1987-2001:

A REMOTE SENSING STUDY

A Thesis

By

CHRISTIAN DEGRASSI

May 2004

APPROVED BY: ______________________________ Michael Mayfield Chairperson, Thesis Committee ______________________________ James Young Member, Thesis Committee ______________________________ Jeffrey Colby Member, Thesis Committee ______________________________ James Young Chairperson, Department of Geography and Planning ______________________________ Judith E. Domer Dean, Graduate Studies and Research

iii

Copyright by Christian Degrassi 2004. All Right reserved.

iv

Abstract

“SNOW AND ICE COVER CHANGES IN THE GLACIERS OF THE

CORDILLERA REAL (BOLIVIA) 1987-2001:

A REMOTE SENSING STUDY.” (May 2004)

Christian Degrassi, B.S., University of Trieste, Italy

M.A., Appalachian State University

Thesis Chairperson: Dr. Michael Mayfield

Fluctuations of mountain glaciers and ice caps are key variables for the

observation of climate-related changes and the detection of enhanced greenhouse effects

(IPCC 2001). Due to the narrow range of climate conditions under which glaciers and ice

caps develop, monitoring the distribution, intensity, and rate of change of ice masses is of

utmost importance in delineating spatial and temporal variations of the Earth’s climatic

system.

The purpose of this research was to analyze snow and ice cover changes of the

glaciers of the Cordillera Real (Bolivia) during the period 1987-2001. The study was

based on satellite remote-sensing technology, and sought to enhance the regional

understanding of trends and dynamics of the glacier retreat observed on benchmark

glaciers. As a result of the research produced in this thesis an estimate was made of the

snow and ice cover changes affecting the glaciers of the Cordillera Real. The results are

in line with field observations of a gradual but quantifiable retreat from 1987. The snow

v

and ice cover of the Cordillera Real were found to be highly sensitive to changes in

regional weather patterns such as those induced by El Niño events, but may also be

related to long-term global change processes in the troposphere such as the increasing

planetary temperature. This study confirmed that, while not developed for glaciological

applications, Landsat platforms are a valuable resource to study snow and ice cover, but

present serious limitations in snow-covered tropical alpine environments.

vi

Acknowledgments

I would like to thank my thesis chair, Dr. Michael Mayfield, from whom I have

learned so many great things. I would like to thank him for helping and supporting this

research project, and for his endless patience. Also, I would like to thank my committee

members, Dr. James Young and Dr. Jeffrey Colby for their insight, guidance and support

throughout my graduate career and the thesis process; I must thank Mr. Baker Perry for

his assistance during the data-mining process.

Within the wonderful faculty of the Department of Geography and Planning, I

have particular thanks for Dr. Richard Crepeau, for being always available for discussion

and for helping solving many problems, for Dr. Kathleen Schroeder, without whom I

would not have been able to make it to the end, and Mr. Arthur Rex, for being always

available.

I must thank Dr. Edelma Huntley, Dr. Judith Domer, and the Graduate School for

supporting my research and providing a thesis research grant. A thank you also to the

GSAS for the research award.

I will never been able to thank enough my wife Allyson for being always close to

me, and for always letting me pursue all my dreams.

vii

Dedication

With Love, to My Wife Allyson.

viii

Table of contents List of tables ..................................................................................................................... xi

List of figures ................................................................................................................. xiii

Chapter 1. Introduction ....................................................................................................... 1

Chapter 2. The Study of Glaciers: Methodology Overview ............................................... 3

On-site Measurements .................................................................................................... 5

Indirect Measurements.................................................................................................... 7

Coring ........................................................................................................................... 11

Chapter 3. Study Area....................................................................................................... 14

Exploration and Historical Data.................................................................................... 14

Area of Interest ............................................................................................................. 16

Chapter 4. Climatic Conditions Affecting Tropical Glaciers of Bolivia .......................... 21

Chapter 5. Remotely Sensed Change Detection of Snow and Ice .................................... 26

Supervised and Unsupervised Classifications .............................................................. 29

Band Ratioing ............................................................................................................... 30

Spectral Un-mixing....................................................................................................... 32

Chapter 6. Spectral environment....................................................................................... 34

Spectral Characteristics of Snow and Ice...................................................................... 34

Landsat platforms.......................................................................................................... 42

ix

TM and ETM+ spectrometers and sensors calibration ................................................. 43

Use of Landsat TM and ETM+ for Snow and Ice Spectrometry.................................. 52

Chapter 7. Preprocessing .................................................................................................. 59

Data Characteristics ...................................................................................................... 59

Geometric Correction and Co-registration.................................................................... 63

Study Area Subset......................................................................................................... 66

Atmospheric Correction................................................................................................ 69

Image Standardization .................................................................................................. 81

Chapter 8. Spectral Analysis and Classification ............................................................... 85

Spectral Characteristics of the Subscenes..................................................................... 85

Spectral Endmembers Definition.................................................................................. 88

Binary classification...................................................................................................... 90

Supervised Classification: Minimum Distance............................................................. 92

Spectral unmixing ......................................................................................................... 94

Chapter 9. Results ............................................................................................................. 97

Binary Classification: Zonal Statistics.......................................................................... 97

Minimum Distance: Zonal Statistics........................................................................... 100

Binary vs. Minimum Distance Classifier.................................................................... 109

Regional vs. Local Observations ................................................................................ 110

Chapter 10. Discussion and Conclusion ......................................................................... 114

Discussion................................................................................................................... 114

Technical, and Physical Limitations of the Results .................................................... 118

Conclusions................................................................................................................. 121

x

Bibliography ................................................................................................................... 124

References Cited ......................................................................................................... 124

References Consulted.................................................................................................. 131

Appendix A. Anisotropic Reflectance Correction and Topographic Normalization of

rugged terrains and snow covered surfaces. ................................................................... 132

Concepts of Anisotropic Reflectance and Topographic normalization ...................... 132

Topographic Normalization: Methods and Techniques.............................................. 133

Case Study: Image Processing .................................................................................... 136

Case Study: Results and Discussion ........................................................................... 139

Limitations and Final Consideration........................................................................... 144

Biographical information .................................................................................................146

xi

List of tables Table 3.1. Glaciers of the Bolivian Andes as November, 1st 1984 (modified after Jordan

1991). .................................................................................................................... 18

Table 6.1. Spectral and spatial specifications for TM and ETM+ spectrometers (NASA 2003). .................................................................................................................... 43

Table 6.2. LTAP gain rules for Landsat ETM+ (modified after NASA 2003). .............. 47

Table 6.3. ETM+ Radiance Dynamic Range (wm-2sr-1µm-1) (modified after NASA 2003). .................................................................................................................... 48

Table 6.4. ETM+ low-gain and TM Radiance Dynamic Range (w•m-2•sr-1•µm-1) (modified after NASA 2003). ............................................................................... 51

Table 7.1. Constraints applied to the selection of suitable Landsat scenes. ..................... 60

Table 7.2 Metadata summary for the selected Landsat scenes. ........................................ 61

Table 7.3. Solar energy path (modified after Schott 1997)............................................... 70

Table 7.4. Manually selected DOS values based on the analysis of dark targets present in the five scenes. The values are represented as digital numbers. ........................... 78

Table 7.5. Atmospheric scattering models (after Chavez 1988)....................................... 79

Table 7.6. Scattering function coefficients obtained for each band in different atmospheric conditions (methodology after Chavez 1988). ................................. 79

Table 7.7. Procedure to predict scattering from TM band 1 reference DOS (methodology after Chavez 1988)................................................................................................ 80

Table 7.8. Observed DOS and final predicted DOS values.............................................. 81

Table 7.9. Post-calibration parameters (after Irish 1998). ................................................ 83

Table 7.10. Mean Solar Exo-atmospheric irradiances (Wcm-2µm-1) (values after Irish 1998). .................................................................................................................... 84

Table 7.11. Earth-Sun distance (astronomic units AU) and Sun elevation (degrees) at the time of acquisition (distance d calculated after Irish 1998). ................................. 84

xii

Table 8.1. Maximum in-scene reflectance registered by the TM and ETM+ sensors for the five subscenes. Note: Reflectance values greater than 100% are possible due to path radiance and background effect illumination............................................ 86

Table 8.2. Percent saturated pixels for each band of the subscenes. ................................ 87

Table 9.1. Snow/ice area measured from the zonal statistics of the binary classification. ............................................................................................................................... 98

Table 9.2. Inter-scene zonal statistics illustrating gain, losses and total net balance. ...... 99

Table 9.3. Inter-scene snow/ice cover percent changes. ................................................... 99

Table 9.4. Snow/ice area measured from the zonal statistics of the minimum distance classification. .......................................................................................................100

Table 9.5. Inter-scene zonal statistics from minimum distance statistics illustrating gain, losses and total net balance. ................................................................................ 102

Table 9.6. Inter-scene snow/ice cover percent changes determined from minimum distance classification. ........................................................................................ 102

Table 9.7a. Change detection statistics from 1987-1989................................................ 103

Table 9.7b. Change detection statistics from 1989-1997................................................ 104

Table 9.7c. Change detection statistics from 1977-2000................................................ 105

Table 9.7d. Change detection statistics from 2000-2001................................................ 106

Table 9.8. Detail of class areal changes (sq. km.)........................................................... 107

Table 10.1. Inter-scene snow/ice cover percent loss....................................................... 116

Table 10.2. Inter-scene snow and ice cover change estimates for Chacaltaya Glacier. . 117

Table 10.3. Inter-scene snow and ice cover change estimates for Zongo Glacier.......... 118

Table 10.4. Inter-scene snow and ice cover change estimates for Laguna Glaciar. ....... 118

Table A.1. Rugged terrain subset.................................................................................... 132

Table A.2. Ancohuma-Illampu glaciers.......................................................................... 132

xiii

List of figures Figure 2.1. Schematic representation of the profile of a glacier. Net balance is positive in

the accumulation zone and negative in the ablation zone. Ice movements transfer mass from the accumulation to the ablation zone constantly reshaping the profile of the glacier (modified after Bennett and Glasser 1996)....................................... 4

Figure 2.2. Ice sample extracted from the coring tool by scientist of Thompson’s research group studying the retreating glaciers of the Kilimanjaro (Tanzania) (NOAA 2002). .................................................................................................................... 12

Figure 3.1. Spatial profile of the Andes in proximity of the study area (elevation data from EDC 2003). (Vertical scale is exaggerated)................................................. 17

Figure 3.2. Bolivia and relative location of the area of study area of study (datasets from EDC 2003). ........................................................................................................... 19

Figure 3.3. Area of study. Landsat image of the Cordillera Real. .................................... 20

Figure 4.1. During La Niña years, the wind pattern is dominated by easterlies rich in moisture collected over the Amazon basin. Easterlies favor snow precipitation and accumulation during the austral summer months. During El Niño events, low moisture westerlies are predominant; the rain shadow effect of the Andean coastal ranges reduces precipitation (Francou, Vuille, Wagnon, Mendoza and Sicart 2003). .................................................................................................................... 25

Figure 6.1. Physical states of H2O (after Martin 2004).................................................... 35

Figure 6.2. Spectral reflectance of snow. Each curve represents the spectrum for a different grain size. The dashed line indicates the center (1.03 µm) of a characteristic absorption band extending from 0.96 µm to 1.08 µm (modified after Nolin and Dozier 2000). ....................................................................................... 38

Figure 6.3. Modeled spectral reflectance of glacier facies (modified after Hall et al. 1988, Zheng et al. 1984). ................................................................................................ 39

Figure 6.4. Mie Scattering (modified after Nave 2000).................................................... 41

Figure 6.5. Relative spectral responses and spectral differences between TM (grey) and ETM+ (black) spectrometers (modified after NASA 2003). ................................ 45

xiv

Figure 6.6. LTAP gain rules for the month of august. The rules depend on expected landcover, seasonality, and illumination. (after NASA 2003).............................. 47

Figure 6.7. Optimization of radiometric resolution based on ETM+ Radiance Dynamic Range adjustments (after NASA 2003). ............................................................... 48

Figure 6.8. Modeled maximum reflectance detected by Landsat 7 ETM+ P01R71 06-26-00 for low- gain mode and high-gain mode.......................................................... 50

Figure 6.9. ETM+ low-gain and TM Radiance Dynamic Range (w•m-2•sr-1•µm-1) (modified after NASA 2003). ............................................................................... 51

Figure 6.10. Maximum reflectance detected by Landsat 5 TM P01R71 06-10-97 (L5) and Landsat 7 ETM+ P1R71 06-26-00 (L7). The ETM+ is preset by LTAP on high-gain mode (HHHHHH) (modified after NASA 2003). ........................................ 52

Figure 6.11. Modeled Directional-Hemispherical reflectance of deep snow (modified after Nolin and Dozier 2000). TM bandpasses are represented by light gray boxes overlapping the spectra. “r” represents grain size................................................. 53

Figure 6.12. Modeled Directional Hemispherical reflectance of glacier facies in TM bandpasses for the visible and NIR domains (modified after Hall et al. 1988, Zheng et al. 1984). ................................................................................................ 54

Figure 6.13. Integrated spectral directional-hemispherical reflectance of deep snow and glacial facies in TM bandpasses (modified after Zheng et al. 1984, Dozier and Mark 1987, Hall et al. 1988, Rosenthal et al. 1996). TM bandpasses are represented by light gray boxes overlapping the spectra. “r” represents snow grain size. ........................................................................................................................55

Figure 6.14. TM and ETM+ reflectance cut-offs (left) and resulting snow and ice signatures from modeled directional hemispherical reflectance (right). Refer to Figure 6.13 for snow and ice classes legend......................................................... 57

Figure 7.1 Relative position of the Cordillera Real within a Path01-Row71 Landsat scene. Approximate coordinates of the scene center are latitude 15.89S, longitude 67.99W...................................................................................................................59

Figure 7.2. ERDAS Image 8.6 Geometric Correction tool. Example of Ground Control Points (GCP) selected in proximity to the study area. .......................................... 64

Figure 7.3. Geometric correction and co-registration of Landsat TM Path 01 – Row 71 8/2/1987. Before (left) and after (right). The original image was a L1R product. ............................................................................................................................... 65

xv

Figure 7.4. Landsat ETM+ Path 01 – Row 71 6/26/2000 and 7/31/2001. Spatial offset at La Paz Airport (approx. 200 m). Both scenes were originally projected in UTM 19S WGS84. ......................................................................................................... 65

Figure 7.5. Glaciers of the Cordillera Real and detail of the Ancohuma-Illampu Massif. Area is calculated as planimetric surface. Glacier classification by Jordan (1999). ................................................................................................................................67

Figure 7.6. Details of the subset procedure based on Jordan’s glacier inventory (1999). ............................................................................................................................... 68

Figure 7.7. Solar energy path (modified after Schott 1997). ............................................ 69

Figure 7.8. Effects of atmospheric transmission and path radiance on resulting total at-sensor reflectance (modified after Schott 1997). .................................................. 71

Figure 7.9. MODTRAN 3. Example of longwave atmospheric transfer model output from standardized input (http://geosci.uchicago.edu/~archer/cgimodels/radiation.html). ............ 75

Figure 8.1. Maximum in-scene reflectance registered by TM and ETM+ sensors in the five subscenes. ...................................................................................................... 86

Figure 8.2. Percent saturated pixels for each band of the subscenes. .............................. 87

Figure 8.3. Accumulation (Zone III and Zone II) and ablation (Zone I) zones of Zongo Glacier as identified by Klein and Isacks (1999).................................................. 89

Figure 8.4. Sample of training sites used to obtain the spectral signatures for the information classes................................................................................................ 89

Figure 8.5. Spectral signatures obtained from the training sites on Zongo Glacier.......... 90

Figure 8.6. Binary classification of Zongo Glacier using the snow/ice mask after Dozier (1989) applied to four Landsat scenes. ................................................................. 92

Figure 8.7. Minimum distance supervised classification of Zongo Glacier from Landsat TM 1987................................................................................................................ 93

Figure 8.8. Minimum distance supervised classification of Zongo Glacier from Landsat ETM+ 2000........................................................................................................... 93

Figure 8.9. Accumulation and ablation endmembers of Zongo Glacier, and AB transect (figure 8.10). The fraction is represented in grey scale with white as the highest and black as lowest . ............................................................................................. 95

Figure 8.10. Accumulation and ablation endmember fractions along the AB transect (figure 8.9). ........................................................................................................... 96

xvi

Figure 9.1. Total snow/ice cover of the Cordillera Real measured by binary classification. ............................................................................................................................... 98

Figure 9.2. Total snow/ice cover of the Cordillera Real measured by minimum distance classifications...................................................................................................... 101

Figure 9.3. Surface changes in km2 for each scenes pair obtained from the minimum distance classification. ........................................................................................ 107

Figure 9.4. Percent change from ablation class to other classes. .................................... 108

Figure 9.5. Comparison of the total snow/ice cover of the Cordillera Real measured by binary and minimum distance classifications. .................................................... 109

Figure 9.6. Laguna Glaciar: Comparison of the total snow/ice measured by binary and minimum distance classifications. ...................................................................... 111

Figure 9.7. Zongo Glacier: Comparison of the total snow/ice measured by binary and minimum distance classifications. ...................................................................... 112

Figure 9.8. Chacaltaya Glacier: Comparison of the total snow/ice measured by binary and minimum distance classifications. ...................................................................... 113

Figure 10.1. ASTER scene (left) acquired on 06/29/2001 and Landsat ETM+ scene (right) acquired on 07/31/2001 representing the central portion of the Cordillera Real after an extended snow fall event (USGS Global Visualization Viewer, http://glovis.usgs.gov). ............................................................................................. 115

Figure A.1. Snow and ice cover sample of approximately 59 km2................................ 129

Figure A.2. Soils and vegetation sample of approximately 43 km2............................... 130

Figure A.3. Soils and vegetation sample before and after topographic normalization... 132

Figure A.4. Snow and ice cover sample before and after topographic normalization.... 133

Figure A.5. Regression analysis for the soil and low vegetation subset (band TM 3) before normalization. .......................................................................................... 134

Figure A.6. Regression analysis for the soil and low vegetation subset (band TM 3) after normalization. ..................................................................................................... 135

Figure A.7. Regression analysis for the snow and ice cover subset (band TM 3) before normalization. ..................................................................................................... 136

Figure A.8. Regression analysis for the snow and ice cover subset (band TM 3) after normalization. ..................................................................................................... 136

xvii

Figure A.9. Regression analysis for the snow and ice cover subset (band TM 5) before normalization.... 137

1

Chapter 1

Introduction

Fluctuations of mountain glaciers and ice caps are key variables for the

observation of climate-related changes and the detection of enhanced greenhouse effects

(IPCC 2001). The World Glacier Monitoring Service (WGMS) and the Global Land Ice

Measurements from Space (GLIMS) projects are currently coordinating a world-wide

research program to continue monitoring glaciers in areas already represented by more

than a century of observations and to obtain direct and inventory data from a number of

under-represented locations, such as the Andes. These areas represent a critical gap in the

existing research network. The analysis of spatially and temporally distributed

glaciological data will contribute to a better understanding of the driving forces and

processes involved in global climate changes.

Glaciers and ice caps develop under a narrow range of climate conditions (defined

temperature range, humidity and precipitation), showing strong links with atmospheric

circulation changes and climate variables. Monitoring the distribution, intensity, and rate

of change of ice masses is of utmost importance in delineating spatial and temporal

variations of the Earth’s climatic system.

The purpose of this research is to analyze snow and ice cover changes of the

glaciers of the Cordillera Real (Bolivia) during the period 1987-2001. The study is based

2

on satellite remote sensing technology (Landsat TM and ETM+), which is able to provide

a regional synoptic view of the study area and discrete data coverage during the defined

period of time. This research seeks to enhance the regional understanding of trends and

dynamics of the glacier retreat observed on benchmark glaciers such as Zongo and

Chacaltaya. This research will highlight the advantages and disadvantages of remote

sensing as a tool to generate relatively fast and inexpensive information about glacial

advance or retreat. Another potential outcome of this research (but not further

investigated) is the development of water resource management policies and planning for

the study area. In fact, glacial melt-water from the Cordillera Real provides vital fresh

water resources for local agriculture, the hydroelectric industry, and the large human

population present in the area. The rate of retreat and glacial mass loss are of vital

concern for the planning and provision of future alternative water resources. Thus,

detecting and monitoring snow and ice mass variations in the Bolivian glaciers are not

only important for delineating spatial and temporal variations of the Earth’s climatic

system, but also for fresh water resource management necessary for sustainable human

living and power generation (Bindschadler, Dowdeswell, Hall, and Winther 2001).

This research seeks to: (1) provide an overview of the climatic variables that

affect tropical glaciers in the Bolivian Andes; (2) describe the most valuable research

methodologies to study glaciers and ice-caps; (3) investigate the application of remote-

sensing based measurements for the development of glacier inventories and the

estimation of regional mass balance variations; and (4) apply a set of remote-sensing

based techniques to Landsat TM and ETM+ imagery to estimate regional snow and ice

cover variations of the Cordillera Real from 1987 to 2001.

3

Chapter 2

The Study of Glaciers: Methodology Overview

By definition, a glacier (or ice-cap) will form from the snow accumulated and

compacted to ice where the rate of accumulation exceeds the rate of ablation.

Accumulation occurs most commonly as a result of snow, hail, and frost precipitation;

avalanches and secondary processes may increase the accumulation rate locally. Ablation

can occur by melting and/or sublimation. The balance between accumulation and ablation

defines the mass balance of a glacier; mass balance is largely dependent on climate

(Bennett and Glasser 1996). Accumulation dominates the upper region of a glacier where

temperatures are constantly (or mostly) below freezing, while ablation dominates the

glacier terminus where temperatures are generally at, or above, the freezing point.

The spatial imbalance between accumulation and ablation determines the surface

morphology and drives ice gravitational flow (Bennett and Glasser 1996). The net

balance gradient is defined as the increase in net balance (accumulation minus ablation)

with altitude; moving from lower to higher elevation, the accumulation of ice exceeds

loss of ice at a greater rate (figure 2.1). The equilibrium of the surface slope of a glacier is

a function of the net balance gradient (Bennett and Glasser 1996). As a result, net balance

gradient is steepest on glaciers that experience warm damp maritime climates, and lowest

in cold dry continental areas.

4

Mass balance measurements are based on the study of glacier inputs and outputs.

Measurements can be achieved with direct and indirect methodologies, but even with the

most accurate procedure, the results are based on interpolations and estimations.

Braithwaite (2002) explained that glacier mass balance forms an important link between

the atmospheric environment and glacier dynamics and hydrology. Glacier mass balance

changes influence oceanic mass and, thus, are of global impact. Locally, changes in ice

mass affect water resources and hydrology (such as surging floods or irreversible drought

in the case of complete melting).

Haeberli, Frauenfelder, Hoelzle and Maisch (1999) emphasized that the final

purpose of mass balance studies is to gain enough understanding to be able to

parameterize unmeasured glaciers and describe ongoing global changes. The following

mass balance measurement techniques are the most frequently used in research programs.

Accumulation wedge(positive net balance)

Ablation wedge (negative net balance)

Equilibrium line

Ice movement

Figure 2.1. Schematic representation of the profile of a glacier. Net balance is positive in the accumulation zone and negative in the ablation zone. Ice movements transfer mass from the accumulation to the ablation zone constantly reshaping the profile of the glacier (modified after Bennett and Glasser 1996).

5

On-site Measurements

On-site measurements produce the most accurate and significant data. Accurate

measurements of mass variation over time and space are essential for the identification of

those processes that modify glacier dynamics. Through direct measurement, it is possible

to collect reliable information to be used as benchmarks to calibrate mass-balance and

energy-balance models. The disadvantages of on-site measurements are mainly the high

demand of resources (technical and human) and the high cost of long-term operations in

extreme environments (Haeberli et al. 1999).

Mass balance measurements can be achieved by on-site surveys of mass density,

volumetric change (geometric), and flow rates. This methodology relies on the

installation of spatially distributed survey stations over the glacier surface to sample snow

and ice thickness variation and ice flow rates throughout a specific period of time; the

survey stations use graduated borehole stakes and snow pits. Densimetric data are

obtained from ice and snow cover samples. Using on-site surveys, it is possible to obtain

direct measurements of the accumulation and ablation zones (especially with regard to

the surface extent), the equilibrium line (in particular its elevation shifts in time), the

thickness of the ice to bedrock, and the velocity vectors of flow (directions and

intensities).

Data collected from the survey stations are spatially interpolated to model the

glacial surface; the resulting surface provides the base to determine mass losses and gains

and the estimation of mass variation per unit of time and space. Nearby meteorological

stations provide data to correlate mass balance with atmospheric parameters (such as

temperature, pressure, moisture, and wind) and the opportunity to monitor glacier

6

response to changes in climatic patterns. Examples of direct mass balance measurements

are extensive, because this methodology was the first applied to the study of glacier

dynamics (Braithwaite 2002).

Another form of direct measurement is to record data relative to the energy fluxes

that control melting and freezing processes. Sensors measuring incoming and outgoing

long- and short-wave radiation, albedo, and latent/sensible heat fluxes (to mention a few

variables) are installed as part of meteorological stations. Mass variation is estimated

from energy fluxes over time and space, and requires complex glacier-climate models.

The models take into consideration variation of mass balance in response to temperature

changes. An important parameter considered in these models is the climate sensitivity of

mass balance (∆balance/∆T), which is the basis for static and dynamic models. Static

models state that temperature changes are reflected by instantaneous balance adjustment

(such as the models used by IPCC’s Sea Level Panel); dynamic models consider the

delayed responses in thickness and area adjustment induced by temperature variations.

One of the most commonly used models of sensitivity is the degree-day model.

As Braithwaite and Zhang (1999) explained, the +1 degree temperature increase

represents a suitable standard for comparing sensitivities from different glaciers and

models applied to the same glacier. The degree-day model takes into account the spatial

distribution of mass balance over the entire glacier area, accounting for the mass balance

variations in elevation bands.

Some examples of degree-day melting factors (Braithwaite and Zhang 1999) are:

8.0 mm day-1 deg-1 for ice and 4.5 mm day-1 deg-1 for snow (millimeters of snow and ice

melting each day for +1°C increase in temperature). The degree day melting factors are

7

highly dependent upon local climatic and morphologic settings. Once the model is

calibrated, it is possible to associate climate changes with mass balance variations and

vice versa. While requiring fewer human resources than direct mass balance

measurements, energy balance and sensitivity models require costly instrumentation and

data processing. As is the case for the direct mass balance measurements, energy balance

methods require long-term continuous climatic and energy flux measurements over the

glacier. Energy balance methods provide a better description of the physical processes

involving climate in mass balance variations.

Indirect Measurements

Indirect mass balance estimation methods rely on remote sensing technology and

ancillary databases. Indirect measurements and estimates are highly cost-effective, but

due to the nature of the data are more prone to measurement error and biases, and are thus

best applied in studies of remote areas, for large area inventories, and for the

development of inventories adjusted and calibrated on benchmarks from direct mass

balance measurements (proxies).

One possible compromise to costly on-site measurements is mass balance

estimation from topographic surveys, which represents a way to estimate thickness

variations and infer flow vectors from superficial flow. The advantage of this method is

that it is possible to rely on ancillary data such as topographic maps generated from

airborne stereo-pairs, digital elevation models (DEM) produced from satellite/airborne

sensors, and GPS/geodetic on-site control points. The method compares topographic

variations over a period of time, which leads to the estimation of volumetric changes and

8

mass balance variations. The method relies on the accuracy and precision of topographic

elevation measurements, which cannot always be estimated for historical topographic

data.

The topographic measurement methodology is of value for the estimation of

regional mass balance variations where direct methods are not cost effective, due to the

extent of the study area, the lack of research resources, or the lack of historical on-site

mass measurements. On the other hand, the area of study must be well covered by

topographic surveys (such as from military agencies, or private and public

airborne/satellite sensors).

Modern technologies (especially SAR-SRTM, LIDAR, and more recently,

ASTER) are improving the accuracy of elevation measurements, thus increasing the

reliability of this methodology. Topographic methods have been successfully applied on

single glaciers (Rivera and Casassa 1999; Lange, Araos and Rivera 2003; Rivera,

Benham, Casassa, Bamber and Dowdeswell 2003) and for the estimation of extensive

regional mass balance (Rignot, Rivera and Casassa 2003; Schnirch, Schneider, Casassa

and Kilian 2003; Rivera et al. 2003; Bamber 2003; Keller, Casassa, Rivera, Forsberg and

Gundestrup 2003).

When topographic data are not available or have limited availability, mass

balance and mass variation can be estimated from surface area changes (indicated

primarily by the retreat or advance of the terminus). Airborne photography and satellite

imagery can be used in conjunction with Geographic Information System (GIS) data to

map the spatial variation of glaciers and ice caps over time. While providing indications

of change, surface area variations do not always represent mass variation in a linear

9

fashion; while easy to detect, surface changes present considerable difficulties of

interpretation (Haeberli 2003).

The accumulation area ratio (AAR = Accumulation area / Total area; an AAR=0.7

corresponds to mass equilibrium at the end of the melting season) and the equilibrium

line altitude (ELA) are two identified parameters that can provide researchers a better

understanding of the interrelation of area and balance variation. The two parameters are

directly connected: the boundary between the accumulation and the ablation zone is

defined by the equilibrium line. The ELA is a function of temperature, and thus may be

considered covariant to mass balance changes (also a function of temperature). Shifts of

the ELA are reflected in changes of surface extent of the accumulation and the ablation

zones.

Bennett and Glasser (1996) discussed how ablation and accumulation zone

geometry is the result of dynamic processes that work to maintain a constant slope; these

processes are functions of climatic conditions. The AAR is thus a property of a glacier in

specific climatic conditions and slope equilibrium: any variation of the AAR through

time reflects rearrangements in the net balance gradient, and thus mass balance.

Numerical estimates of mass balance from the AAR are based on proxy data and

regression from neighboring glaciers. The AAR and the ELA are relatively easy to

determine through remote sensing and allow regional estimates of mass balance variation

once calibrated with measurements on benchmark glaciers.

The error of estimate determined through remote sensing is a complex function of

proximity, geomorphology, and local climatic conditions between the study area and the

benchmark glaciers. A considerable number of assumptions must be taken into

10

consideration, but the net advantage of the method is that it provides an acceptable

estimate of mass balance variations relatively quickly, inexpensively, and on a regional

scale. Remote sensing based estimates reduce human resource requirements, and costs.

These factors compensate for the loss of accuracy. Also, it is the only methodology that

can reasonably be applied for preliminary estimates on remote or unexplored areas such

as alpine glaciers.

Remotely sensed imagery can provide other important glaciological information.

Skvarca, Stuefer and Rott (1999), for instance, successfully determined flow velocity

vectors from the observation of change of shape of surface debris from rock falls. These

measurements are of utmost importance for verifying the amount of mass displaced by

flow. The rate of change of flow velocity is a function of dynamic processes that maintain

slope equilibrium within the glacier, and thus is associated with mass balance variation

(in particular, mass displacement) and climatic variability.

A more sophisticated indirect estimate of mass balance relies on the integration of

topographic measurements (such as thickness variation from DEM and topographic

maps, as previously discussed) and surface change detection. Haeberli and Hoelzle

(1994) presented a model based on a series of easily calculated glacier parameters to

estimate mass balance. The estimate is based on four simple geometric parameters that

are commonly found in detailed inventory data: length of the glacier, maximum and

minimum altitude, and total surface area of the investigated glacier. From these

parameters, mean altitude, vertical extent, and average surface slope are derived.

Analysis involves estimation of total volume, glacier thickness, and the mean flow

velocity, which are obtained from empirical models. Glacier length changes for given

11

changes in mass balance are also related to a characteristic dynamic response time. The

time interval between first reaction and full response is called relaxation time (Haeberli

and Hoelzle 1994).

Haeberli and Hoelzle (1994) argued that the simplicity of the model is reflected in

a number of uncertainties. While the methodology presents inherited error of calculation

of derived flow velocities, response time, and other factors, the results are acceptable to

determine mass balance for medium to large glaciers. The authors suggested that

empirical and regional approaches must be investigated to tune the parameters of the

model for the different sites and conditions of mountain ranges around the world.

Haeberli and Hoelzle (1994, 11) stated that the most important potential benefit of

the methodology (and in general of all indirect methodologies) “is given with the

possibility of quantitatively inferring average decadal mass balance for unmeasured

glaciers by analyzing cumulative length changes from field evidence;” the field evidence

includes moraines, satellite imagery, aerial photography, and long-term observation. In

addition, Haeberli and Hoelzle (1994, 11) stated that regional inventories open new

opportunities “to provide the basis for evaluating scenarios of global warming.”

Coring

Borehole surveys and ice coring represent other important methodologies for

studying glaciers and ice caps. Coring focuses on the analysis and reconstruction of

climatological and paleo-ecological information preserved in glaciers and ice caps.

Thompson and his research team at Ohio State University (Thompson 2000; Thompson

and Davis 1998) are probably the most accomplished scientific group using ice coring.

12

Since the beginning of the 1980s, Thompson has undertaken coring survey expeditions

on a considerable number of glaciers and ice caps worldwide (figure 2.2). The purpose of

his research is to develop a global climate change chronological map suitable for

analyzing the changes in time and the pattern of changes in space of the Earth’s climate

(Thompson 2003).

Climatological and atmospheric information are retrieved from ice cores,

primarily through the analysis of particulate matter and isotopes trapped in the snow

precipitation and subsequently diagenized in ice. Best results are obtained from relatively

stable glaciers, where melting processes do not obliterate the climatic records.

Ice coring is suitable for the reconstruction of time series of temperature,

precipitation, particulate composition, and atmospheric composition; time series might

help determine if the global climatic change measured over the past 200 years is a

Figure 2.2. Ice sample extracted from the coring tool by scientist of Thompson’s research group studying the retreating glaciers of the Kilimanjaro (Tanzania) (NOAA 2002).

13

consequence of natural variability or the result of anthropic activity. Glaciological time

series derived from ice cores present a unique opportunity to understand glacier processes

as a function of global and regional climatic variability. Time series make it possible to

study glacier response to climate change and the relation between these changes at a

global scale. El Niño/Southern Oscillation (ENSO) events, for instance, present a

characteristic signature that has been identified world wide in ice cores. Time series have

shown that the magnitude and frequency of ENSO events have a strong influence on

glacier dynamics at a global scale.

Global climatic maps based on time series from ice cores eventually might help to

identify those mechanisms behind events such as ENSO. In addition, the development of

models to predict future variability as a function of natural patterns and anthropic

perturbation on climate might become possible.

14

Chapter 3

Study Area

Exploration and Historical Data

The glaciers of Bolivia were first described by d'Orbigny (1835-1847); no

previous record of the physiography of the glaciers is available from the native

populations, mostly due to the lack of a native writing system (Jordan 1999). During the

early 20th century, scientific expeditions aimed at “eliminating blank places on maps of

the Earth” (Jordan 1999) produced early maps of the Bolivia’s glaciated mountains.

These early research projects were directed toward the identification and mapping of

snowline and glacier terminus locations, the understanding of glacier morphology, and

the distribution of glaciers. Jordan (1999) listed some of the important researchers of the

time including Conway in 1900, Hauthal in 1911, and Herzog between 1913 and 1915. In

1927-28, a successful climbing expedition led by Carl Troll of the German-Austrian

Alpine Association added important cartographic details to the available maps of the area

using terrestrial photogrammetric methods. Troll specialized in photographs and

triangulation measurements; he was able to produce a number of general and detailed

topographic maps of the northern Cordillera Real (Illampu area) at a scale of 1:50,000

(Troll and Finsterwalder 1935). Later, geological studies provided general maps of the

distribution of glaciers.

15

The most recent topographic map of the Illampu area was developed from aerial

photos taken in 1963 and 1975 (Finsterwalder 1987). The accuracy of the topographic

maps from Troll and Finsterwalder (1935) and Finsterwalder (1987) allowed an initial

quantification of mass balance. At the beginning of the 1990s, Jordan and Finsterwalder

produced another updated cartographic map of the southern Cordillera Real (Illimani

area) at a scale of 1:50,000 (Finsterwalder 1990; Jordan and Finsterwalder 1992).

Only part of the La Paz region and the Cordillera Occidental has been covered by

published maps as of early 2004 (Jordan 1999). In the 1940s, the U.S. Army Map

Service, and later the Instituto Geográfico Militar (IGM) in La Paz, started official

mapping of Bolivia based on vertical aerial photographs; the maps have scales of

1:50,000 and 1:250,000. These maps are unsuitable for glacier studies, because no

distinction is made between snow patches and glaciers (Jordan 1999). However, the

original vertical aerial photographs are suitable for analyzing distribution of glaciers as

they existed at the time of the survey.

In 1967, Mercer conducted surveys of the glaciers of Bolivia using mass-balance

data and energy records, including measurements of ice movement, precipitation,

temperature, evaporation, ablation, and glacier runoff. Mercer charted his observations

from direct measurements and terrestrial photogrammetric photographs. In 1980, Mercer

participated in the compilation of a glacier inventory for Bolivia (Jordan, Brockman

Fernandez, Alvarez and Jacobsen 1980); the inventory was based on aero-triangulations

of Bolivia's glaciated regions carried out from the northern Cordillera Real across the

Cordillera Apolobamba to the Peruvian border (Jordan 1991; Herrmann 1993). The data

are now part of the "World Glacier Inventory" of the United Nations Environment

16

Program/United Nations Educational, Scientific, and Cultural Organization/International

Commission on Snow and Ice (UNEP/UNESCO/ICSI), now the World Glacier

Monitoring Service in Zürich, Switzerland (Jordan 1999).

The first valuable regional image of Bolivian glaciers was taken by the Gemini 9

space flight in June, 1966. The 1972 launch of Landsat 1 carrying the Multi Spectral

Scanner (MSS) marked the beginning of a new era for remote sensing studies. The

glaciers of Bolivia have been regularly observed from space since that time.

Area of Interest

Glaciers in Bolivia are found in two main ranges of the Andes, the Cordillera

Occidental (latitude 18°03' S to 18°25' S, longitude 68°53' W to 69°09' W) along the

western border with Chile, and the southern extension of the Cordillera Oriental (figure

3.1) extended from latitude 14°37' S to 17°04' S and longitude 67°13' W to 69°14' W.

Bolivia lies completely within the tropics, so glaciers are confined to the highest peaks of

the Andes Mountains (Francou 1993) as ice caps, valley glaciers, and mountain glaciers.

Most of the glaciers in Bolivia are found on the Cordillera Oriental, with 1,826 glaciers

covering 591.6 km2 (Jordan 1999); this region includes the Cordillera Apolobamba,

Cordillera Real, Cordillera Tres Cruces, Cordillera de Muñecaus, and the Nevado Santa

Vera Cruz (table 3.1). Annual precipitation within Bolivia is variable, but generally

decreases toward the southwest, where it reaches its seasonal minimum (Jordan 1999);

south of latitude 18°30' S even the highest peaks (6,000 m and above) cannot sustain

glaciers due to the lack of precipitation. The southern limit of glaciers roughly matches

17

the northern line of the large salars (salt pans), such as Salar de Coipasa and Salar de

Uyuni in the central plateau (Altiplano) of the Andes (Jordan 1999).

The Altiplano (extending about 200 km wide between the Cordillera Occidental

and the Cordillera Oriental) does not have sufficient elevation or precipitation to sustain

glaciers. This research focuses on the glaciers of the Cordillera Real (latitude 15°45' S to

16°40' S, longitude 67°40' W to 68°34' W), a section of the Cordillera Oriental (figure 3.2

and figure 3.3). The total glaciated area surveyed on November 1, 1984 (Jordan 1991)

amounted to 323.603 km2 (964 glaciers), corresponding to 54% of the glaciated surface

area of the Cordillera Oriental (53% of the number of glaciers). At the time of the survey,

the glaciers extended from a maximum elevation of 6,436 m to a lowest terminus

elevation of 4,420 m. The entire area is included in a single Landsat scene (MSS, TM and

ETM+ Path 01-Row 71). The study of the glaciers of the Cordillera Real is particularly

important due to the presence of a large population concentration (especially in the La

Paz district) that relies on the melt-waters of the glacier as a fresh water resource. It is

important to understand local glacier variations in relation to global change, but the

analysis of the historical evolution, trends, and pattern of the glacier mass balance of the

Cordillera Real also might help in the development of planning policies and mitigation

projects to optimize administration of water resources in the area.

Ele

vatio

n (m

)

Pacific Ocean Amazon Basin

Cordillera Occidental

Cordillera OrientalAltiplano

Ele

vatio

n (m

)

Pacific Ocean Amazon Basin

Cordillera Occidental

Cordillera OrientalAltiplano

Figure 3.1. Spatial profile of the Andes in proximity of the study area (elevation data from EDC 2003). (Vertical scale is exaggerated).

18

Area

(sq.

km

)Pe

rcen

tN

umbe

r of

glac

iers

Perc

ent

CO

RD

ILLE

RA

OR

IEN

TAL

14°3

7'-1

7°04

'67

°13'

-69°

14'

591.

600

100.

0018

26.0

0010

0.00

6436

.043

11.0

CO

RD

ILLE

RA

AP

OLO

BA

MBA

14°3

7'-1

5°04

'68

°58'

-69°

14'

219.

804

37.2

065

2.00

036

.00

6059

.043

11.0

Cha

upi O

rko

14°4

0'69

°10'

129.

357

21.9

034

6.00

019

.00

6059

.043

65.0

Col

olo

14°5

0'69

°06'

43.0

727.

3013

5.00

07.

5057

74.0

4311

.0U

lla K

haya

15°0

0'69

°03'

47.3

758.

0017

1.00

09.

5056

69.0

4390

.0C

OR

DIL

LER

A D

E M

EC

AS

15°2

0'-1

5°38

'68

°33'

-68°

55'

0.68

40.

1016

.000

1.00

5237

.048

28.0

Mor

ocol

lu15

°20'

68°5

5'0.

148

0.03

8.00

00.

5051

56.0

4828

.0C

uchu

15°3

8'68

°33'

0.53

60.

108.

000

0.50

5237

.048

86.0

CO

RD

ILLE

RA

REA

L15

°45'

-16°

40'

67°4

0'-6

8°34

'32

3.60

354

.70

964.

000

53.0

064

36.0

4420

.0

NO

RTH

ER

N C

OR

DIL

LER

A R

EA

L15

°45'

-16°

20'

68°0

1'-6

8°34

'26

2.76

644

.40

784.

000

43.0

064

36.0

4420

.0Ill

ampu

-Anc

ohum

a15

°50'

68°3

0'10

3.09

917

.40

147.

000

8.00

6436

.044

38.0

Cal

zada

-Chi

aroc

o- C

hach

acom

ani

16°0

0'68

°20'

94.0

7215

.90

251.

000

14.0

061

27.0

4676

.0N

igru

ni-C

ondo

riri

16°0

8'68

°15'

40.8

686.

9024

1.00

013

.00

5752

.044

20.0

Saltu

ni-H

uayn

a Po

tosí

16°1

5'68

°08'

14.5

042.

5050

.000

3.00

6088

.048

04.0

Zong

o-C

umbr

e- C

haca

ltaya

16°1

8'68

°05'

10.2

231.

7095

.000

5.00

5519

.045

78.0

SO

UTH

ER

N C

OR

DIL

LER

A R

EA

L16

°20'

-16°

40'

67°4

0'-6

7°58

'60

.837

10.3

018

0.00

010

.00

6414

.044

99.0

Ham

patu

ri-Ta

ques

i16

°26'

67°5

2'11

.685

2.00

70.0

004.

0055

48.0

4723

.0M

urur

ata

16°3

0'67

°47'

17.2

072.

9075

.000

4.00

5836

.045

92.0

Illim

ani

16°3

8'67

°44'

31.9

455.

4035

.000

2.00

6414

.044

99.0

CO

RD

ILLE

RA

TR

ES C

RU

CE

S (Q

UIM

SA C

RU

Z)16

°47'

-16°

09'

67°2

2'-6

7°32

'45

.276

7.70

177.

000

9.50

5760

.047

08.0

Cho

quet

anga

16°5

4'67

°22'

6.99

21.

2021

.000

1.00

5541

.048

12.0

Hig

h re

gion

of T

res

Cru

ces

16°5

6'67

°24'

38.2

846.

5015

6.00

08.

5057

60.0

4708

.0N

EV

AD

O S

AN

TA V

ER

A C

RU

Z17

°03'

-17°

04'

67°1

3'-6

7°14

'2.

233

0.40

17.0

001.

0055

60.0

4853

.0

CO

RD

ILLE

RA

OC

CID

ENTA

L18

°03'

-18°

25'

68°5

3'-6

9°09

'10

.000

100.

00--

--65

42.0

5100

.0

Nev

ado

Saja

ma

18°0

6'68

°53'

4.00

040

.00

----

6542

.051

00.0

Nev

ados

Pay

acha

ta18

°09'

69°0

9'4.

000

40.0

0--

--62

22.0

5500

.0C

erro

s Q

uim

sach

ata

18°2

3'69

°03'

2.00

020

.00

----

6032

.055

00.0

Low

est

term

inus

(m

eter

s)

Area of Interest

CO

RD

ILLE

RA

OC

CID

ENTA

L

CO

RD

ILLE

RA

OR

IEN

TAL

Gla

cier

Mou

ntai

n gr

oup

Latit

ude

(sou

th)

Long

itude

(w

est)

Hig

hest

el

evat

ion

(met

ers)

Tabl

e 3.

1. G

laci

ers o

f the

Bol

ivia

n A

ndes

as N

ovem

ber,

1st 1

984

(mod

ified

afte

r Jor

dan

1991

).

19

Figure 3.2. Bolivia and relative location of the area of study area of study (datasets from EDC 2003).

La Paz

Sucre

Santa Cruz de la Sierra

BoliviaLake Titicaca

70 W

65 W

10 S

15 S

Tropic of Capricorn

15 S

20 S

Area of Study

La Paz

Sucre

Santa Cruz de la Sierra

BoliviaLake Titicaca

70 W

65 W

10 S

15 S

Tropic of Capricorn

15 S

20 S

Area of Study

20

Cordillera Real

La Paz

LakeTiticaca

Landsat Path 01 – Row 71

Cordillera Real

La Paz

LakeTiticaca

Landsat Path 01 – Row 71

Figure 3.3. Area of study. Landsat image of the Cordillera Real.

21

Chapter 4

Climatic Conditions Affecting Tropical Glaciers of Bolivia

Tropical climates are characterized by a homogenous atmosphere, infrequent

frontal activity, and one or two distinct and pronounced precipitation seasons, thus

presenting a characteristic impact on tropical glaciers (Kaser 1999). By definition (Kaser

1999), a tropical glacier must be located within the tropics, have a daily temperature

variation that exceeds the annual temperature variation (thermal delimitation), and be

within the Inter-tropical Convergence Zone (ITCZ) oscillation area (hygric delimitation).

Another important boundary condition characteristic of tropical climates is the thermal

homogeneity which is reflected in a fairly stable elevation of the 0º C isotherm.

Kaser (1999) argued that, in the absence of seasonal thermal variability, the

vertical mass balance gradient on the tongues of tropical glaciers is weaker than that

observed on mid-latitude glaciers. As a result, accumulation area ratio (AAR) is higher

and the terminus is more sensitive to equilibrium line altitude (ELA) shifts; changes in

temperature affect ablation (due to sensible heat flux) and accumulation zone (due to

ELA shift).

Historically, tropical glaciers had the most recent maximum extent during the

Little Ice Age (LIA; estimated from 1350 to 1900), and have receded since the middle of

the 19th century. Starting in the 1930s, it was possible to observe a drastic glacier retreat

22

that slowed, but did not stop, during the 1950s to 1970s period. After the 1970s, the rate

of retreat has consistently increased; only a few locations have been marked by terminus

advance (Kaser 1999; Hooker and Fitzharris 1999; Haeberli and Hoelzle 1999; Rivera

and Casassa 1999; Thompson 1997; Thompson and Davis 1998; Thompson 2000). The

general trend of tropical glacier fluctuation agrees with the pattern of mid-latitude

glaciers, leading to the assumption that the processes behind observed change are of

global scale (Kaser 1999; Thompson 2000). This assumption becomes evident when

considering that air temperature and water vapor pressure, the two main factors affecting

tropical glacier fluctuations, are the result of the energetic state of the planetary climatic

system.

Kaser (1999) explained that in tropical glaciers temperature changes affect all

respective mass balance components uniformly, while air moisture changes have

different effects in dry and humid areas and seasons. As an example, an increase of

humidity during the dry season may reduce sublimation, but increase long wave radiation

with a general increase in negative mass balance; in humid areas (or during a wet season)

a reduction of humidity may cause mass loss due to increased insolation and decreased

precipitation. Due to the thermal stability of the tropical climate, it may be assumed that

fluctuations of tropical glaciers are associated with vapor pressure and air humidity

oscillation, thus making tropical glaciers a suitable benchmark for global change studies

(Kaser 1999).

In agreement with Kaser’s observations, Francou, Vuille, Wagnon, Mendoza and

Sicart (2003) highlighted the influence of precipitation variability and air humidity on the

mass balance of tropical glaciers. From a decade-long study of direct mass balance

23

measurements on Zongo and Chacaltaya glaciers in Bolivia, Francou et al. (2003)

observed that there was a close relationship between mass balance and atmospheric

condition during the austral summer (October to April), in particular from December to

February. During this period (the wet season), humidity and temperature are high and

melting prevails over sublimation. Net all-wave radiation is the dominant factor which

governs ablation. Albedo is a function of solid precipitation: a lack of precipitation (low

cloudiness) reduces the snow cover, exposing bare ice (characterized by lower albedo

properties), thus increasing energy available for melting processes. This process is

particularly important in the Bolivian tropical glaciers where the ablation and

accumulation season coincides with the austral summer. This is quite distinct from mid-

latitude glaciers, where there is a clear distinction between accumulation and ablation

seasons. The tropical austral summer presents the highest temperatures (thus energy

available for melting) and highest air humidity, resulting in the most intense snow

precipitation of the year.

Francou et al. (2003) argued that temperature is a key variable for long term

variations of tropical glaciers, because temperature governs all climatic fluxes on a global

scale. However, air humidity and precipitation patterns better predict mass balance

variation over short periods (annual/inter-annual) for the study area. Francou and his

colleagues observed that a lack of precipitation (reduced snow cover), reduced cloudiness

(low albedo), and increased net incoming solar radiation during the December-January-

February (DJF) months resulted in negative mass balance. During the dry winter months,

sublimation is the dominant ablation process, which is less effective than melting. If

abundant snow precipitation is available during the DJF months, the snow cover will

24

maintain high albedo during dry winters, reducing the energy available for melting.

Tropical dry winters represent a period of mass conservation. Francou et al. (2003)

showed that, in fact, the relationship between climatic/meteorological conditions and

mass balance variations is highly influenced by the summer season variability and, in

particular they found that mass balance and climatic/meteorological condition present a

more significant correlation when the dry season observations are removed from the

annual series.

Variability during the summer season has been associated with Sea Surface

Temperature (SST) oscillation (in this case of the western pacific coast of South

America) in conjunction with El Niño Southern Oscillation (ENSO) events. In particular,

glacial mass balance in the Bolivian Andes is highly associated with the increased

frequency of ENSO during recent decades (Thompson and Davis 1998; Kaser 1999;

Thompson 2000; Francou et al. 2003).

During La Niña years, the wind pattern is dominated by moisture rich easterlies

(moisture is collected over the Amazon basin), which favor snow precipitation and

accumulation during the austral summer months (figure 4.1). During El Niño events, low

moisture westerlies are predominant; the rain shadow effect of the Andean coastal ranges

reduces precipitation. Reduced precipitation and increased overall temperature negatively

affect the annual mass balance of Bolivian glaciers (Francou et al. 2003). The higher

frequency and changed spatial-temporal distribution of El Niño events since the 1970s,

and a general tropospheric warming, might explain the recent rapid glacial retreat in the

Bolivian Andes. Thompson and Davis (1998) showed similar results based on a

comparative analysis of ice cores (containing 25,000 years of paleo-climatologic data)

25

from Bolivian glaciers. Time series from ice cores have shown strong correlations

between drought in the Bolivian Andes and El Niño events; the effect is visible in the

reduced seasonal ice thickness and an overall increase in tropospheric temperature from

measurements of oxygen isotopes (18O).

El Niño drought episodes, corresponding to reduced or negative glacier mass

balance, are characterized by a high concentration of dust particulates, high temperatures,

and low precipitation. Ice cores also confirm the general historical trend of deglaciation

in the area, only temporarily delayed by the relatively colder episode of the LIA. Analysis

of the composition of particulate matter allowed Thompson and Davis (1998) to state that

during La Niña events the main source of moisture is localized over the Amazon basin, in

accordance with recent climatologic models.

Figure 4.1. During La Niña years, the wind pattern is dominated by easterlies rich in moisture collected over the Amazon basin. Easterlies favor snow precipitation and accumulation during the austral summer months. During El Niño events, low moisture westerlies are predominant; the rain shadow effect of the Andean coastal ranges reduces precipitation (Francou, Vuille, Wagnon, Mendoza and Sicart 2003).

Ele

vatio

n (m

)

Pacific Ocean Amazon Basin

Cordillera Occidental

Cordillera OrientalAltiplano

El Niño La Niña

Ele

vatio

n (m

)

Pacific Ocean Amazon Basin

Cordillera Occidental

Cordillera OrientalAltiplano

El Niño La Niña

26

Chapter 5

Remotely Sensed Change Detection of Snow and Ice

Remotely sensed imagery offers an unequaled possibility to study landforms from

a synoptic point of view, thus making remote sensing a suitable technology for placing

results from local research within a regional context. Four remote sensing technologies

can be applied to glaciological studies: visible (analog and digital), multi-spectral, radar,

and laser. All four sensors can be carried by two different vectors: airborne (low

elevation flying missions) and satellite (high elevation).

Airborne platforms (equipped with any of the four sensors) have the advantage of

providing very high spatial resolution, but the disadvantage of being relatively costly and

not suitable for extremely remote and unexplored areas. On the other hand, satellite

platforms offer a generally continuous coverage of the Earth’s surface. Satellite imagery

is a relatively low cost product, and the low resolution typical of older systems is being

replaced by modern high spatial resolution sensors. Within the four technologies, each

sensor is best suited for different research questions, but the digital format of modern

products allows the integration of heterogeneous sensors in order to increase the amount

of available information.

Imagery derived from visible wavelengths is mainly used for feature detection; in

glaciology, the imagery is used as base maps for change detection of landforms (such as

moraines), terminus position, areal distribution, and surface flow. Airborne photography

27

was the first remote sensing technology utilized in glaciology for local and regional

studies, and thus, there are numerous historical image databases that depict glaciers since

the invention of photography (Braithwaite 2002). These databases are of utmost

importance and, in some cases, allow researchers to extend change detection time series

back past the beginning of the 20th century.

Multispectral satellite imagery is the most direct descendent from airborne

photography. Originally used for military purposes, multispectral imagery is based on

optical sensors (mechanical and/or digital) that simultaneously record the visible and

infrared domains. Each sensor is sensitive to a particular electromagnetic frequency (or

range of frequencies) and thus provides specific information about the reflector. In fact,

depending on the physical properties of an object, electromagnetic energy is reflected,

absorbed, and transmitted in different ways for each wavelength of the spectrum.

Spaceborne multispectral sensors record the electromagnetic response (radiance) of an

object on the surface of the Earth irradiated by the incoming solar energy, assign to this

measurement a scaled digital value, and store the value in the corresponding pixel of the

image. In glaciology, multispectral imagery is essential for the detection of snow and ice

cover (at the local, regional, and global scales) and to distinguish between various types

of snow and ice facies; thus, multispectral imagery provides a suitable tool for change

detection studies.

Radar imagery is mainly applied as a geomorphologic tool; it provides selective

feature detection, especially based on relief and shape contrast. Most recently, multiband

radar sensors have opened new opportunities to approach classification and change

detection through non-optical methodologies. The most important application of radar

28

technology in glaciology is known as interferometry. This new technology is based on the

assumption that phase changes detected from multi-temporal radar scans of a specific

scene are a function of topographic or morphologic variations on the surface. As a result,

interferometry can be applied to detect velocity flow of the surface of glaciers (JPL 2003;

Rignot et al. 2003).

In recent years, especially in response to widespread use of GIS technology,

satellite imagery has been expanded and applied to develop a new class of remotely

sensed information: digital elevation models (DEMs) and terrain models (DTMs).

Historically, topographic maps and elevation models were created from stereo

photography sampled by airborne sensors. Most recently, new sampling techniques and

processing capabilities have allowed the development of DEMs and DTMs from virtually

any of the four sensor types (visible optical, multispectral, radar, and laser). Starting with

radar sensors, followed by laser altimetry (LIDAR), and currently by

optical/multispectral sensors such as the Advanced Spaceborne Thermal Emission and

Reflection (ASTER), it is possible to produce a DEM of virtually any place on the

Earth’s surface. The full extent of the new DEM potential will be known within a few

years, after the generation of time-based digital elevation series of the surface of the

Earth.

In the meantime, DEMs and DTMs allow geographers and glaciologists to detect

topographic changes by comparing digital information to historical topographic maps,

calculating volumetric or mass changes over time; this is particularly important (as

previously discussed) for the detection of mass balance variation over time. Differential

topographic measurements are the key to obtaining the best possible estimate of mass

29

balance from remote sensing imagery. While topographic changes and topographic

parameters (see Haeberli and Hoelzle 1994) are essential for the study of glacier

dynamics, it is also very important to be able to detect accumulation and ablation zone

extents and their variation over time (to calculate AAR), and to determine the shifts and

migration of ELA to quantify change detection of glaciers. This is the final purpose of

spectral classification of remotely sensed multispectral imagery.

Like every other substance on the Earth’s surface, snow and ice present a

characteristic response to electro-magnetic radiation. Unfortunately, apart from specific

global snow and ice satellite systems (with very low spatial resolution), the most common

multispectral sensors utilized for land cover change detection (Multi-Spectral Scanner

[MSS], Thematic Mapper [TM] and Enhanced Thematic Mapper [ETM+] on Landsat) do

not offer specific tools or observation sensors targeted for snow and ice. The best

possible classification results come from the integration of heterogeneous information

(different sensors) and ancillary data. Nevertheless, a set of theories and techniques have

been developed to classify the surface of glaciers using the available multispectral and

radar sensors. The main classification and change detection approaches found to be

effective for classifying snow and ice facies of glaciers include supervised and

unsupervised classifications based on spectral reflectance, band ratioing, and spectral un-

mixing.

Supervised and Unsupervised Classifications

Multispectral classification implements spectral pattern recognition algorithms to

identify groups of pixels that have common spectral characteristics (spectral signatures),

30

and then arranges the pixels in a finite number of information classes. Pattern recognition

algorithms require a set of training sites, or spectral classes; the training sites may be

selected manually (supervised methods), or automatically (unsupervised methods)

(ERDAS 1997).

Supervised classification is controlled by the analyst, who provides a set of

spectral classes from in-scene training sites or from spectral libraries. Unsupervised

classification is automated, but the user may control a set of general options, such as the

number of final information classes. This research aims to develop a classification based

on the recognition of physical properties of snow and ice facies of glaciers; for this

purpose, one of the most suitable methodologies is a supervised classification based on

spectral reflectance libraries. The spectral libraries for this research were derived from

the integration of previous research in the area (Klein and Isacks 1999), in-scene training

sites, and theoretical spectral signatures (Rosenthal and Dozier 1996).

Band Ratioing

Band rationing is a technique based on the development of indices from relative

reflectance ratios between different bands. Band ratioing removes (or reduces) spectral

differences resulting from variations of illumination and is thus useful for removing the

shadowing effect typical of high-relief glacial environments in alpine areas (Schott 1997).

Landsat TM and ETM+ imagery are used for land cover and land use change

detection, and have been widely implemented in glaciology to map temporal changes of

terminus position and glacial albedo factors. Within the seven TM (and ETM+)

wavelength bands, the visible bands (1, 2 and 3) tend to saturate on snow and ice fields.

31

These bands are used to detect visible features to isolate ice and snow from non-glaciated

areas using digital masking (Klein and Isacks 1999). Band 4 (near infrared) can be used

to distinguish between ablation and accumulation zones; however, to avoid illumination

variations, the best results can be obtained from TM4/TM5 band ratioing (Hall, Ormsby,

Bindschadler and Siddalingaiah 1987; Paul, Kaab, Maisch, Kellenberger, and Haeberli

2002). Gao and Liu (2001) reported that TM3/TM4 ratioing can be used to enhance blue

ice from snow, and TM3/TM5 ratioing is suitable to enhance snow grain size variations.

Normalized difference (ND) ratios represent a more sophisticated ratioing

technique and allow a better filtering of atmospheric and illumination effects (Schott

1997). Five main ND indices are currently used to classify glaciated areas.

1. 5252

TMTMTMTMNDSI

+−

= (equation 5.1)

known as Normalized Difference Snow Index (NDSI) (Dozier 1989; Hall, Riggs,

and Salomonson 1995; Hall, Foster, Verbyla, and Klein 1998);

2. 5353

TMTMTMTMNDSII

+−

= (equation 5.2)

known as Normalized Difference Snow/Ice Index (NDSII) (Dozier 1989; Xiao,

Shen and Qin 2001);

3. 5454

TMTMTMTM

+− (equation 5.3)

which is a grain size index for finer grains (higher values represent larger grains)

(Dozier 1989);

32

4. 4242

TMTMTMTM

+− (equation 5.4)

another grain size index but for all sizes (higher values represent large grains)

(Dozier 1989);

5. 2121

TMTMTMTM

+− (equation 5.5)

is used as contamination index (higher values represent cleaner snow) (Dozier

1989).

All these indices have the advantage of requiring little digital processing.

Spectral Un-mixing

One of the main problems in producing an accurate estimate of snow and ice

cover in alpine terrain is that snow and ice cover varies at spatial resolutions finer than

the instantaneous field of view (IFOV) of most satellite platforms. At the pixel level

(IFOV), snow and ice might mix with other landcover types, such as forests, exposed

rock, and bare soil, resulting in a mixed and complex spectral response; atmospheric

properties and anisotropic reflectance also might vary at sub-pixel level. The resulting

spectral response is a complex function of the reflective properties of landcover types that

can be solved through spectral un-mixing techniques. Spectral mixture analysis assumes

that the energy measured by satellite sensors is a linear combination of the radiances

reflected by the landcovers in the IFOV (Painter, Dozier, Roberts, Robert, and Green.

2003). Spectral un-mixing analysis uses least square modeling to describe the linear

combination of pure components known as endmembers (Mather 1999).

33

Rosenthal and Dozier (1996) found that satellite derived spectral un-mixing

classification estimated snow and ice fractions with accuracy comparable to that of aerial

photography, but in less time, and suggested that spectral un-mixing is suitable to study

rugged terrain environments where there is extreme variation of landcover types over a

short horizontal distance; spectral un-mixing also could normalize terrain effects and

shadowing when spectral signatures vary in amplitude, but not in shape.

Nolin, Dozier, and Mertes (1993) pointed out that spectral un-mixing models

assume a linear combination of the spectral endmembers and argued that this assumption,

while making for easier computation and interpretation, induces error of estimate due to

the non-linear nature of the spectral mixing phenomenon. The error of estimate is mostly

related to atmospheric effects and anisotropic reflectance, and must be taken into account

when evaluating landcover percentages from spectral un-mixing analysis.

Klein and Isacks (1998; 1999) proposed a method to detect AAR and ELA

change based on spectral mixture classification. This technique was found to be

extremely effective for glacier change detection, due to the transitional nature of snow

and ice properties; in fact, snow and ice present an almost unlimited gradient of water

content, grain size, density, particulate concentration and texture structure. Spectral

mixture analysis has been successful in the estimation of fractions of snow cover within

pixels. In estimating the transient snowline position (ELA), spectral mixture analysis was

superior to the use of single bands and band ratios, and in most cases, endmembers from

benchmark glaciers can be exported for regional inventory change detection as proxy data

(Klein and Isacks 1999).

34

Chapter 6

Spectral environment

Spectral Characteristics of Snow and Ice

The study of the spectral characteristics of snow and ice is necessary in order to

perform snow cover classification based on spectral signatures. Changes in snow

reflectance and optical properties observed through remotely sensed imagery have been

recognized as indicators of environmental changes associated with local and global

energy fluxes such as mass and energy balance, diagenetic and metamorphic processes,

and facies changes (Hall, Chang, and Siddalingaiah 1988; Hall, Chang, Foster, Benson,

and Kovalick 1989). Reflectance of snow cover is also an important parameter used in

the evaluation of albedo. As discussed earlier, glaciers are characterized by an area of

accumulation and one of ablation. The accumulation area can be further subdivided into

three distinct facies: a dry snow area, usually at higher elevation, where little melting

occurs; a percolation area where local or sporadic melt occurs; and a wet facies

characterized by intense melting during the warm season (Hall et al. 1989). The

identification of the three accumulation facies and the ice cover of the ablation area are

the base parameters for any estimate of mass balance, starting with the simplest

measurement of accumulation area ratio (AAR). To properly identify the snow facies

based on their spectral signatures, it is important to understand the spectral and optical

properties of snow and ice.

35

While of simple chemical composition (H20), snow presents a complex structure

due to the physical properties of water. Water presents a number of triple phase points at

a temperature approaching 273°K (0°C) as a function of the atmospheric pressure (figure

6.1). Triple points are characterized by the presence of coexisting multiple phases (liquid,

vapor and solid) of a substance. Snow is a form of solid precipitation composed of ice

crystals or grains developing from direct conversion of water vapor into ice (McKnight

2000).

The open crystalline structure of snow allows the inclusion of high air content

during the deposition process, leaving enough room for other water phases to exist. As a

consequence, depending on temperature and pressure, snow is characterized by the

presence of ice grains, ice crystals, air and water vapor, and a variable fraction of water.

During deposition, snow develops around nuclei of atmospheric particulates and

chemical impurities such as dust, pollen, soot and ions. The number and types of phases

Figure 6.1. Physical states of H2O (after Martin 2004).

36

present in the snow, the specific structure and density of the snow itself, and the influence

of impurities are important factors that affect the optical properties of snow (Dozier

1989). In fact, the optical properties of snow result from the integration of the optical

properties of all the components present in the snowpack. The complex optical properties

of snow are due to the fact that the resulting combination is not always a linear function

of the fractional components present in the snowpack, but a more complex interaction of

absorption, refractive, and reflective indices of each contributing component (Nolin and

Dozier 2000).

Study of the spectral properties of snow and ice through remote sensing platforms

must be addressed in term of reflectance (this subject will be thoroughly discussed in the

next chapter). Optical satellite imaging registers the amount of energy reflected by an

object on the Earth’s surface. This measurement, also defined as at-sensor-radiance, can

be used to calculate an object’s reflectance. Reflectance is a physical property of a

reflector and is defined by the bidirectional reflectance function (BRDF) (Schott 1997)

][),(),( 100 −= sr

EL

ii φθφθ

ρ (equation 6.1)

where ),( 00 φθL is the radiance from the object toward the sensor (wm-2sr-1µm-1),

and ),( iiE φθ is the incoming solar irradiance (wm-2µm-1) that illuminates the object; both

terms are dependent on the relative azimuth (φ ) and elevation (θ ) of illumination and the

observer’s location. The advantages of using reflectance, rather than recorded pixel

brightness values, are to remove illumination angle effects due to solar elevation and to

37

compensate for the exo-atmospheric irradiance variability registered by optical sensors

along the solar spectrum (Teillet, Barker, Markham, Irish, Fedosejevs, and Storey 2001).

Dozier (1989) observed that the optical properties of ice and water are similar in

the visible and near-infrared wavelengths (VNIR), and that the reflectance of the

snowpack in the VNIR domain is a function of the variations of the refractive index of

ice, the snow grain size, and the percent of impurities present with refractive indices

significantly different from ice. In particular, in the near-infrared wavelengths (NIR: 0.9

µm to 1.3 µm), where ice has significant absorption, snow reflectance depends on the

microstructure of the snow matrix, which is strictly a function of diagenetic processes of

compaction and metamorphic processes that take place during melting-refreezing cycles.

In the visible domain, ice is highly transparent; thus reflectance is inversely proportional

to the amount of impurities (Dozier 1989; Davis, Nolin, Jordan, and Dozier 1993; Nolin

and Dozier 2000). A low, but significant, reduction of reflectance in the visible

wavelengths can be observed when 10 part-per-million by weight (ppmw) of desert dust

or 0.1 ppmw of carbon soot are present as snow impurities (Warren, 1982; Hall et al.

1988).

The variability of the ice absorption index is responsible for the extreme

variability of snow reflectance over the solar irradiance spectrum. In fact, snow

reflectance may vary from nearly complete reflectance to nearly complete absorption

(Nolin and Dozier 2000). In general, snow reflectance decreases as grain size increases

(figure 6.2). Snow reflectance is high in the visible domain, but decreases after

metamorphic processes have occurred (Hall et al. 1988). The NIR domain is particularly

sensitive to grain size, offering a good basis from which to approach remote-sensing

38

based classification of firn and fresh snow over the accumulation zone of glaciers (Hall et

al. 1988). Figure 6.3 shows modeled spectral reflectance for the four glacial facies

defined by Hall et al. (1988; 1989); reflectance decreases as diagenesis and metamorphic

processes transform fresh snow to firn and eventually to glacial ice.

Painter et al. (2003) observed that fresh snowpack presents smaller grain size,

lower density, and lower water content than weathered snow. Firn is the result of

metamorphic and diagenetic processes and is characterized by larger ice grains, higher

density and higher water content. Grain size has a strong negative correlation with

elevation, and thus temperature mean and range. At higher elevation finer grains are

preserved due to the lower rate of melting-refreezing, while at lower elevation large

temperature oscillations allow snow matrix reorganization. Since ice and liquid water

present similar refraction indexes, the presence of liquid water itself does not directly

affect snow reflectance. However, liquid water induces reorganization of the snow matrix

Figure 6.2. Spectral reflectance of snow. Each curve represents the spectrum for a different grain size. The dashed line indicates the center (1.03 µm) of a characteristic absorption band extending from 0.96 µm to 1.08 µm (modified after Nolin and Dozier 2000).

39

structure, resulting in a general increase of snow grain size and lower reflectance (Dozier

1989). Another important observation related to grain size and density is that up to 650

kg/m3, density does not directly affect reflectance, because the center of reflection, the

distance between the grains, is usually greater than the wavelengths of the solar spectrum

(Dozier 1989). As snowpack density increases, grain size changes affect snow reflectance

variations until the packing reaches a threshold at which ice grains fuse together,

obliterating the snow matrix.

As expressed in equation 6.1, the reflectance ratio is a function of the incoming

solar irradiance and of the reflected radiance. Solar irradiance is controlled by the solar

elevation, which determines the amount of energy provided for illumination, while the

radiance is a function of the landcover, observer location, and surface orientation, and is

Ref

lect

ance

Wavelength (µm)

Fresh snow

Firn

Glacier ice

Dirty Glacier ice

Figure 6.3. Modeled spectral reflectance of glacier facies (modified after Hall et al. 1988, Zheng et al. 1984).

40

thus strongly affected by topography. Fresh snow is considered a Lambertian reflector

due to its fine granular structure that diffuses energy equally in all directions (Knap and

Reijmer 1998). As snow ages, diagenetic and metamorphic processes induce grain

growth and uneven layering of the snowpack, and the snow becomes a strong anisotropic

reflector with pronounced forward-scatter and forward-reflection (Knap and Reijmer

1998). Correction based on surface analysis, such as a topographic and anisotropic

normalization model, can be implemented to reduce the effect of forward scattering and

illumination angle (Hall et al. 1988; Dozier 1989). Knap and Reijmer (1998) presented an

alternative solution after observing that, for nadir sensors such as Landsat, the reflectance

error measured over snow and ice fields is consistent and directional and can be solved

by an empirical linear equation that describes the effective reflectance in the form:

),,(),,( φθθφθθρρ ssa c+= (equation 6.2)

where ),,( φθθρ sa represents the apparent at-sensor reflectance registered by the

sensor, and ),,( φθθ sc represents a correction constant obtained from regression (θs is the

solar zenith, and θ and Φ are the bidirectional components of the viewing angle). The

implementation of equation 6.2 is of great impact and benefit for remote sensing

applications for areas where terrain models are unavailable or inaccurate to correct

topographic and anisotropic reflectance of snow and ice cover.

Knap and Reijmer’s (1998) observations are in agreement with Dozier (1989) and

Nolin and Dozier (2000), who suggested that snow reflectance should be modeled as a

multiple scattering problem; the interrelationship between snow reflectance and grain size

is explainable by the Mie scattering theory. The Mie scattering theory and the non-

41

selective scattering model can be applied to calculate scattering and absorption of ice

crystals, assuming the shape of ice grains to be sphere-equivalent (Nolin and Dozier

2000). Mie scattering theory is extremely complex to implement; the main idea is that

Mie scattering occurs when the size of the particle is comparable to the incident

wavelength. The angle and intensity of the scattering is partly a function of wavelength λ,

and presents a characteristic forward-scatter pattern (Schott 1997) (figure 6.4) as

observed by Knap and Reijmer (1998). Non-selective scattering is wavelength

independent and it manifests itself when the size of the particle is considerably greater

than the incident wavelength, resulting in a more uniform scattering of energy in all

directions (Schott 1997).

To model snow reflectance for classification purposes, a number of determining

parameters must be taken in account. Those parameters include ice grain size and

structure (snow matrix), optical properties of the ice grains, chemical-physical

composition of the particulate fractions, illumination and observation geometry, and

energy fluxes.

Direction of incident light

Figure 6.4. Mie Scattering (modified after Nave 2000).

42

Landsat platforms

The first Landsat satellite, originally known as ERTS-1 (Earth Resources

Technology Satellite 1) was launched July 23, 1972. Landsat 1 carried the first MSS

spectrometer (Multi-Spectral Scanner) capable of acquiring three visible channels and

one near-infrared channel. The following missions, Landsat 2 (ERTS B) and Landsat 3,

were launched in January 22, 1975, and March 5, 1978. Landsat 3 was designed to

acquire data in a new thermal band (10.4-12.6 µm wavelength). The MSS spectrometer

provided digital data with a spatial resolution of 79m and a radiometric resolution of 6

bits (later expanded to 7 bits).

During the first half of the 1980s, new micro-technology breakthroughs allowed

the development of a generation of more sophisticated spectrometers. Landsat 4 (July 16,

1982) and Landsat 5 (March 1, 1984) were equipped with the old MSS, as well as the

new Thematic Mapper (TM) spectrometers. Thematic Mapper offered three visible

bands, one near-infrared band, two mid-infrared channels, and a thermal long-wave

infrared band. Some of the most important improvements were the finer spatial resolution

that allowed the TM to sample at 30m in the visible and near/mid-infrared bands, and to

return data with a radiometric resolution of 8 bits.

Landsat 6 was launched on October 5, 1993, but a launch failure resulted in

complete loss of the platform. Landsat 6 carried the Enhanced Thematic Mapper (ETM)

which was designed to provide a panchromatic band at a spatial resolution of 15m. It was

only in April 15, 1999, that the ETM spectrometer became operational with in its new

version (ETM+) on Landsat 7. In addition to the sensors aboard Landsat 5, the ETM+

provides two thermal bands with improved spatial resolution to 60m (with low and high

43

gain mode) and a high-resolution panchromatic band. Table 6.1 provides an overview of

the spatial and spectral specifications of TM and ETM+ spectrometers. In May 31, 2003,

Landsat 7 ETM+ was affected by an instrumentation failure involving the Scan Line

Corrector (SLC); the problem produces an anomaly that invalidates the sampled imagery.

The SLC has been recently remotely deactivated and the validity and quality of the

product are currently under investigation.

TM and ETM+ spectrometers and sensors calibration

Landsat TM and ETM+ do not present significant spectral resolution and spectral

bandwidth differences in the visible, near infrared (NIR), and mid-infrared (MIR)

domains (table 6.1). Although outstanding efforts have been made to maintain calibration

accuracy and over-time consistency between Landsat sensors, there are significant

differences in the radiometric response and detector calibrations that must be taken into

account when cross-correlating reflectance measurements from TM and ETM+ sensors

(Teillet et al. 2001; Chander and Markham 2003; Irish 1998; NASA 2003).

Table 6.1. Spectral and spatial specifications for TM and ETM+ spectrometers (NASA 2003).

Band TM ETM+ TM ETM+1 (Blue) 30 30 0.45-0.52 0.45-0.522 (Green) 30 30 0.52-0.60 0.53-0.613 (Red) 30 30 0.63-0.69 0.63-0.694 (Near IR) 30 30 0.76-0.90 0.78-0.905 (Middle IR) 30 30 1.55-1.75 1.55-1.756 (Thermal IR) 120 60 10.4-12.5 10.4-12.5 LG6 (Thermal IR) 60 10.4-12.5 HG7 (Middle IR) 30 30 2.08-2.35 2.09-2.358 (Panchromatic) 15 0.52-0.90

Spectral resolution (µm)Spatial resolution (m)

44

Figure 6.5 shows the spectral and radiometric response of the TM and ETM+

spectrometers. It can be observed that ETM+ presents a significant improvement on the

band-edge responses compared to TM. The spectral shifting that occurs on the TM sensor

is attributed to filter outgassing, which has been mostly corrected in the ETM+ sensors

(USGS 2003). TM and ETM+ cross-calibration studies that seek to obtain validation of

physical measurements (such as reflectance) have shown that the effects and the

magnitude of the radiometric and spectral differences on measured apparent-reflectance

are dependent on in-scene surface reflectance characteristic (i.e. land cover), and on

variations of exo-atmospheric solar irradiance and atmospheric conditions (Teillet et al.

2001).

Adjustments for spectral and radiometric differences might thus require

information about in-scene spectral content such as surface radiometric measurements.

Teillet et al. (2001) suggested a methodology to obtain calibration coefficients to correct

spectral differences of matching Landsat TM and ETM+ scenes, but emphasized that the

results are, in fact, scene-dependent. The error measured was within 3% for the VNIR

bands, but was less predictable for the MIR bands; these observations must be taken into

account in any attempt to develop standardized cross-platform spectral libraries to be

implemented in change detection and classification studies. Spectral and radiometric

differences, such as those discussed above, are dependent on technological and physical

variables and strictly associated with the sensors’ characteristics. It also is important to

consider those differences that are operator-controlled and are part of acquisition plan

strategies.

45

The Long-Term Acquisition Plan (LTAP) is a global strategy to optimize the use

of Landsat 7 ETM+ to collect and create an archive to document Earth’s land cover

processes (USGS 2003). The LTAP is based on three major sampling strategies that can

be overridden in case of major natural disasters, national security, or user requests. The

first strategy is based on seasonality and is designed to study vegetation distribution and

accomplish change detection. As a consequence, areas of the world that present

Figure 6.5. Relative spectral responses and spectral differences between TM (grey) and ETM+ (black) spectrometers (modified after NASA 2003).

46

interesting vegetation patterns are sampled at the maximum temporal resolution (16

days), while areas subject to low frequency change are sampled rarely. The second

strategy is based on cloud cover. Predicted cloud cover based on NOAA models

influence the acquisition priority of a specific scene. While not directly affecting the

radiometric settings of the platform, the seasonality and cloud cover scheme assumes that

LTAP targets specific land covers. In particular, the spectrometers must be optimized to

fit the reflective properties of vegetation.

The third strategy is the most important in term of defining the spectral and

radiometric settings of the sensors in relation to land cover and in-scene reflective

properties. LTAP defined calibration guidelines that adjust per-band sensitivity of the

ETM+ spectrometers (gain settings) to the expected reflective properties of the scenes;

adjustments are based on geography, seasonality, and land cover.

The gain setting strategy is based on fixed categories: Land (non-desert, non-ice),

Desert, Ice/Snow, Sea Ice, Water, and Volcano/Night. Table 6.2 summarizes the gain

settings based on land cover and geography (solar illumination). Bands 1 to 3 (visible

domain) change gain settings together, and a similar strategy if followed for band 5 and 7

(MIR domain). Each Landsat 7 scene belongs to one of the defined categories. Solar

elevation varies depending on geography (latitude), season, and acquisition time. The

Landsat 7 Science Data Users Handbook (Irish 1998) provides worldwide temporal maps

of the gain rules; for example, figure 6.6 represents the month of August.

47

Categories Exceptions Visible NIR MIR

Land (non-desert, non-ice) High < 45, High HighHigh > 45, Low High

Desert < 28, High < 45, High < 38, High> 28, Low > 45, Low > 38, Low

Ice/Snow/Sea Ice < 19, High < 31, High High> 19, Low > 31, Low High

Water/Coral Reef High High High

Volcano/Night High High Low

(including solar elevation in degrees)

Table 6.2. LTAP gain rules for Landsat ETM+ (modified after NASA 2003).

Figure 6.6. LTAP gain rules for the month of august. The rules depend on expected landcover, seasonality, and illumination. (after NASA 2003).

48

The purpose of the gain rules is to maximize the instrument's 8 bit radiometric

resolution without saturating the detectors (figure 6.7); the low-gain dynamic range is

about 1.5 times that of the high-gain, thus the low-gain setting is applied to regions with

expected high brightness, and the high-gain mode is applied when the in-scene expected

brightness is low (Irish 1998). Table 6.3 and figure 6.7 show that the maximum measured

radiance depends on the gain setting of the spectrometers.

Reflectance measurements depend on solar irradiance; therefore, the maximum

reflectance registered by ETM+ must be calculated on a scene-by-scene basis. Figure 6.8

represents the modeled maximum reflectance measured by ETM+ on the present area of

study (Path 01-Row 71 – 06-26-00). The Long-term Acquisition Program (LTAP)

classified Path 01- Row 71 as Land based on predicted landcover. The solar elevation at

acquisition time was 39.85 degrees. Following the gain rules from table 6.2, it can be

Figure 6.7. Optimization of radiometric resolution based on ETM+ Radiance Dynamic Range adjustments (after NASA 2003).

Low Gain High GainTM1 285.7 190.0TM2 291.3 193.7TM3 225.0 149.6TM4 225.0 149.6TM5 47.3 31.5TM7 16.7 11.1

Table 6.3. ETM+ Radiance Dynamic Range (wm-2sr-1µm-1) (modified after NASA 2003).

49

seen that the gain is set to “high” for all the bands. It can be argued that there is a

significant difference between the represented curves in figure 6.8. As a result, the gain

setting rules have a great impact in classification of features based on spectral signature,

especially when the preset gain mode overrides the spectral reflectance of target

landcovers, such as snow and ice (as will be discussed in the next chapter). Landsat TM

was not subject to any major gain adjustment over mid and low latitudes. The gain

settings for TM are comparable to the low gain settings of ETM+ (table 6.4 and figure

6.9). Figure 6.10 compares the modeled maximum reflectance for Path 01 - Row 71 from

a TM (06-10-97) and ETM+ (06-20-00) scene. Illumination conditions are similar, but

the high gain setting of ETM+ drastically affects the measurable spectral reflectance

range. This difference has a major impact on the definition of exportable snow and ice

spectral libraries between TM and ETM+.

50

0%

20%

40%

60%

80%

100%

120%

Band

Ref

lect

ance

Low Gain 74% 80% 73% 109% 106% 103%

High Gain 49% 53% 49% 73% 71% 69%

TM1 TM2 TM3 TM4 TM5 TM7

Figure 6.8. Modeled maximum reflectance detected by Landsat 7 ETM+ P01R71 06-26-00 for low- gain mode and high-gain mode.

51

Figure 6.9. ETM+ low-gain and TM Radiance Dynamic Range (w•m-2•sr-1•µm-1)(modified after NASA 2003).

0

50

100

150

200

250

300

350

Band

Rad

ianc

e D

ynam

ic R

ange

ETM+ Low Gain 285.7 291.3 225.0 225.0 47.3 16.7

TM 152.1 296.8 204.3 206.2 27.2 14.4

TM1 TM2 TM3 TM4 TM5 TM7

ETM+ TMLow Gain

TM1 285.7 152.1TM2 291.3 296.8TM3 225.0 204.3TM4 225.0 206.2TM5 47.3 27.2TM7 16.7 14.4

Table 6.4. ETM+ low-gain and TM Radiance Dynamic Range (w•m-2•sr-1•µm-1) (modified after NASA 2003).

52

Use of Landsat TM and ETM+ for Snow and Ice Spectrometry

Landsat imagery has been used for snow and ice studies since its first mission in

1972. Landsat TM and ETM+ have been used to study and monitor glaciers and to

measure snow and ice cover extent for global change, hydrology and climatic research.

Nevertheless, Landsat TM and ETM+ were not developed specifically to monitor snow

and ice spectral characteristics, but primarily to monitor vegetation (USGS 2003).

Landsat TM and ETM+ can be employed to monitor glaciers, but a number of important

issues arise when comparing the spectral characteristics of snow and ice to the spectral

0%

20%

40%

60%

80%

100%

Band

Ref

lect

ance

L5 39.7% 83.1% 67.2% 101.8% 64.7% 91.2%

L7 49.3% 54.1% 49.9% 76.4% 69.7% 66.7%

TM1 TM2 TM3 TM4 TM5 TM7

Figure 6.10. Maximum reflectance detected by Landsat 5 TM P01R71 06-10-97 (L5) and Landsat 7 ETM+ P1R71 06-26-00 (L7). The ETM+ is preset by LTAP on high-gain mode (HHHHHH) (modified after NASA 2003).

53

and radiometric capability of the sensors. Landsat ETM+, in particular, presents

important limitations in the discrimination of snow facies when gain settings override the

reflectance characteristic of snow and ice.

Figure 6.11 represents the spectral characteristics of modeled directional-

hemispherical reflectance of four classes of snow based on grain size (Nolin and Dozier

2000); TM and ETM+ bandpasses are represented by gray boxes. The best discriminator

for snow grain size and facies is the wavelength interval from 1.1 to 1.4 µm (MIR), where

the reflectance range is at a maximum (Davis et al. 1993; Nolin and Dozier 2000; Painter

et al. 2003); TM and ETM+ do not provide spectral bands in this interval. Band 4 (NIR)

is the closest option but the reflectance range is considerably reduced and more sensitive

to instrument noise and processing errors. Band 5, while not centered on the highest peak

Figure 6.11. Modeled Directional-Hemispherical reflectance of deep snow (modified after Nolin and Dozier 2000). TM bandpasses are represented by light gray boxes overlapping the spectra. “r” represents grain size.

TM in

tegr

ated

dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

TM bandpass

0

20

40

60

80

100

TM 1 2 3 TM 4 TM 5 TM 7

r = 50 µmr = 200 µm

r = 500 µm

r = 1000 µm

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4Wavelength (µm)

TM in

tegr

ated

dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

TM bandpass

0

20

40

60

80

100

TM 1 2 3 TM 4 TM 5 TM 7

TM in

tegr

ated

dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

TM bandpass

0

20

40

60

80

100

TM 1 2 3 TM 4 TM 5 TM 7

TM in

tegr

ated

dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

TM bandpass

0

20

40

60

80

100

0

20

40

60

80

100

TM 1 2 3 TM 4 TM 5 TM 7TM 1 2 3 TM 4 TM 5 TM 7

r = 50 µmr = 200 µm

r = 500 µm

r = 1000 µm

r = 50 µmr = 200 µm

r = 500 µm

r = 1000 µm

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.40.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4Wavelength (µm)

54

of the far MIR range, occupies a favorable interval and can be used for classification and

discrimination purposes, such as to discriminate clouds from snow due to the high

reflectance of clouds in the MIR. However, Nolin and Dozier (2000) observed that snow

reflectance in this range decreases quickly as grain size increases, approaching low

reflectance values proximate to instrument noise (low signal to noise ratio). Band 7 is

spectrally similar to Band 5. The visible domain, where snow has very high reflectance,

is not sensitive to grain size, and not particularly suitable for snow facies classification.

Figure 6.12 illustrates that TM bands 1 to 3 could potentially be used to distinguish

between snow and glacial ice (Hall et al. 1988) due to the low reflectance of ice and dirt

ice in the ablation zone.

Figure 6.12. Modeled Directional Hemispherical reflectance of glacier facies in TM bandpasses for the visible and NIR domains (modified after Hall et al. 1988, Zheng et al. 1984).

TM bandpass

Dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

0

20

40

60

80

100

TM 1 2 3 TM 4

Fresh snow

Firn

Glacier ice

Dirty Glacier ice

0.4 0.6 0.8 1.0 1.2Wavelength (µm)

TM bandpass

Dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

0

20

40

60

80

100

TM 1 2 3 TM 4

Fresh snow

Firn

Glacier ice

Dirty Glacier ice

0.4 0.6 0.8 1.0 1.2Wavelength (µm)

0.4 0.6 0.8 1.0 1.20.4 0.6 0.8 1.0 1.2Wavelength (µm)

55

Reflectance registered within each TM bandpass can be mathematically integrated

to obtain modeled directional-hemispherical spectral reflectance signatures for snow

(Rosenthal and Dozier 1996) and glacier facies (Hall et al. 1988) as shown in figure 6.13.

Modeled visible and NIR could be effectively used to discriminate glacier facies, but this

model does not consider the cut-off effect induced by the gain settings of TM and ETM+.

The spectral signatures need to be modified in relation to the spectral and radiometric

limitation of Landsat spectrometers.

TM and ETM+ spectral cut-off models for reflectance can be developed from

radiance dynamic ranges (table 6.3 and table 6.4) to produce scene dependent apparent

reflectance ranges (figure 6.8 and figure 6.10). The apparent reflectance obtained is the

Figure 6.13. Integrated spectral directional-hemispherical reflectance of deep snow and glacial facies in TM bandpasses (modified after Zheng et al. 1984, Dozier and Mark 1987, Hall et al. 1988, Rosenthal et al. 1996). TM bandpasses are represented by light gray boxes overlapping the spectra. “r” represents snow grain size.

TM in

tegr

ated

dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

TM bandpass

0

20

40

60

80

100

TM 1 2 3 TM 4 TM 5 TM 7

r = 50 µmr = 100 µmr = 200 µmr = 500 µmr = 1000 µm

FirnGlacial IceDirty Ice

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4Wavelength (µm)

TM in

tegr

ated

dire

ctio

nal-h

emis

pher

ical

Ref

lect

ance

(%)

TM bandpass

0

20

40

60

80

100

TM 1 2 3 TM 4 TM 5 TM 7

r = 50 µmr = 100 µmr = 200 µmr = 500 µmr = 1000 µm

FirnGlacial IceDirty Ice

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4Wavelength (µm)

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.40.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4Wavelength (µm)

56

maximum reflectance detectable by the TM and ETM+ sensors for specific in-scene

illumination angle and solar irradiance, and can be defined as TM and ETM+ reflectance

cut-off curves. In figure 6.14 reflectance cut-off curves (left) based on TM P01-R71 06-

10-97 and ETM+ P01-R71 06-26-00 are intersected with the modeled curves from figure

6.13 to obtain platform dependent spectral signatures of glacier facies (right).

It can be argued that the spectral signature patterns of TM and ETM+ low-gain

are essentially identical except for band 1, leaving open the possibility for the

development of standardized spectral signatures for inter-platform classification

purposes. On the other hand, ETM+ high-gain presents a limited reflectance range that

substantially affects the modeled spectral signature of glacier facies. The resulting

spectral curves allow limited interpretation for detection of snow, firn, and ice in the

visible and NIR domain, leaving only the noise sensitive MIR range for facies

discrimination.

The effect of reflectance compression observed for TM and ETM+ low-gain band

1 and band 3, and for visible and NIR bands for ETM+ high-gain, is described in the

literature as band saturation (Hall et al. 1987, 1988, and 1989; Dozier 1989; Fily,

Bourdelles, Dedieu, and Sergent 1997; Knap and Reijmer 1998; Klein and Isacks 1999).

Band saturation is the physical result of surface reflectance going beyond the detector

capability, and, in case of TM and ETM+, confirms that Landsat platforms were not

developed or programmed to measure snow and ice landcovers over low and mid

latitudes. USGS (2003) reported that the main purpose of Landsat TM and ETM+

missions was to monitor vegetation. The spectrometers were developed to maximize

dynamic response over vegetated areas.

57

Figure 6.14. TM and ETM+ reflectance cut-offs (left) and resulting snow and ice signatures from modeled directional hemispherical reflectance (right). Refer to Figure 6.13 for snow and ice classes legend.

TM

ETM+ Low Gain

ETM+ High Gain

1 2 3 4 5 7TM bandpass

0.5 1.0 1.5 2.0 2.4Wavelength (µm)Wavelength (µm)

0 0.5 1 1.5 2 2.5

0

20

40

60

80

100

120

Ref

lect

ance

(%)

0

20

40

60

80

100

120

Ref

lect

ance

(%)

0

20

40

60

80

100

120

Ref

lect

ance

(%)

TM

ETM+ Low Gain

ETM+ High Gain

1 2 3 4 5 7TM bandpass

1 2 3 4 5 7TM bandpass

0.5 1.0 1.5 2.0 2.40.5 1.0 1.5 2.0 2.4Wavelength (µm)Wavelength (µm)

0 0.5 1 1.5 2 2.50 0.5 1 1.5 2 2.5

0

20

40

60

80

100

120

Ref

lect

ance

(%)

0

20

40

60

80

100

120

Ref

lect

ance

(%)

0

20

40

60

80

100

120

Ref

lect

ance

(%)

0

20

40

60

80

100

120

Ref

lect

ance

(%)

0

20

40

60

80

100

120

Ref

lect

ance

(%)

0

20

40

60

80

100

120

Ref

lect

ance

(%)

58

The most important outcomes of the analysis presented in this chapter are:

The effect of band saturation conceals important spectral information that would

allow a correct distinction between snow and ice facies;

Spectral reflectance signatures based on physical parameters obtained from

Landsat TM may not perform correctly when applied to Landsat ETM+ imagery (and

vice versa) depending on ETM+ gain mode;

For snow and ice classification purposes, Landsat ETM+ in high-gain mode

presents significant limitations for the interpretation of glacier facies due to the high

saturation of the visible and NIR bands and consequent compression of the spectral

signatures in these domains.

59

Chapter 7

Preprocessing

Data Characteristics

Figure 7.1 shows the relative position of the study area within a Path 01 – Row 71

Landsat scene. Five Landsat scenes were selected on the basis of temporal, seasonal,

cloud cover and cost criteria (table 7.1). The temporal criteria were based on TM satellite

image availability, and to allow the comparison of the results with previous field

observations in the area of study for the same time span. The seasonal constraint is linked

to the position of the equilibrium line, which, for this area, corresponds to the limit of the

firn facies at the end of the dry season (winter) (Klein and Isacks 1999).

+ 67.99 W - 15.89

Path 01 - Row 71 UTM WGS 84 Zone 19 S

Cordillera Real

Figure 7.1 Relative position of the Cordillera Real within a Path01-Row71 Landsat scene. Approximate coordinates of the scene center are latitude 15.89 S, longitude 67.99 W.

60

Table 7.2 summarizes the characteristics (metadata) for the five scenes. Three

Landsat TM images were acquired from the Earth Observatory System (EOS) Data

Gateway for $15 each. Of these, the Landsat 4 TM image belonged to the datasets

developed by the Landsat Pathfinder Humid Tropical Forest Project (HTFP), a project

oriented to establish long-term, medium- to high-resolution databases for global change

research (EDC 2003). The 1987 and 1997 Landsat 5 TM scenes were part of the NASA

Landsat Data Collection (NLDC), which contains MSS and TM data purchased by

NASA investigators. NLDC and HTFP make available only those scenes that were

previously used for the specific projects; thus there was a limited selection of scenes that

fit the requirements for this study. For year 1997, the only available image was acquired

in early June, slightly outside the seasonality requirements.

The two Landsat 7 ETM+ scenes were purchased from the USGS at $600 each.

The USGS Landsat ETM+ datasets where selected using the source criteria as the TM

scenes from NLDC and HTFP. Due to the Long-Term Acquisition Plan (LTAP)

priorities, Path 01 – Row 71 scenes are sampled at a lower frequency than the typical 16

day (Landsat 7 maximum temporal resolution).

Least possible due to limited research fundsCost

Less than 1%Cloud cover (on study area)

austral winter; late winter if possibleSeasonality

from mid 80s to todayTemporal Distribution

Landsat dataset query constraints

Least possible due to limited research fundsCost

Less than 1%Cloud cover (on study area)

austral winter; late winter if possibleSeasonality

from mid 80s to todayTemporal Distribution

Landsat dataset query constraints

Table 7.1. Constraints applied to the selection of suitable Landsat scenes

61

Table 7.2 Metadata summary for the selected Landsat scenes.

ID 5001071008721410Platform Landsat 5

Sensor TMAcquired 8/2/1987

Cloud Cover 10%(over study area) < 1%

Sun Elevation 39.23Sun Azimuth 49.68

ID 4001071008921110Platform Landsat 4

Sensor TMAcquired 7/30/1989

Cloud Cover 10%(over study area) < 1%

Sun Elevation 40.00Sun Azimuth 49.00

ID 5001071009716110Platform Landsat 5

Sensor TMAcquired 6/10/1997

Cloud Cover 20%(over study area) < 1%

Sun Elevation 37.79Sun Azimuth 42.90

ID 7001071000017850Platform Landsat 7

Sensor ETM+Acquired 6/26/2000

Cloud Cover 31%(over study area) < 1%

Sun Elevation 39.85Sun Azimuth 39.44

ID 7001071000121250Platform Landsat 7

Sensor ETM+Acquired 7/31/2001

Cloud Cover 32%(over study area) < 1%

Sun Elevation 42.59Sun Azimuth 45.58

62

As a consequence, there is a very limited amount of imagery available once

proper cloud cover and the narrow 3 month window necessary to study glacier

phenomena are considered. For the year 2000, for example, an early winter scene was the

only image available.

Landsat images are available to the public in three preprocessed versions. Level

0R (L0R) is a raw format provided along with ancillary acquisition information allowing

the user to perform first order processing. ETM+ imagery was initially meant to be

provided only in L0R format, but marketing and research demand forced NASA to

systematically correct the data to ensure availability and accuracy (Irish 1998). The

systematic correction applied to L0R is the Level 1R (L1R), which consists of

radiometrically corrected imagery. A further processing step is available through Level

1G (L1G), which provides systematically and geometrically corrected data. NLDC and

HTFP TM scenes are available in L1R or L1G format, depending on the original project;

the ETM+ scenes were purchased in L1G format to be used as a reference for geometric

correction and co-registration of all the other images.

Note:

For more information on HTFP and NLDC projects refer to respectively:

http://edcdaac.usgs.gov/pathfinder/pathpage.html

and

http://edcdaac.usgs.gov/pathfinder/nldc/nldc.html.

63

Geometric Correction and Co-registration

Geometric correction assigns scale and projection properties to remotely sensed

data, while co-registration is a process of matching the coordinate systems of two or more

images collected at different times or by different sensors (Mather 1999). Geometric

correction and co-registration of the five images was necessary in order to perform a

meaningful comparison of the distribution of the final snow and ice classes from each

scene (Mather 1999); the analysis of the dynamic variations of the glaciers occurred in

the specific period of time elapsed between the acquisition of the scenes, and is based on

quantified changes of snow and ice facies.

The first geometrically corrected image acquired for this project was the 2001

ETM+ image. This image was used as master image to co-register and geometrically

correct the remaining four scenes using the image-to-image geometric correction tool of

ERDAS Imagine 8.6. The correction was based on a final selection of 25 to 30 suitable

ground control points (GCPs) on the ETM+ image (figure 7.2); the operation was

rendered particularly difficult due to the lack of clearly defined anthropogenic features.

The final Total Root Means Square (TRMS) error obtained for each image was below

0.5, corresponding to less than one-half the cell size (14.25m) (ERDAS 1997). The best

results were obtained by distributing the control points in the vicinity of the study area.

While still providing a TRMS below 0.5, the initial attempt to select control points

distributed over the whole scene resulted in noticeable local error over the study area.

The final spatial adjustment was obtained using a first order polynomial rigorous

transformation and a nearest neighbor (NN) re-sampling algorithm; the resulting four co-

registered images had a spatial resolution of 28.5m to match the 2001 ETM+ cell size

64

(figure 7.2). Nearest Neighbor resampling ensured that the final images preserved

original digital values; other algorithms, such as bilinear interpolation and cubic

convolution, substitute the original data with interpolated or spatially averaged digital

values, resulting in the loss of the original information and introduction of digital artifacts

(Mather 1999).

Two important considerations must be taken into account: (1) the USGS ETM+

L1G scenes (geometrically corrected from the source) were expected to be spatially co-

registered, but in fact, appeared not to correspond geometrically (figure 7.4); the 2000

image had to be co-registered to the 2001 image as were the other images; (2) geometric

co-registration does not correct for relief effects and distortion induced by the angle of

view of the sensor, but assures that all the images represent the same geometric structure.

Figure 7.2. ERDAS Image 8.6 Geometric Correction tool. Example of Ground Control Points (GCP) selected in proximity to the study area.

65

Figure 7.3. Geometric correction and co-registration of Landsat TM Path 01 – Row 71 8/2/1987. Before (left) and after (right). The original image was a L1R product.

Figure 7.4. Landsat ETM+ Path 01 – Row 71 6/26/2000 and 7/31/2001. Spatial offset at La Paz Airport (approx. 200 m). Both scenes were originally projected in UTM 19S WGS84.

ETM+ 2001 ETM+ 2000

2001

2000

66

Study Area Subset

The glaciers of the Cordillera Real have been studied and catalogued since the

1970s. Jordan (1991 and 1999) and colleagues (Jordan at al. 1980) produced an inventory

of Bolivian glaciers available from the USGS publication Satellite Image Atlas of

Glaciers of the World - SOUTH AMERICA (USGS 1999). Jordan’s inventory is available

as a vector file in ArcInfo E00 format (figure 7.5). The polygons describing the glaciers

can be used to define spatial reference units for zonal statistics and to filter out ice and

snow classified covers that do not represent glaciers.

The vector polygons containing the spatial description of the glaciers were

rasterized and converted to a binary mask (value 1 = glacier, value 0 = non-glacier) and a

zonal layer (zones defined by attribute USGS_ID of the glacier inventory); a three pixel

buffer to better evaluate the transition from glacier to non-glacier classes was added to

the binary mask. The five Landsat scenes were then clipped with the binary mask to

obtain subset images of the glaciers of the Cordillera Real (figure 7.6). Removing

unnecessary data from the image also shortens the computational time of the

classification process.

67

Number of glaciers: 893

Smallest Glacier (sqm) 923.1

LargestGlacier (sqm) 8,298,608.0

Number of glaciers: 893

Smallest Glacier (sqm) 923.1

LargestGlacier (sqm) 8,298,608.0

Number of glaciers: 893

Smallest Glacier (sqm) 923.1

LargestGlacier (sqm) 8,298,608.0

Number of glaciers: 893

Smallest Glacier (sqm) 923.1

LargestGlacier (sqm) 8,298,608.0

Figure 7.5. Glaciers of the Cordillera Real and detail of the Ancohuma-Illampu Massif. Area is calculated as planimetric surface. Glacier classification by Jordan (1999).

68

Inventory

Binary mask

Landsat Source

Glaciers

Figure 7.6. Details of the subset procedure based on Jordan’s glacier inventory (1999).

69

Atmospheric Correction

Atmospheric correction of remotely sensed imagery is needed when performing

measurements seeking to determine a physical property of the surface of the Earth, such

as reflectance, and when it is necessary to standardize the at-sensor radiance to compare

imagery acquired with different atmospheric conditions (Mather 1999). To obtain a

measurement of the effective radiance leaving a surface and calculate its true reflective

properties, it is necessary to remove the optical effects induced by the atmosphere. Figure

7.7 shows the solar energy paths through the atmosphere and table 7.3 summarizes the

characteristics of each path. Thermal radiance terms will not be considered in this

research since they do not influence the energy measured in the visible, NIR, and MIR

domains of the Landsat spectrometers. There are two main types of radiance: (1) surface-

reflected solar energy that reaches the sensor

(LA); and (2) path radiance, which includes

radiances of type B, C, G, and I, and represents

the amount of solar energy that reaches the

sensor, but is not a function of the surface

properties (Schott 1997). Atmospheric

correction algorithms and models aim to

remove the effect of path radiance from

remotely sensed imagery. Figure 7.7. Solar energy path (modified after Schott 1997)

70

The relationship between effective radiance (LA), path radiance (LP = LB + LC +

LG +LI), and at-sensor (or apparent) radiance (LS) can be generalized as

LS = LP + LA; (equation 7.1)

This equation describes the common observation that during hazy days the

apparent radiance (LS) increases as scattering (haziness) increases (figure 7.8). The

equation also describes the effect of diffuse illumination in shadows, where, in the

absence of direct illumination, LA = 0 and LS = LP, radiance becomes a direct function of

scattering and background reflection. Using LOWTRAN models, Schott (1997) observed

that, in the visible and NIR domain, the G and I terms are significantly smaller than A, B,

and C. The magnitude of the path radiance versus effective radiance varies, depending on

the reflectance of the surface. Path radiance significantly increases the radiance of

surfaces with low reflectance values (reflectance lower than 5% in figure 7.8) and the

effect of path radiance is gradually reduced by high surface reflectors.

Table 7.3. Solar energy path (modified after Schott 1997).

Path Radiance Type Description

Object RadianceA L A Effective Radiance Measurement of solar energy directly reflected by an objectD L D Self emitted Radiance Radiance from the object, self emission, usually thermal

Downwelled RadianceB L B Skylight Radiance Reflected radiance from atmospheric scatteringE L E Sky thermal Radiance Reflected thermal radiance from atmospheric self emission

Upwelled RadianceC L C Uppwelled Radiance Solar energy scattered by the atmosphere reaching the sensorF L F Upwelled thermal Radiance Self emitted atmospheric thermal energy reaching the sensor

Background RadianceG L G Background Radiance Illumination provided by neighboring surfaces H L H Background thermal Radiance Thermal energy from neighboring surfaces I L I Adjacency Effect Scattered radiance from neighborsing surfaces

71

For dark surfaces, with reflectance from 1% to 3%, the path radiance may

represent 50% of the total at-sensor radiance; as a consequence, a small error in the

computation of upwelled radiance results in a large error of estimate of the effective

radiance (Schott 1997). For highly reflective surfaces such as fresh snow, path radiance is

a relatively minor component of total radiance. The magnitude of path radiance is also a

function of atmospheric conditions; Schott (1997) observed that in clear sky conditions,

the A term is up to seven times greater than the downwelled radiance, but the terms

become comparable in hazy atmospheric conditions. It is thus very important to account

for the reflective properties of the investigated surface and the atmospheric condition at

the time of acquisition when generalizing atmospheric correction models to be applied to

satellite imagery.

Figure 7.8. Effects of atmospheric transmission and path radiance on resulting total at-sensor reflectance (modified after Schott 1997).

Radiance leaving the ground (W*cm-2*sr-1)0.001 0.002 0.003 0.004 0.0050

Rad

ianc

e re

achi

ng th

e gr

ound

(W*c

m-2*s

r-1)

0.001

0.003

0.002

0.005

0.004

0

Ideal atmosphere

Visibility = 20 km

Visibility = 5 km

Visibility = 2.5 km

Visibility = 1 km

Radiance leaving the ground (W*cm-2*sr-1)0.001 0.002 0.003 0.004 0.0050

Rad

ianc

e re

achi

ng th

e gr

ound

(W*c

m-2*s

r-1)

0.001

0.003

0.002

0.005

0.004

0

Radiance leaving the ground (W*cm-2*sr-1)0.001 0.002 0.003 0.004 0.0050

Radiance leaving the ground (W*cm-2*sr-1)0.001 0.002 0.003 0.004 0.0050

Rad

ianc

e re

achi

ng th

e gr

ound

(W*c

m-2*s

r-1)

0.001

0.003

0.002

0.005

0.004

0

Rad

ianc

e re

achi

ng th

e gr

ound

(W*c

m-2*s

r-1)

0.001

0.003

0.002

0.005

0.004

0

Rad

ianc

e re

achi

ng th

e gr

ound

(W*c

m-2*s

r-1)

0.001

0.003

0.002

0.005

0.004

0

0.001

0.003

0.002

0.005

0.004

0

Ideal atmosphere

Visibility = 20 km

Visibility = 5 km

Visibility = 2.5 km

Visibility = 1 km

Ideal atmosphere

Visibility = 20 km

Visibility = 5 km

Visibility = 2.5 km

Visibility = 1 km

Reflectance = 5%

Reflectance = 25%

72

The effective radiance LA for a specific wavelength can be described as a function

of the total incoming solar energy (Htot), the atmospheric transmittance (T) in the specific

wavelength, and the reflective properties of the irradiated surface (Mather 1999):

LA = HtotρAT; (equation 7.2)

and rearranging

THL

tot

AA =ρ (equation 7.3)

where ρA describes the effective reflectance in absence of atmospheric path

radiance. By definition, for full atmospheric transmittance (T = 1), equation 7.3

represents the theoretical optical reflectance of a surface, described by the ratio of

outgoing exitance and incoming irradiance (equation 6.1). The value of this ratio is a

physical property of a surface, and the desired target measurement of remote sensing

studies. To better describe the contribution of the path radiance to at-sensor

measurements, equation 7.2 can be substituted to LA in equation 7.1

LS = LP + HtotρAT (equation 7.4)

and then solved for effective reflectance (ρA)

THLL

tot

PSA

)( −=ρ (equation 7.5)

Equation 7.5 can be further elaborated to obtain

TH

LTH

L

tot

P

tot

SA −=ρ . (equation 7.6)

73

Since in can be assumed that (Schott 1997; Mather 1999)

THL

tot

SS =ρ , (equation 7.7)

it possible to write equation 7.6 as

cSA −= ρρ , where TH

Lc

tot

P= . (equation 7.8)

Equation 7.5 and equation 7.8 are of utmost importance in the solution of

atmospheric correction models. Equation 7.5 is commonly applied to in-scene

atmospheric correction models (such as the Dark Object Subtraction method) and

illustrates that the effective reflectance (ρA) of an object can be obtained by subtracting

the estimated path radiance (LP) from the at-sensor radiance measurements (LS). Equation

7.8 shows that the effective reflectance (ρA) can also be obtained by subtracting a

correction value (c) from at-sensor apparent reflectance measurements (ρS). Three main

methodologies are available to estimate LP and c: ground station corrections, atmospheric

propagation models, and in-scene calibration.

Measurement of reflectance by using onsite spectrometers is the most accurate

procedure to calculate c, and is based on the comparison of ground reflectivity

measurements (ρG) and at-sensor reflectance. The difference between the measurements

is the path radiance:

c = ρS – ρG; (equation 7.9)

74

The obtained constant is a direct function of LP, because the ground

measurements are corrected for all the components of the path radiance. While extremely

accurate, this procedure requires field measurements using costly equipment and the

availability of spectrally pure target surfaces in the satellite image to avoid spectral

mixing. This method is also difficult to implement in remote and less accessible areas.

Atmospheric propagation models (APM) can be used to estimate atmospheric

radiative transfer properties from field measured parameters or from standardized

atmospheric conditions (Mather 1999). Atmospheric propagation models can be

computationally intensive and a number of software packages have been developed to

produce the most accurate atmospheric modeling, such as LOWTRAN, MODTRAN, and

5S/6S. Atmospheric propagation models are platform- and scene-independent and can be

applied to an area with little or no information about the atmospheric conditions present.

Processing parameters from field measurements that describe the atmospheric conditions

at acquisition time provides extremely accurate, but extremely costly results; therefore,

processing of standardized atmospheric conditions is more likely to be applied to the

study of remote areas, such as alpine environments (Schott 1997). Schott also observed

that APMs offer a range of acceptable results based on the quality of the input

parameters, but warned that the concept of “garbage in, garbage out” is of pertinent

application in APMs processing. Figure 7.9 shows an example of an online MODTRAN3

application offered by the Department of Geophysical Sciences at the University of

Chicago (http://geosci.uchicago.edu/~archer/cgimodels/radiation.html).

Atmospheric corrections that take advantage of the characteristics of the data

within multispectral imagery are known as in-scene calibrations (Schott 1997). In-scene

75

calibrations may be solved by comparing imagery from different altitudes, different

angles or different wavelengths to solve equation 7.8. Dark object subtraction (DOS)

(Chavez 1988) is an in-scene calibration methodology commonly used to correct satellite

imagery for path radiance. The theory behind DOS is based on the concept expressed in

equation 7.5 that the effective reflectance (ρA) can be obtained by subtracting the path

Model Output

Intensity (W/(m2-sr-wavenumber)) vs. wavenumber Atmospheric Profiles vs. Height, km

MODTRAN 3

Atmospheric Longwave Radiation Modeler

University of ChicagoDepartment of Geophysical Science

Model Input •pCO2 = 330 •CH4 = 1.7 •Tropospheric Ozone (ppb) = 28 •Stratospheric Ozone Scale Factor = 1 •T offset = 0 •Constant Water Vapor Pressure •H2O vapor twiddle factor = 1 •Tropical Atmosphere Model •No Clouds •Sensor Altitude = 750 Looking down •User Supplied Tag = Christian+Degrassi •Model Run Tag Number, 19133837

Figure 7.9. MODTRAN 3. Example of longwave atmospheric transfer model output from standardized input (http://geosci.uchicago.edu/~archer/cgimodels/radiation.html).

76

radiance (LP) from at-sensor apparent radiance measurements (LS), or adjusting the

apparent reflectance (ρS) by a correction factor c (equation 7.8). In the absence of

external references to solve equation 7.9, the only way to establish a radiance absolute

reference is to identify an in-scene feature that presents known spectral characteristics.

Such opportunities are offered by dark objects, features that, in absence of path radiance,

would present radiance equal or proximate to zero radiance (thus appearing black):

c(LP) = ρSDO – ρADO (equation 7.10)

where ρSDO is the apparent reflectance of a dark object and ρADO is the effective

reflectance of a dark object, which is equal to 0%; thus for ρADO = 0

c = ρSDO (equation 7.11)

Shadowy, clear lakes or recent basalt flows are examples of dark objects. The

correction factor (c) obtained equals the apparent reflectance of a dark object (Chavez

1989). This is in line with the practical observation that shadows and black objects appear

brighter in presence of haze (scattering). For a dark object, the apparent brightness is

produced by the path radiance.

While simple in theory, the Dark Object Subtraction (DOS) method requires an

accurate selection of the dark features to avoid under- and over-estimation of the

correction factor. Chavez (1988) explained that there is a high probability that at least a

few pixels within an image will present 0% reflectance, such as shadows from

topography or clouds. There are a number of techniques that can be implemented to

extract and model the correction factor for DOS calibration. The simplest of these

77

methods is based on the analysis of the band frequency histograms; a more complex

methodology takes into account the central wavelength of each multispectral band to

calculate atmospheric scattering effects (Chavez 1988). Other DOS models include a

more thorough correction for atmospheric absorption, transmittance and downwelling

radiance (Moran, Jackson, Slater and Teillet 1992; Chavez 1996). Three DOS techniques

were coordinated and implemented to remove path radiance from the Landsat TM and

ETM+ scenes of the Cordillera Real in this project.

A first order DOS correction was performed by identifying the lowest non-zero

frequency for each of the band histograms, assuming that it represented the shifted

radiance of the darkest object in the scene (Chavez 1988). However, very low non-zero

histogram values are sensitive to instrument noise and may provide a false dark object,

resulting in underestimation of the path radiance. Selecting the frequency at which the

histogram slope starts increasing more consistently may help to filter out background and

instrument noise, but the selection of the ideal frequency can be highly biased, subject to

analyst interpretation, and may result in inconsistent under- or over-estimation of the path

radiance. The method used in this research was based on the identification of in-scene

dark targets, such as lakes, deep topographic shadows and cloud shadows. A maximum of

200 of the darkest object pixels were sampled and analyzed through descriptive statistics.

Table 7.4 shows the resulting final set of DOS values for the five images. The

atmospheric conditions, and thus the path radiances, are very similar in the five images.

The main difference is observed between TM and ETM+; this is due differences in sensor

calibration and spectral response, as discussed in chapter 7.

78

Chavez (1988) argued that the selection of uncorrelated Dark object Subtraction

values from the spectral bands might not reflect the real relative scattering model; in fact,

the manual selection of independent DOS band values does not have a physical

foundation, resulting in a possible distortion of the natural atmospheric scattering

phenomenon. It is possible to increase DOS performance by introducing an atmospheric

scattering correction model to obtain a correction consistent with scattering optical and

physical phenomena. Relative atmospheric scattering models assume that scattering is

inversely proportional to the wavelength by a power factor dependent on the type of

scattering. Rayleigh scattering is inversely proportional to the fourth power of the

wavelength, while the Mie scattering power factor may vary between 4, representing

clear atmospheric conditions, and 0 for complete cloud cover or maximum scattering

(Chavez 1988). Chavez suggested that for observed real atmospheric conditions, the

power factor ranges within 2 and 0.7, representing a mixture of Rayleigh and Mie

scattering, and between 0.7 and 0.5 for increasing hazy conditions. Dark Object

Subtraction techniques cannot correct for heavy haze conditions (power factor lower than

0.5). The relative scattering model uses the central wavelength of each TM band, a

starting DOS value from one representative band, and a standardized model for

atmospheric conditions to estimate DOS correction for all the bands (Chavez 1988).

1987 1989 1997 2000 2001TM1 40.0 35.0 36.0 36.0 37.0TM2 11.0 11.0 11.0 22.0 23.0TM3 5.0 7.0 6.0 14.0 15.0TM4 3.0 3.0 3.0 9.0 10.0TM5 1.0 1.0 1.0 7.0 7.0TM7 1.0 1.0 1.0 6.0 6.0

TM ETM+

Table 7.4. Manually selected DOS values based on the analysis of dark targets present in the five scenes. The values are represented as digital numbers.

79

Table 7.5 shows the power factor corrections corresponding to five standardized

atmospheric scattering models. Table 7.6 illustrates the resulting coefficients estimated

for the atmospheric scattering model for each Landsat TM and ETM+ band obtained

substituting the central wavelength of the bands for each atmospheric scattering model.

TM and ETM+ band 1 DOS values from table 7.4 were used as the reference starting

values to estimate the relative scattering for each band.

Table 7.7 illustrates, as an example, the calculations and results for the 1987

Landsat TM image used in this research. It can be observed that the scattering function

coefficients for each atmospheric condition are normalized on band 1 (the selected

reference band). The predicted haze values for each band and atmospheric condition are

calculated by multiplying the observed reference DOS value from band 1 (40 in the

Table 7.5. Atmospheric scattering models (after Chavez 1988).

Atmospheric Models

Wavelength Power Factor

Very Clear λ-4

Clear λ-2

Moderate λ-1

Hazy λ-0.7

Very Hazy λ-0.5

Table 7.6. Scattering function coefficients obtained for each band in different atmospheric conditions (methodology after Chavez 1988).

Min Max Central Very Clear Clear Moderate Hazy Very HazyTM1 0.450 0.520 0.485 TM1 18.073 4.251 2.062 1.660 1.436TM2 0.520 0.600 0.560 TM2 10.168 3.189 1.786 1.501 1.336TM3 0.630 0.690 0.660 TM3 5.270 2.296 1.515 1.338 1.231TM4 0.760 0.900 0.830 TM4 2.107 1.452 1.205 1.139 1.098TM5 1.550 1.750 1.650 TM5 0.135 0.367 0.606 0.704 0.778TM7 2.080 2.350 2.215 TM7 0.042 0.204 0.451 0.573 0.672

ETM+1 0.450 0.520 0.485 ETM+1 18.073 4.251 2.062 1.660 1.436ETM+2 0.530 0.610 0.570 ETM+2 9.473 3.078 1.754 1.482 1.325ETM+3 0.630 0.690 0.660 ETM+3 5.270 2.296 1.515 1.338 1.231ETM+4 0.780 0.900 0.840 ETM+4 2.009 1.417 1.190 1.130 1.091ETM+5 1.550 1.750 1.650 ETM+5 0.135 0.367 0.606 0.704 0.778ETM+7 2.090 2.350 2.220 ETM+7 0.041 0.203 0.450 0.572 0.671

Scattering FunctionWavelength

80

example) by the normalized coefficient. The atmospheric conditions for the five Landsat

scenes used in this research study were estimated to be “very clear” under the conditions

proposed by Chavez’ models (Chavez 1988).

A third correction can improve the quantification of DOS values. Chavez (1989)

suggested that natural surfaces rarely present a perfect 0% reflectance (complete

absorption) and are more likely to show a minimum of 1% to 2% reflectance. For each

TM and ETM+ band it is possible to estimate the radiance value (or digital number)

corresponding to 1% reflectance. The number is then added to the previously estimated

DOS values. Table 7.8 summarizes the results for the five scenes after adjusting the

predicted values for sensor calibration, band specific spectral responses, and 1%

minimum reflectance.

Table 7.7. Procedure to predict scattering from TM band 1 reference DOS (methodology after Chavez 1988).

min max avg Very Clear Clear Moderate Hazy Very HazyTM1 0.450 0.520 0.485 TM1 18.073 4.251 2.062 1.660 1.436TM2 0.520 0.600 0.560 TM2 10.168 3.189 1.786 1.501 1.336TM3 0.630 0.690 0.660 TM3 5.270 2.296 1.515 1.338 1.231TM4 0.760 0.900 0.830 TM4 2.107 1.452 1.205 1.139 1.098TM5 1.550 1.750 1.650 TM5 0.135 0.367 0.606 0.704 0.778TM7 2.080 2.350 2.215 TM7 0.042 0.204 0.451 0.573 0.672

Very Clear Clear Moderate Hazy Very HazyTM1 1.000 1.000 1.000 1.000 1.000TM2 0.563 0.750 0.866 0.904 0.931TM3 0.292 0.540 0.735 0.806 0.857TM4 0.117 0.341 0.584 0.687 0.764TM5 0.007 0.086 0.294 0.424 0.542TM7 0.002 0.048 0.219 0.345 0.468

Reference Very Clear Clear Moderate Hazy Very HazyTM1 40 40 TM1 40.000 40.000 40.000 40.000 40.000TM2 11 TM2 22.505 30.003 34.643 36.170 37.225TM3 5 TM3 11.664 21.600 29.394 32.240 34.289TM4 3 TM4 4.664 13.658 23.373 27.462 30.577TM5 1 TM5 0.299 3.456 11.758 16.976 21.686TM7 1 TM7 0.092 1.918 8.758 13.814 18.717

Observed 1987

Haze Function

Normalized Haze Function Model

Predicted Value

Wavelength

81

Chavez (1989) argued that the predicted values for bands TM5 and TM7 are

usually too high due to sensor calibration parameters; a DOS equal to 1 should be used to

correct MIR TM bands. Unfortunately there is not sufficient literature discussing

predicted values for ETM+; therefore, in this project, the observed values were

substituted for those predicted for ETM+ 5 and ETM+ 7.

Image Standardization

Change detection and quantification of physical parameters of the Earth’s

environment depend on the ability of satellite sensors to provide calibrated and consistent

measurements (Markham and Chander 2003). Huang, Yang, Homer, Wylie, Vogelman

and DeFelice (2003) argued that one of the major challenges to obtaining satellite based

change detection is the removal of noise arising from sensor calibrations, instrument

errors, atmospheric effects, and illumination geometry; such noise may render unreliable

and inconsistent the information extracted from satellite imagery. Preprocessing

Table 7.8. Observed DOS and final predicted DOS values.

Observed Predicted Observed Predicted Observed PredictedTM 1 40 40 35 35 36 36TM 2 11 13 11 12 11 12TM 3 5 10 7 9 6 9TM 4 3 5 3 5 3 5TM 5 1 5 1 5 1 5TM 7 1 4 1 3 1 3

2000Observed Predicted Observed Predicted

ETM +1 36 36 37 37ETM +2 22 22 23 23ETM +3 14 18 15 19ETM +4 9 12 10 12ETM +5 7 9 7 9ETM +7 6 9 6 9

1987 1989 1997

2001

82

techniques help correct for calibration and atmospheric effects, but a further step must be

taken to standardize information before change detection analysis can take place. Huang

et al. (2003) showed that converting digital numbers (DN) to at-satellite reflectance not

only provides a physical parameter for describing land surfaces, but also provides more

relevant temporal information about the target (such as illumination). At-satellite

reflectance transformation does not introduce error and its use makes it possible to

standardize images that are separated spatially (mosaic) or temporally (change detection).

Markham and Barker (1986) described the methodology for obtaining at-sensor

reflectance from DN. This methodology was successfully applied to the five Landsat

scenes used in this thesis research. The first step (equation 7.12) consists of converting

DN to at-sensor radiance, a physical measure of the energy registered by the satellite

sensors:

Lλ = Gainλ * DN + Biasλ (equation 7.12)

where λ represents the spectral band and Gain and Bias are post-launch

calibration parameters provided in the Landsat scene header files. Gain and Bias can be

calculated from post-calibration coefficients as shown in equation 7.13 and 7.14.

CALMAX

MINMAXA Q

LLGain )( −= (equation 7.13)

MINA LBias = (equation 7.14)

where QCALMAX is the range of rescaled values (usually 255), LMIN is the spectral

radiance for DN = 0, and LMAX is the spectral radiance for QCALMAX. Table 7.9 summarizes

Gain and Bias pairs, and post-calibration LMIN and LMAX for each TM and ETM+ band.

83

The second step is to transform at-sensor radiance to at-sensor reflectance.

Reflectance is a measurement of the relationship between incoming solar irradiance and

exiting radiance from a surface (indicated by equation 6.1 and equation 7.3). The

bidirectional reflectance factor (BDRF; ρBDRF) simplifies the bidirectional reflectance

distribution function (BRDF; ρBRDF) to a unitless measure of diffuse radiance (Schott

1997) and corresponds to at-sensor reflectance (ρS) measurements:

ρBDRF = π ρBRDF = ρS (equation 7.15)

The ρBRDF term corresponds to equation 6.1 and can be substituted in equation

7.15 to obtain a general representation of unitless at-sensor reflectance:

),(),( 00

iiS E

Lφθφθπρ = (equation 7.16)

From the concept expressed by equation 7.16, Markham and Barker (1986) developed

equation 7.17 to calculate the resulting at-sensor reflectance for Landsat imagery:

θπρ

cos

2

SUN

SS E

dL= ; (equation 7.17)

Table 7.9. Post-calibration parameters (after Irish 1998).

Gain Bias Lmin LmaxTM1 0.6024314 -1.5200000 -1.5200000 152.1000070TM2 1.1750981 -2.8399999 -2.8399999 296.8100156TM3 0.8057647 -1.1700000 -1.1700000 204.2999985TM4 0.8145490 -1.5100000 -1.5100000 206.1999950TM5 0.1080784 -0.3700000 -0.3700000 27.1899920TM6 0.0569804 -0.1500000 -0.1500000 14.3800020

ETM+1 0.7756863 -6.1999969 -6.1999969 191.6000096ETM+2 0.7956862 -6.3999939 -6.3999939 196.4999871ETM+3 0.6192157 -5.0000000 -5.0000000 152.9000035ETM+4 0.6372549 -5.1000061 -5.1000061 157.3999934ETM+5 0.1257255 -0.9999981 -0.9999981 31.0600044ETM+7 0.0437255 -0.3500004 -0.3500004 10.8000021

84

where ESUN represents the incoming solar irradiance for a specific TM and ETM+

band (table 7.10), and d is the date-related Earth-Sun distance. It is necessary to factor in

a correction of up to 3.5% for changes in solar irradiance due to the eccentric orbit of the

Earth (table 7.11). The cosine of the solar zenith (cosθ) is introduced to remove the angle

of illumination effect.

It is important to emphasize that equation 7.17 assumes that the reflector is a flat

Lambertian surface (Markham and Barker 1986). However, only fresh snow cover

presents Lambertian reflectance characteristics. Topographic normalization and

anisotropic reflectance corrections may be applied to remove relief and reflective errors

(Colby 1991) using DEMs and Minnaert coefficients to normalize the measured at-sensor

radiance. Due to the lack of accurate DEMs for the study area, these corrections could not

be performed and the inherent errors must be taken into account during the analysis of the

classification results. (Appendix A provides a case study that discusses the application of

topographic normalization and anisotropic reflectance to a small subset of the study area).

Table 7.10. Mean Solar Exo-atmospheric irradiances (Wcm-2µm-1) (values after Irish 1998).

Landsat 4 Landsat 5 Landsat 7TM1 1957.00 1957.00 ETM+1 1969.00TM2 1825.00 1826.00 ETM+2 1840.00TM3 1557.00 1554.00 ETM+3 1551.00TM4 1033.00 1036.00 ETM+4 1044.00TM5 214.90 215.00 ETM+5 225.70TM6 80.72 80.67 ETM+7 82.07

Mean Solar Exoatmospheric irradiances

Date Julian Day Distance d Sun ElevationTM 1987 08/02/87 214 1.014627391 39.23TM 1989 07/30/89 211 1.015023836 40.00TM 1997 06/10/97 161 1.015323647 37.79ETM+ 2000 06/26/00 178 1.016587389 39.85ETM+ 2001 07/31/01 212 1.014896088 42.59

Table 7.11. Earth-Sun distance (astronomic units AU) and Sun elevation (degrees) at the time of acquisition (distance d calculated after Irish 1998).

85

Chapter 8

Spectral Analysis and Classification

Spectral Characteristics of the Subscenes

After preprocessing, descriptive statistics were calculated for the resulting five

Landsat scene subsets to obtain information on the spectral characteristics of each image.

In particular, the statistics concerning band saturation provided useful information

regarding the separability of snow facies in the accumulation zone (as discussed in

chapter 6). Table 8.1 and figure 8.1 compare the maximum reflectance measured by the

TM and ETM+ sensors for the five subscenes. It can be observed that in the ETM+

scenes (2000 and 2001) the maximum reflectance measured is significantly lower that

that measured by TM spectrometers. This is due to the high-gain mode setting employed

on the ETM+ spectrometer and it confirms the theoretical spectral response of ETM+

described in chapter 6. This observation also raises a question related to the amount of

information that has actually been compressed and obliterated by the band saturation

effect.

Table 8.2 and figure 8.2 show the statistics describing the band saturation effect

for the subscenes. ETM+ visible (VIS) and near-infrared (NIR) domains are significantly

saturated and will drastically affect the possibility to identify different snow facies in the

accumulation zone. The spectral signatures necessary to perform a supervised

86

classification were thus selected from the TM images, which provided a wider reflectance

range, and less band saturation, and thus preserved more spectral information to identify

snow and ice facies.

Figure 8.1. Maximum in-scene reflectance registered by TM and ETM+ sensors in the five subscenes.

0%

20%

40%

60%

80%

100%

120%

TM1 TM2 TM3 TM4 TM5 TM7

Band

Ref

lect

ace

19871989199720002001

Band TM 1987 TM 1989 TM 1997 ETM+ 2000 ETM+ 20011 34.87% 35.11% 36.68% 44.51% 41.77%2 80.49% 79.95% 83.53% 51.77% 48.85%3 66.17% 65.29% 68.39% 48.67% 45.93%4 101.38% 100.51% 105.19% 75.70% 71.91%5 64.67% 63.71% 66.84% 67.34% 65.67%7 70.57% 72.29% 70.69% 61.73% 58.26%

Table 8.1. Maximum in-scene reflectance registered by the TM and ETM+ sensors for the five subscenes. Note: Reflectance values greater than 100% are possible due to path radiance and background effect illumination

87

0%

5%

10%

15%

20%

25%

30%

35%

40%

45%

50%

TM1 TM2 TM3 TM4 TM5 TM7Band

Perc

ent o

f sat

urat

ed p

ixel

s

TM 1987TM 1989TM 1997ETM+ 2000ETM+ 2001

Figure 8.2. Percent saturated pixels for each band of the subscenes.

Band TM 1987 TM 1989 TM 1997 ETM+ 2000 ETM+ 20011 34.21% 43.59% 38.07% 29.47% 38.20%2 5.36% 6.98% 4.16% 25.35% 33.25%3 14.59% 17.70% 15.58% 28.53% 36.96%4 0.73% 0.70% 0.33% 8.44% 11.06%5 0.02% 0.02% 0.01% 0.01% 0.01%7 0.00% 0.00% 0.00% 0.00% 0.00%

Table 8.2. Percent saturated pixels for each band of the subscenes.

88

Spectral Endmembers Definition

Classification of satellite imagery is the process of remotely identifying target

types of ground covers; the information preserved in the pixels of multispectral imagery

(spectral signatures) is used to match spectral classes (groups of pixels with similar

spectral signature) to information classes (categories of land cover under investigations)

(Mather 1999). The purpose of this research was to identify and classify a set of

landcovers that may be used to describe the dynamics of glaciers.

Hall et al. (1987) and Hall et al. (1988) performed classifications of Landsat TM

scenes of the Grossglockner Glacier (Austria) using three information classes (Zone I, II,

III) to study environmental changes in alpine glaciers. For the accumulation zone Hall et

al. (1987) recognized a fresh snow zone (Zone III) and a wet snow zone (Zone II); the

ablation zone was classified as glacier ice (Zone I). Similarly, Rosenthal and Dozier

(1996) identified two major accumulation facies for the Sierra Nevada (USA) and

described them as winter snow (corresponding to fresh snow) and spring snow (a wet

aged snow facies). Klein and Isacks (1999) adopted a similar classification scheme to

analyze changes of the Zongo glacier (Cordillera Real, Bolivia), and argued that the

spectral classes associated with the glacier facies of Zongo Glacier may be exported to

perform a regional scale classification. Klein and Isacks (1999) suggested that it was

possible to apply locally developed spectral signatures at a regional scale. Extrapolation

of such spectral signatures is the basis of the study of this research.

Five information classes, based on a comparative reconstruction of the Klein and

Isacks (1999) study of Zongo Glacier were developed to classify the snow and ice cover

of the Cordillera Real. The spectral classes were based on a set of in-scene training sites

89

matching the glacier facies identified by Klein and Isacks (1999). Figure 8.3 shows the

three glacier facies Hall et al. (1987) identified on Zongo Glacier after Klein and Isacks

(1999). Figure 8.4 is a sample of some of the training sites over Zongo Glacier from the

1987 Landsat TM Image.

To increase the classifier algorithms’ performance in discriminating the glacier

facies, three more information classes were added (Klein and Isacks 1999): vegetation,

rock-soil, and shadows. The shadow class is necessary to obtain correct spectral un-

mixing in shaded areas, while vegetation and rock-soil are necessary to avoid

misclassification of these facies when present nearby or within glaciated covers (Klein

and Isacks 1999). Figure 8.5 shows the final spectral signatures applied to perform snow

Zone IZone II

Zone III

Figure 8.3. Accumulation (Zone III and Zone II) and ablation (Zone I) zones of Zongo Glacier as identified by Klein and Isacks (1999).

Figure 8.4. Sample of training sites used to obtain the spectral signatures for the information classes.

90

and ice cover classification on a regional (Cordillera Real) and temporal (multiple scene)

scale.

Binary classification

The simplest approach to identify snow and ice cover is based on binary

classifiers. The two resulting information classes isolate snow and ice covers from all

other materials. This methodology is useful as a first order classification to produce

snow-ice masks. Dozier (1989) proposed a technique based on the spectral characteristics

of snow and ice to develop a snow/ice mask from Landsat TM imagery. Klein and Isacks

Figure 8.5. Spectral signatures obtained from the training sites on Zongo Glacier.

Accumulation - Zone III

Accumulation - Zone II

Ablation - Zone I

Rock/Soil

Vegetation

Shadows

91

(1999) applied this technique to remove unwanted landcovers from Zongo Glacier. The

procedure is also particularly effective in identifying and removing clouds from glaciated

regions.

Dozier (1989) observed that snow and ice can be isolated from other land covers

on the base of the three following conditions:

Reflectance of TM band 1 greater than 15% (or a selected threshold between 15%

and 20%); this threshold isolates surfaces with high brightness, including non-snow

surfaces that are removed by the following two rules;

Reflectance of TM band 5 lower than 25% (or a selected threshold between 20%

and 25%); this rule distinguishes between clouds and snow, being that clouds are highly

reflective in the MIR domain, and snow is highly absorbing.

Normalized Difference Snow Index (NDSI) calculated from reflectance values

must be greater than 0.4; this procedure filters out bright soils, rocks and clouds.

The selection of the thresholds can be obtained by analysis of the atmospheric

conditions and spectral response of the sensors at the acquisition time, or by in-scene

observations of reflectance values (Dozier 1999). This binary classification was

effectively applied to the five Landsat scenes to estimate regional variations in snow and

ice landcover over time. Figure 8.6 shows a sample of the resulting binary classification

for Zongo Glacier. The binary classification was applied to the five project images but

the 2001 Landsat ETM+ was excluded from the figure for reasons explained in the next

chapter. The results are also discussed in the next chapter, but it is already evident from

figure 8.6 that from 1987 to 1997 there is a visible retreat of the glacier terminus, and a

loss of snow cover along the edges of the glaciated areas.

92

Supervised Classification: Minimum Distance

The minimum distance classifier was selected from among the possible

supervised classification techniques. One of the driving reasons for the selection of this

method was its computational power, being the fastest decision rule to calculate (ERDAS

1997). Another reason for selecting this method was the simple implementation of the

algorithm that does not require detailed statistical information about the endmembers; the

minimum distance classifier calculates the Euclidean distance of an unclassified pixel

from the spectral vectors of endmembers; endmembers can be obtained from spectral

libraries as predefined spectral signatures, and fed to the minimum distance algorithm

Figure 8.6. Binary classification of Zongo Glacier using the snow/ice mask after Dozier (1989) applied to four Landsat scenes.

Zongo Glacier terminus

93

without involving complex in-scene statistics (such is the case with the maximum

likelihood classifier) (RSI 2001). Since the endmembers from this project had to be

exported from one scene to the others, the minimum distance classifiers offered the best

and simplest classification solution. Figure 8.7 and figure 8.8 show samples of the

resulting minimum distance classification of Zongo Glacier.

The ETM+ band saturation is clearly visible in figure 8.8. The saturation

obliterates information for bright reflectors such as fresh snow; as a result, the classifiers

cannot detect the difference between Zone III and Zone II and returns Zone III as a global

class for the accumulation endmembers.

Ablation – Zone I

Accumulation – Zone III

Vegetation

Accumulation – Zone II Rock and Soil

Ablation – Zone IAblation – Zone I

Accumulation – Zone IIIAccumulation – Zone III

VegetationVegetation

Accumulation – Zone IIAccumulation – Zone II Rock and SoilRock and Soil

Figure 8.7. Minimum distance supervised classification of Zongo Glacier from Landsat TM 1987.

Ablation – Zone I

Accumulation – Zone III

Vegetation

Accumulation – Zone II Rock and Soil

Figure 8.8. Minimum distance supervised classification of Zongo Glacier from Landsat ETM+ 2000.

94

Spectral unmixing

Klein and Isacks (1999) found that the spectral un-mixing technique performed

well in defining the position of the transient snowline (or equilibrium line). Interpretation

of the resulting classification is more complex than that of supervised classifiers due to

the fact that the resulting information layers represent transitional fractions of

endmembers rather than absolute hard class membership. Each classified pixel, thus,

needs to be interpreted as a function of its endmember fractions. It is important to observe

that spectral unmixing is highly effective in obtaining information in shaded area and

transitional zones where hard classifiers, such as minimum distance, have to output an

absolute decision about pixel membership. Through spectral un-mixing, for example, it is

possible to obtain the fractional amount of snow and ice present in proximity to (or

within) vegetation, rock, and soil landcovers, or in shaded areas. Rosenthal and Dozier

(1996) successfully applied the spectral un-mixing technique to obtain detailed

information relative to snow fraction in the Sierra Nevada (USA).

Figure 8.9 and figure 8.10 represent the results of a spectral un-mixing

classification performed on the 1987 Landsat TM. The results of the classification match

the results obtained by Klein and Isacks (1999) and confirm that the spectral signatures of

the endmembers are correctly calibrated to describe the information classes for the

accumulation and ablation zones. From the spatial transect (figure 8.10) it can be

observed that the transient snowline is estimated at the intersection of the accumulation

and ablation endmember fraction curves. In spatial terms, this spectral intersection

represents the location at which there is a transition from snow to ice cover, and thus

from accumulation to ablation.

95

Spectral un-mixing failed to give meaningful results when the endmembers

obtained from the Landsat TM scenes were exported to the Landsat ETM+ scenes due to

the distortion induced by the band saturation on the spectral signatures. This result

enforced the idea that ETM+ spectrometers in high-gain mode are not fit for snow and ice

facies classification based on spectral information.

A

B

A

B

Ablation endmember Accumulation endmember

Figure 8.9. Accumulation and ablation endmembers of Zongo Glacier, and AB transect (figure 8.10). The fraction is represented in grey scale with white as the highest and black as lowest .

Transient snowline

96

Figure 8.10. Accumulation and ablation endmember fractions along the AB transect (figure 8.9).

Accumulation Ablation

Transient snowline

A BSpatial transect

97

Chapter 9

Results

Zonal statistics provide a meaningful measure of the magnitude and spatial

pattern of the environmental changes of the glaciers of the Cordillera Real. Zonal

functions are used to calculate raster datasets where each pixel value depends on the

value of the source cell (information layer) and the association that the source cell has

within a cartographic zone (zone layer). In particular, zonal functions can be used to

calculate descriptive statistics of classified satellite imagery based on predefined

cartographic areas (ESRI 2003). For this project, the binary and the supervised minimum

distance classification outputs provided the information layers, and Jordan’s 1984 glacier

inventory (Jordan 1991) the zone layer. Each zone was defined by the glacier ID found in

Jordan’s 1984 inventory. ESRI ArcGIS 8.3 Spatial Analyst offered a powerful tool to

obtain zonal statistics of the classification results.

Binary Classification: Zonal Statistics

Zonal statistics applied to the snow and ice binary classification provided the

number of snow and ice covered pixels within each glacier zone. From the summary of

the statistics (table 9.1 and figure 9.1) it is clear that the amount of snow and ice cover in

98

the Cordillera Real was at minimal levels in 2000, with greater amounts of snow and ice

in 1987, 1997, 1989, 2001 (figure 9.1).

Relative snow and ice cover differences were calculated for the year pairs 1987-

1989, 1989-1997, 1997-2000, and 2000-2001 in order to quantify total and zonal surface

gains and losses. Table 9.2 illustrates the changes in glacial ice and snow cover for the

respective time periods by number and percent. It can be observed that between 1989 and

Figure 9.1. Total snow/ice cover of the Cordillera Real measured by binary classification.

1987

1989

1997

2000

2001

210

220

230

240

250

260

270

1986 1988 1990 1992 1994 1996 1998 2000 2002

Year

Tota

l sno

w/ic

e co

ver (

sq. k

m)

Table 9.1. Snow/ice area measured from the zonal statistics of the binary classification.

257,839,828220,392,666 236,320,889250,747,261233,632,341Total Snow/ice cover (m2)

317,439271,336 290,946308,707287,636Total Snow/ice cover (pixels) 20012000 199719891987Reference year

99

1997, and between 1997 and 2000, 55% and 63% of the glaciers of the Cordillera Real

lost snow and ice surface respectively. From 1987 to 1989, and from 2000 to 2001, the

number of glaciers that gained snow and ice surface was estimated to be 72% and 84%

respectively. The maximum surface loss was registered from 1997 to 2000 (15,928,223

m2), followed by the 1989-1997 negative balance (14,426,372 m2) for a total surface

reduction of 30,354,595 m2 from 1989 to 2000. Between 1987 and 1989, and between

2000 and 2001, the glaciers registered a net gain in area. The inter-scene percent of snow

and ice cover lost or gained is shown in table 9.3.

Reference year 1987 1989 1997 2000 2001Total Snow/ice cover (pixels) 287,636 308,707 290,946 271,336 317,439 Total Snow/ice cover (sqm) 233,632,341 250,747,261 236,320,889 220,392,666 257,839,828

1987-1989 1989-1997 1997-2000 2000-20001

Net areal balance (pixels) 21071 -17761 -19610 46103Net areal balance (sqm) 17,114,920 -14,426,372 -15,928,223 37,447,162

Percent change of snow/icecover from reference year 7.3% -5.8% -6.7% 17.0%

Table 9.3. Inter-scene snow/ice cover percent changes.

1987-1989 1989-1997 1997-2000 2000-20001Number of Glaciers w ith Snow/ice loss 135 495 563 54Percent of Glaciers w ith Snow/ice loss 15.1% 55.4% 63.0% 6.0%

Number of Glaciers w ith Snow/ice gain 640 300 215 748Percent of Glaciers w ith Snow/ice gain 71.7% 33.6% 24.1% 83.8%

Number of Glaciers w ith Snow/ice unchanged 118 98 115 91Percent of Glaciers w ith Snow/ice unchanged 13.2% 11.0% 12.9% 10.2%

Total area losses (pixels) -1653 -24208 -25987 -787Total area losses (sqm) -1,342,649 -19,662,948 -21,107,941 -639,241

Total area gains (pixels) 22724 6447 6377 46890Total area gains (sqm) 18,457,569 5,236,576 5,179,718 38,086,403

Net areal balance (pixels) 21071 -17761 -19610 46103Net areal balance (sqm) 17,114,920 -14,426,372 -15,928,223 37,447,162

Table 9.2. Inter-scene zonal statistics illustrating gain, losses and total net balance.

100

The losses between 1997 and 2000 corresponded to 6.7% of the cover registered

in 1997. Similarly, the losses between 1989 and 1997 were 5.8% of the 1989 glaciated

surface.

Minimum Distance: Zonal Statistics

A second analysis applied zonal statistics to the results of the supervised

minimum distance classification. The global trend of changes highlighted by the zonal

statistics applied to the supervised minimum distance classification was similar to the

binary classification results (table 9.4 and figure 9.2). Year 2000 presented the lowest

snow and ice cover extent followed by 1987, 1997, 1989 and 2001.From an inter-scene

point of view (table 9.5 and table 9.6), zonal statistics showed that the largest amount of

snow and ice cover was lost between 1989 and 1997, when 56% of the glaciers lost a

total of 14,471,046 m2 of snow and ice surface, corresponding to 5.6% of the 1989

surface extent; between 1997 and 2000, 57% of the glaciers showed a reduction of the

snow and ice surface for a total of 5,263,380 m2, and corresponding to 2.2% of year

1997. The periods from 1987 to 1989 and from 2000 to 2001 had snow and ice surface

gain 6.6% of 1987 and 13.4% of 2000 respectively.

Reference year 1987 1989 1997 2000 2001Total Snow/ice cover (pixels) 296,836 316,353 298,537 282,174 321,701Total Snow/ice cover (sqm) 241,105,041 256,957,724 242,486,678 229,195,832 261,301,637

Table 9.4. Snow/ice area measured from the zonal statistics of the minimum distance classification

101

Figure 9.2. Total snow/ice cover of the Cordillera Real measured by minimum distance classifications.

2000

2001

1997

1989

1987

210

220

230

240

250

260

270

1986 1988 1990 1992 1994 1996 1998 2000 2002

Year

Tota

l sno

w/ic

e co

ver (

sq. k

m)

102

Detailed change detection analysis (table 9.7a through 9.7d) provided information

about the interclass variations; the columns present the classes in the initial state, while

the rows represent the class distribution at final state. Class changes reflect environmental

dynamics and help to identify change anomalies and misclassifications. For example, a

summary of the net class differences between the scenes suggested that the surface gains

registered between 1987-1989 and 2000-2001, counterbalanced by a loss of rock-soil and

1987-1989 1989-1997 1997-2000 2000-20001Number of Glaciers w ith Snow/ice loss 129 500 577 60Percent of Glaciers w ith Snow/ice loss 14.4% 56.0% 64.6% 6.7%

Number of Glaciers w ith Snow/ice gain 646 282 190 735Percent of Glaciers w ith Snow/ice gain 72.3% 31.6% 21.3% 82.3%

Number of Glaciers w ith Snow/ice unchanged 118 111 123 96Percent of Glaciers w ith Snow/ice unchanged 13.2% 12.4% 13.8% 10.8%

Total area losses (pixels) -1304 -23895 -21887 -784Total area losses (sqm) -1,059,174 -19,408,714 -17,777,716 -636,804

Total area gains (pixels) 20821 6079 5524 40311Total area gains (sqm) 16,911,857 4,937,668 4,486,869 32,742,610

Net areal balance (pixels) 19517 -17816 -16363 39527Net areal balance (sqm) 15,852,683 -14,471,046 -13,290,847 32,105,806

Table 9.5. Inter-scene zonal statistics from minimum distance classification illustrating gain, losses and total net balance.

Reference year 1987 1989 1997 2000 2001Total Snow/ice cover (pixels) 296,836 316,353 298,537 282,174 321,701Total Snow/ice cover (sqm) 241,105,041 256,957,724 242,486,678 229,195,832 261,301,637

1987-1989 1989-1997 1997-2000 2000-20001Net areal balance (pixels) 19,517 -17,816 -16,363 39,527Net areal balance (sqm) 15,852,683 -14,471,046 -13,290,847 32,105,806

Percent change of snow/icecover from reference year 6.6% -5.6% -5.5% 14.0%

Table 9.6. Inter-scene snow/ice cover percent changes determined from minimum distanceclassification.

103

shadows classes, might not represent glacial activity but extended snow falls events (table

9.8 and figure 9.3).

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 89,329 1,466 23,063 16,534 23,215 153,607Vegetation 796 151 243 83 1,088 2,361Rock-Soil 11,267 563 104,735 1,336 12,474 130,375Accumulation 39,005 168 7,578 152,840 1,828 201,419Shadow 10,244 1,153 8,927 672 129,738 150,734

Class Total 150,641 3,501 144,546 171,465 168,343Class Changes 61,312 3,350 39,811 18,625 38,605Image Difference 2,966 -1,140 -14,171 29,954 -17,609

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 59.3% 41.9% 16.0% 9.6% 13.8%Vegetation 0.5% 4.3% 0.2% 0.0% 0.6%Rock-Soil 7.5% 16.1% 72.5% 0.8% 7.4%Accumulation 25.9% 4.8% 5.2% 89.1% 1.1%Shadow 6.8% 32.9% 6.2% 0.4% 77.1%

Class Total 100.0% 100.0% 100.0% 100.0% 100.0%Class Changes 40.7% 95.7% 27.5% 10.9% 22.9%Image Difference 2.0% -32.6% -9.8% 17.5% -10.5%

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 72,557,480 1,190,759 18,732,922 13,429,742 18,856,384 124,767,286Vegetation 646,551 122,650 197,377 67,417 883,728 1,917,722Rock-Soil 9,151,621 457,297 85,071,004 1,085,166 10,132,007 105,897,094Accumulation 31,681,811 136,458 6,155,231 124,144,290 1,484,793 163,602,583Shadow 8,320,689 936,524 7,250,956 545,832 105,379,691 122,433,692

Class Total 122,358,152 2,843,687 117,407,489 139,272,446 136,736,602Class Changes 49,800,672 2,721,038 32,336,485 15,128,156 31,356,911Image Difference 2,409,134 -925,965 -11,510,395 24,330,137 -14,302,910

Pixel Counts (1987-1989)

Percentages (1987-1989)

Area (Square Meters) (1987-1989)

Table 9.7a. Change detection statistics from 1987-1989

104

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 89,207 966 16,754 29,656 13,230 149,813Vegetation 1,934 106 1,216 114 1,864 5,234Rock-Soil 14,698 201 95,110 1,157 5,119 116,285Accumulation 21,083 135 3,449 168,192 1,656 194,515Shadow 26,685 953 13,846 2,300 128,865 172,649

Class Total 153,607 2,361 130,375 201,419 150,734Class Changes 64,400 2,255 35,265 33,227 21,869Image Difference -3,794 2,873 -14,090 -6,904 21,915

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 58.1% 40.9% 12.9% 14.7% 8.8%Vegetation 1.3% 4.5% 0.9% 0.1% 1.2%Rock-Soil 9.6% 8.5% 73.0% 0.6% 3.4%Accumulation 13.7% 5.7% 2.6% 83.5% 1.1%Shadow 17.4% 40.4% 10.6% 1.1% 85.5%

Class Total 100.0% 100.0% 100.0% 100.0% 100.0%Class Changes 41.9% 95.5% 27.0% 16.5% 14.5%Image Difference -2.5% 121.7% -10.8% -3.4% 14.5%

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 72,458,386 784,634 13,608,437 24,088,086 10,746,068 121,685,609Vegetation 1,570,892 86,099 987,696 92,597 1,514,034 4,251,317Rock-Soil 11,938,451 163,262 77,253,098 939,773 4,157,908 94,452,491Accumulation 17,124,667 109,654 2,801,450 136,613,952 1,345,086 157,994,809Shadow 21,674,891 774,074 11,246,414 1,868,175 104,670,596 140,234,150

Class Total 124,767,286 1,917,722 105,897,094 163,602,583 122,433,692Class Changes 52,308,900 1,831,624 28,643,996 26,988,631 17,763,095Image Difference -3,081,677 2,333,594 -11,444,603 -5,607,774 17,800,459

Pixel Counts (1989-1997)

Percentages (1989-1997)

Area (Square Meters) (1989-1997)

Table 9.7b. Change detection statistics from 1989-1997

105

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 75,309 1,308 11,914 28,875 19,005 136,411Vegetation 1,302 131 533 433 806 3,205Rock-Soil 23,027 1,008 89,608 8,574 13,123 135,340Accumulation 27,022 148 1,029 152,669 1,816 182,684Shadow 23,153 2,639 13,201 3,964 137,899 180,856

Class Total 149,813 5,234 116,285 194,515 172,649Class Changes 74,504 5,103 26,677 41,846 34,750Image Difference -13,402 -2,029 19,055 -11,831 8,207

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 50.3% 25.0% 10.2% 14.8% 11.0%Vegetation 0.9% 2.5% 0.5% 0.2% 0.5%Rock-Soil 15.4% 19.3% 77.1% 4.4% 7.6%Accumulation 18.0% 2.8% 0.9% 78.5% 1.1%Shadow 15.5% 50.4% 11.4% 2.0% 79.9%

Class Total 100.0% 100.0% 100.0% 100.0% 100.0%Class Changes 49.7% 97.5% 22.9% 21.5% 20.1%Image Difference -8.9% -38.8% 16.4% -6.1% 4.8%

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 61,169,735 1,062,423 9,677,147 23,453,719 15,436,811 110,799,835Vegetation 1,057,550 106,405 432,929 351,704 654,674 2,603,261Rock-Soil 18,703,681 818,748 72,784,098 6,964,232 10,659,157 109,929,915Accumulation 21,948,620 120,213 835,805 124,005,395 1,475,046 148,385,079Shadow 18,806,024 2,143,528 10,722,512 3,219,759 112,008,463 146,900,286

Class Total 121,685,609 4,251,317 94,452,491 157,994,809 140,234,150Class Changes 60,515,874 4,144,912 21,668,393 33,989,414 28,225,688Image Difference -10,885,775 -1,648,055 15,477,424 -9,609,730 6,666,136

Pixel Counts

Percentages

Area (Square Meters)

Table 9.7c. Change detection statistics from 1997-2000

106

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 106,199 1,506 30,755 18,830 41,507 198,797Vegetation 584 81 739 54 1,600 3,058Rock-Soil 5,341 428 83,925 231 13,076 103,001Accumulation 26,444 176 5,258 140,409 2,060 174,347Shadow 7,230 487 8,422 292 142,862 159,293

Class Total 145,798 2,678 129,099 159,816 201,105Class Changes 39,599 2,597 45,174 19,407 58,243Image Difference 52,999 380 -26,098 14,531 -41,812

Ablation Vegetation Rock-Soil Accumulation ShadowAblation 72.8% 56.2% 23.8% 11.8% 20.6%Vegetation 0.4% 3.0% 0.6% 0.0% 0.8%Rock-Soil 3.7% 16.0% 65.0% 0.1% 6.5%Accumulation 18.1% 6.6% 4.1% 87.9% 1.0%Shadow 5.0% 18.2% 6.5% 0.2% 71.0%

Class Total 100.0% 100.0% 100.0% 100.0% 100.0%Class Changes 27.2% 97.0% 35.0% 12.1% 29.0%Image Difference 36.4% 14.2% -20.2% 9.1% -20.8%

Ablation Vegetation Rock-Soil Accumulation Shadow Class TotalAblation 86,260,138 1,223,249 24,980,749 15,294,668 33,714,061 161,472,863Vegetation 474,354 65,792 600,253 43,862 1,299,600 2,483,861Rock-Soil 4,338,227 347,643 68,168,081 187,630 10,620,981 83,662,562Accumulation 21,479,139 142,956 4,270,811 114,047,210 1,673,235 141,613,351Shadow 5,872,568 395,566 6,840,770 237,177 116,039,660 129,385,739

Class Total 118,424,426 2,175,206 104,860,663 129,810,546 163,347,536Class Changes 32,164,288 2,109,413 36,692,582 15,763,336 47,307,877Image Difference 43,048,438 308,655 -21,198,101 11,802,805 -33,961,797

Pixel Counts (2000-2001)

Percentages (2000-2001)

Area (Square Meters) (2000-2001)

Table 9.7d. Change detection statistics from 2000-2001

107

Figure 9.3. Surface changes in km2 for each scenes pair obtained from the minimum distance classification.

-40

-30

-20

-10

0

10

20

30

40

50

1987-1989 1989-1997 1997-2000 2000-2001

Years

Surf

ace

chan

ges

(sq.

Km

)

AblationAccumulationShadowRock-SoilVegetation

Ablation Accumulation Shadow Rock-Soil Vegetation1987-1989 2,409,134 24,330,137 -14,302,910 -11,510,395 -925,9651989-1997 -3,081,677 -5,607,774 17,800,459 -11,444,603 2,333,5941997-2000 -3,261,184 -28,184,263 23,113,386 10,408,172 -2,076,1112000-2001 43,048,438 11,802,805 -33,961,797 -21,198,101 308,655

Table 9.8. Detail of class areal changes (sq. km)

108

The interclass changes of the ablation zones follow two main patterns (figure 9.4).

The first pattern is common for the pairs 1989-1997 and 1997-2000, where the general

loss of snow and ice cover correspond in the loss of ablation zone and a gain of rock-soil,

accumulation and shadows surface (which includes few misclassified water pixels); The

second pattern, common to the pairs 1987-1989 and 2000-2001, shows a general gain in

snow and ice surface, and is characterized by a net loss of ablation class in favor of

accumulation class.

Observing the variation of the rock-soil class (figure 9.3 and figure 9.4) it could

be argued that the snow and ice cover gains in the 1987-1989 and 2000-2001 pairs are not

the result of glacial processes, but the product of widespread episodes of late season snow

fall that simulate an increase of accumulation area and a reduction of coverage of all the

Figure 9.4. Percent change from ablation class to other classes.

0.0%

10.0%

20.0%

30.0%

40.0%

50.0%

60.0%

70.0%

80.0%

90.0%

100.0%

Ablation Vegetation Rock-Soil Accumulation Shadow

Class

Perc

ent C

hang

e

Ablation 87-89

Ablation 89-97

Ablation 97-00

Ablation 00-01

109

other classes. This phenomenon is supported by field and remote sensed observation that

will be discussed in chapter 10.

Binary vs. Minimum Distance Classifier

The trend vectors of the two classifications methods are parallel in direction and

different in magnitude (figure 9.5); this result is significant because it is the outcome of

classifications based on two very different methodologies: the binary classification results

from constraint imposed on spectral reflectance, while the minimum distance classifier

operates on pattern recognition of spectral signatures.

The trend directions are parallel, so it can be argued that the surface changes

measured by the two classifications are, in fact, due to physical changes detected from the

Figure 9.5. Comparison of the total snow/ice cover of the Cordillera Real measured by binary and minimum distance classifications.

2001

2000

1997

1987

1989

210

220

230

240

250

260

270

1986 1988 1990 1992 1994 1996 1998 2000 2002

Year

Tota

l sno

w/ic

e co

ver (

sq. k

m)

Binary

Minimum Distance

110

five scenes. The difference in magnitude, in this case, is less relevant because the

direction of the change is the more significant variable for understanding environmental

dynamics. The difference in the magnitude can be attributed to the sensitivity of the

methodologies applied to detect the change. The binary classification was found to be

highly selective, because its rules operate as a bandpass filter with well defined cut-off

thresholds; this was particularly noticeable in shaded areas, where the binary classifier

underestimates snow and ice cover. On the other hand, the minimum distance classifier is

more sensitive and less selective, allowing the recognition of snow and ice cover in

shaded and mixed pixel areas, resulting in a larger number of pixels classified as snow

and ice cover. By varying the reflectance thresholds, the binary classification could be

tuned to match the minimum distance results.

Regional vs. Local Observations

Zongo Glacier (68.14°W, 16.27°S, approximately 3 km2), Chacaltaya Glacier

(68.12°W, 16.35°S, approximately 0.5 km2), and the unnamed glacier feeding Laguna

Glaciar (68.55°W, 15.83°S, approximately 4.3 km2) (from here referred as Laguna

Glaciar for simplicity) represent three retreating glaciers of the Cordillera Real that have

been observed under field monitoring since the 1980s. Mass balance records for Zongo

and Chacaltaya cover the period from 1991 to 2002 (Francou et al 2003). During the past

6 years, Laguna Glacial retreat has been observed in the field, but not measured, by the

Geography Department of Appalachian State University, NC.

Laguna Glaciar (figure 9.6) and Zongo Glacier (figure 9.7) present a surface

variation pattern that closely reflects the regional trend of the Cordillera Real. Chacaltaya

111

Glacier (figure 9.8), while showing a general retreat pattern, presents some anomalies.

The apparent surface increase of Chacaltaya Glacier registered from 1987 to 1997, and

the sudden drop in 2000 can be explained by the fact that the glacier surface is

considerably smaller than the other two glaciers, rendering the estimate of the real glacier

surface particularly sensitive to sudden environmental changes such as local snowfall

events. However, similar to the regional trend, the three glaciers presented a starting low

snow and ice cover surface in the year 1987, followed by an increase between 1987 and

1989, and a significant surface loss during the period 1997-2000. In the year 2001, a

significant increase of the snow and ice cover was recorded for the three glaciers

confirming the observations conducted at a regional scale.

3

3.05

3.1

3.15

3.2

3.25

3.3

3.35

3.4

3.45

3.5

1986 1988 1990 1992 1994 1996 1998 2000 2002

Year

Tota

l sno

w/ic

e co

ver (

sq. k

m)

Binary maskMinimum Distance

Figure 9.6. Laguna Glaciar: Comparison of the total snow/ice measured by binary and minimum distance classifications.

112

Figure 9.7. Zongo Glacier: Comparison of the total snow/ice measured by binary and minimum distance classifications.

2.3

2.35

2.4

2.45

2.5

2.55

2.6

2.65

1986 1988 1990 1992 1994 1996 1998 2000 2002

Year

Tota

l sno

w/ic

e co

ver (

sq. k

m)

Binary maskMinimum Distance

113

0.2

0.22

0.24

0.26

0.28

0.3

0.32

0.34

1986 1988 1990 1992 1994 1996 1998 2000 2002

Year

Tota

l sno

w/ic

e co

ver (

sq. k

m)

Binary maskMinimum Distance

Figure 9.8. Chacaltaya Glacier: Comparison of the total snow/ice measured by binary and minimum distance classifications.

114

Chapter 10

Discussion and Conclusion

Discussion

The results of the snow and ice classification presented in chapter 9 showed the

presence of two predominant trends in the dynamic evolution of the glaciers of the

Cordillera Real recorded by the five Landsat scenes. From 1987 to 1989 and from 2000

to 2001 the snow and ice cover seems to increase, while during the period 1989 to 1997

and from 1997 to 2000, a progressive reduction of the snow and ice cover is registered.

From the analysis of change detection statistics, it can be argued that the snow and ice

cover variations registered for the pairs 1987-1989 and 2000-2001 are, in fact, not the

result of glacial processes but extended short-term snow events. Fresh snow presents

similar spectral characteristics to the accumulation classes describes as Zone II and Zone

III. As a result, snow falls are classified as accumulation zone and the global snow and

ice cover for that scene is larger. Also, fresh snow that reaches elevations below the

glacier terminus obliterates the spectral signatures of other classes by covering ice, rock-

and soil and vegetation. These effects are visible in the global change detection statistics

for the pair 1987-1989 and 2000-2001, where the general increase of snow and ice cover

is counterbalanced by a reduction of the ablation and rock-soil classes. On the other hand,

retreat of glacier surfaces should be reflected as a general reduction of the snow and ice

115

cover balanced by a relative increase of non-glacier classes such as rock-soil, as visible

for the pairs 1989-1997 and 1997-2000.

These important observations obtained from in-scene analysis are confirmed by

independent satellite data and field measurements. An ASTER satellite scene acquired

June 29th, 2001 confirmed that a conspicuous snowfall event took place before Landsat

ETM+ acquired the July 31, 2001 scene of the Cordillera Real (figure 10.1). The

extended snowfall maintained a substantial cover at elevations lower than the glacier

termini, resulting in an apparent increase of the glacial surface. The snow event is also

reported in field measurements by Francou et al. (2003) who registered a positive mass

balance for Chacaltaya glacier in the year 2000. A similar snowfall event has been

assumed to describe the snow and ice cover increase for the pair 1987-1989. It was

observed that the pattern of changes for the ablation class detected for the pair 1987-1989

is consistently similar to that the pair 2000-2001 (chapter 9).

Figure 10.1. ASTER scene (left) acquired on 06/29/2001 and Landsat ETM+ scene (right) acquired on 07/31/2001 representing the central portion of the Cordillera Real after an extended snow fall event (USGS Global Visualization Viewer, http://glovis.usgs.gov).

116

It may thus be argued that the 1989 Landsat TM and 2001 Landsat EMT+ scenes

are not suitable to identify glacial retreat processes for the study area. The 1989 snowfall

recorded by the Landsat TM image appears to be more localized and the image could be

used for local studies of unaffected areas, while the 2001 event in the Landsat ETM+

presented a widespread regional disturbance. Consequently, within the five images

dataset, the 1987 and the 2000 scenes represent the closest portrait of undisturbed snow

and ice cover extend for the Cordillera Real, and will represent the reference to estimate

the total snow and ice cover change during the period of time.

Table 10.1 illustrates the snow and ice areal losses within the Cordillera Real

considering 1987 as the initial state and the year 2000 as the final; it was possible to

estimate a total snow and ice cover loss of 5.7% (binary classification) and 4.9%

(minimum distance) during this 13 year interval. The general trend is in agreement with

observations reported for example, by Thompson and Davis (1998), Kaser (1999),

Haeberli (2003), and Francou et al (2003). The observations conducted at local scale on

Zongo, Chacaltaya and Laguna Glaciar reflect the general regional trend, but also

emphasize the net dynamic contrast between larger glaciers (surface greater than 0.5 km2)

and smaller glaciers (less than 0.5 km2).

Classification TypeReference year 1987 2000 1987 2000Total Snow/ice cover (pixels) 287,636 271,336 296,836 282,174Total Snow/ice cover (sqm) 233,632,341 220,392,666 241,105,041 229,195,832

1987-2000 1987-2000Net areal balance (pixels) -16,300 -14,662Net areal balance (sqm) -13,239,675 -11,909,210

Percent change of snow/ice cover -5.7% -4.9%

Table 10.1. Inter-scene snow/ice cover percent losses.

Binary Minimum Distance

117

Francou et al. (2003) report that 80% percent of the glaciers of the Cordillera Real

have a surface area less than 0.5 km2 (79% in the subset of the Cordillera Real used for

this research thesis). Due to the ongoing climatic and environmental changes, these

glaciers are close to their critical mass, beyond which, the retreating process becomes

irreversible and the solid precipitation is not able to support the generation of new ice

mass (Kaser 1999, Haeberli et al. 1999, Thompson 2000, Francou et al. 2003).

Chacaltaya Glacier is a representative example of a glacier with surface area near 0.5

km2. It can be observed that the surface loss for Chacaltaya Glacier was estimated

between 11.9% and 18.6% (table 10.2), values significantly higher than the regional

estimates. The snow and ice cover loss measurements performed by Landsat imagery for

Chacaltaya is in agreement with field observations that predict the extinction of the

glacier within the next 30 years (Thompson and Davis 1998, Thompson 2000, Francou et

al. 2003).

The higher sensitivity of small glaciers such as Chacaltaya is also supported by

statistical observations performed on a subset of glaciers of the Cordillera Real

characterized by surface areas between 0.25 km2 and 0.75 km2. Depending on the

classification methodology (binary or minimum distance), the zonal statistics applied to

the 122 glaciers subset (representing 14% of the total number) showed a total loss of

Binary Mask Minimum DistanceNet areal balance (pixels) -39 -63Net areal balance (sqm) -31,678 -51,172

Percent change of snow/icecover from reference year -11.9% -18.6%

Chacaltaya1987-2000

Table 10.2. Inter-scene snow and ice cover change estimates for Chacaltaya Glacier.

118

snow and ice cover of 9% to 10%. Also, of the 122 glaciers, 75% to 79% presented snow

and ice cover losses, with an average surface loss of 16%.

On the other hand, Zongo Glacier (table 10.3), Laguna Glaciar (table 10.4), and a

subset of 20 glaciers with area between 2 km2 and 5 km2 presented a surface loss close to

the regional estimate. The total surface loss was estimated between 2% to 3%, but 76% to

80% of the 20 glaciers showed an average surface loss of 4%. These results are in

agreement with field measurements performed by Francou et al. (2003) on Zongo

Glacier, and confirm that these glaciers, while retreating, could still recover ice mass in

favorable environmental and climatic conditions.

Technical, and Physical Limitations of the Results

Satellite imagery from the Landsat platforms TM and ETM+ have been used in

many projects to study single glaciers and measure snow and ice cover extent in alpine

environments. The relationship between snow and ice cover changes detected by satellite

Binary Mask Minimum DistanceNet areal balance (pixels) -215 -140Net areal balance (sqm) -174,634 -113,715

Percent change of snow/icecover from reference year -6.9% -4.4%

Zongo Glacier1987-2000

Table 10.3. Inter-scene snow and ice cover change estimates for Zongo Glacier.

Table 10.4. Inter-scene snow and ice cover change estimates for Laguna Glaciar.

Binary Mask Minimum DistanceNet areal balance (pixels) -79 -219Net areal balance (sqm) -64,168 -177,883

Percent change of snow/icecover from reference year -2.0% -5.5%

Laguna Glaciares1987-2000

Glacial

Binary Mask Minimum DistanceNet areal balance (pixels) -79 -219Net areal balance (sqm) -64,168 -177,883

Percent change of snow/icecover from reference year -2.0% -5.5%

Laguna Glaciares1987-2000

Glacial

119

platforms at a regional scale and the estimate of glacier mass balance is, however, subject

to complex interactions between numerous environmental variables (Haeberli 2003).

Remote sensing based estimates of mass balance changes provide the advantage of being

faster than regional scale field studies, a major advantage when considering accelerating

global warming processes that are shortening the life of a number of glaciated regions of

the world. In addition, many alpine glaciers are remote and difficult to access on foot.

Also, in Bolivia, difficult political events during the latter half of the 2003 have further

reduced access for field studies in the Cordillera Real. This research study has found a list

of issues that affect the interpretation of quantitative estimates of glacier mass balance at

the regional scale from snow and ice cover change detection. The issues identified were

of both a technical and a physical nature.

The first problem encountered in this research study was related to the technical

limitations of the Landsat TM and ETM+ spectrometers. Landsat platforms were

developed to study vegetation; thus, any attempt to implement Landsat imagery for other

purposes may go beyond the capability of the sensors. This research emphasized that

Landsat platforms present important spectral limitations that affect the accuracy of the

detection of snow and ice facies within the tropics. Landsat TM sensors saturate in the

visible and NIR domain over snow cover, but maintain sufficient spectral response to

allow the identification of major glacial facies such as fresh snow, firn, and glacial ice.

Landsat ETM+ performance is dramatically affected by LTAP defined gain rules; this

study showed that ETM+ spectrometers in high-gain mode failed to discriminate the

accumulation facies (fresh snow to firn) due to the high saturation in the visible and NIR

bands. As a result, the identification of snow and ice cover is inherently limited by

120

classification errors. Also, it was observed that snow and ice spectral libraries developed

from TM imagery perform poorly when applied to ETM+ scenes, rendering impossible

the development of glacial models based on physical measurements such as reflectance

beyond the available Landsat TM dataset.

Another limitation in the estimation of snow and ice cover to study glacier

dynamic is of a physical nature. Satellite scenes are single shot images that represent the

Earth within well defined time and space coordinates, and do not contain information to

describe dynamic processes. In the specific case of glaciers, it may be correctly argued

that the temporal resolution of Landsat satellites is higher than the rate of change of the

glaciers. Unfortunately, physical atmospheric processes, such as cloud cover, greatly

affect the availability of clear images, especially in alpine environments. Also, if seasonal

constraints are added to the selection of suitable imagery for glaciological application, the

number of possible images available in the Landsat dataset may be reduced from a few to

none in any given year. In the case of Landsat ETM+, acquisition plans such as LTAP

may further reduce availability of imagery for remote areas.

Another physical limitation is related to atmospheric events such as extended

snowfall events that hide the glaciated area with a uniform cover. In extreme cases, such

as for the 2001 Landsat image in this study, early winter snowfall obliterated the real

extent of accumulation and ablation zones and provided false results about the effective

snow and ice cover extent of the glaciers of the Cordillera Real. In less extreme cases,

such as for the 1989 scene (and partially for the 1997 scene), snowfall may affect only a

portion of the study area, or provide misleading detection of the equilibrium line of the

glaciers. Again, the sum of the physical limitations may drastically reduce the availability

121

of suitable imagery. As a consequence, a large amount of ancillary data (precipitation and

snow falls, cloud cover, etc.) is required to determine the most appropriate year, season

and day to acquire a suitable satellite image, which may not correspond to the acquisition

plans, temporal resolution, and availability of a specific satellite platform dataset.

Ancillary data may not be available for remote areas where remote sensing studies are

best applied, leading to a try-fail process for the selection of images from a dataset. All

the physical limitations mentioned were encountered during the acquisition and the

analysis of the data for this research study.

While methodological limitations concerning classification techniques can be

improved from ground control samples, and glacier dynamic estimation models can be

optimized with further understanding of the processes involved in mass balance

variations, the lack of suitable data and spectral capability cannot be readily fixed,

especially when the research needs to address past events.

Conclusions

The research conducted by me and reported herein produced an estimate of the

snow and ice cover changes affecting the glaciers of the Cordillera Real. The results are

in line with field observations of a gradual but quantifiable retreat from 1987. The snow

and ice cover of the Cordillera Real is highly sensitive to changes in regional weather

patterns such as those induced by El Niño events, but the progressive retreat of the

glaciers may also be related to long-term global change processes in the troposphere such

as the increasing planetary temperature.

122

The glaciers of the Cordillera Real developed in a tropical climate which is

characterized by a homogenous atmosphere and well defined temperature ranges.

Consequently, it could be argued that the measured loss of snow and ice surface, the

reduction of glacial mass, and the migration of the equilibrium line at higher elevation,

reflect re-adjustments of tropical environmental balances at regional scale. While the

cause of these environmental changes is currently under debate, it is important to observe

that the changes are significantly visible and have been monitored and measured.

Remote sensing based technologies may be implemented to identify and classify

glacier facies of remote areas, and to estimate environmental changes from the

measurement of the variation of physical properties of the glacier surface. A number of

limitations, though, affect the quality of the estimates, and must be taken into account

when discussing the resulting environmental change estimates. The error of estimate is

characteristic of any remotely sensed measurement due to the nature of the measurement

itself, which is not directly taken from on the object, but inferred from afar.

The accuracy of remotely sensed snow and ice cover is difficult to assess due to

the transitional nature of snow, and the continued overlapping variation of its surface

extent which requires continuous in-field monitoring. This research focused on the

necessity of developing a methodology to readily obtain measurements of snow and ice

changes to describe glacier retreat in remote areas, accepting the inherent error of

estimate resulting from the study of unexplored and extended areas.

This study confirmed that, while not developed for glaciological applications,

Landsat platforms are a valuable resource to study snow and ice cover, but present

serious limitations in snow covered tropical alpine environments. Mass balance variations

123

of the glaciers have been recognized as an optimum indicator of global environmental

changes; glaciers are also an important source for water for hydroelectric power and

water supply for a significant amount of population in alpine areas worldwide. It is thus

necessary to plan and develop satellite platforms dedicated to measure spatial and

temporal variations of the snow and ice cover of the glacier; dedicated satellite systems

would make it possible to estimate quickly and more accurately the environmental

variations driven by global changes processes, and would provide higher quality data to

ensure the development of prompt response plans to mitigate those changes that could

have negative impacts on human and natural environments.

124

Bibliography

References Cited

Bamber, J. 2003. A review of current and future remote sensing techniques for mass balance determination. Symposium on mass balance of Andean glaciers and 1st mass balance workshop on Andean glaciers. Centro de Estudios Cientificos, Valdivia (Chile).

Bennet, M.R., N.F. Glasser. 1996. Glacial Geology: Ice sheets and Landforms. John Wiley & Sons, New York.

Bindschadler, R., J. Dowdeswell, D. Hall and J. Winther. 2001. Glaciological applications with Landsat-7 imagery: Early assessment. Remote Sensing and Environment 78: 163-179.

Bishop, M.P., and J.D. Colby. 2002. Anisotropic reflectance correction of SPOT-3 HRV imagery. International Journal of Remote Sensing 23(10):2125-2131

Braithwaite, R.J. 2002. Glacier mass balance: the first 50 years of international monitoring. Progress in Physical Geography 26 (1): 76-95.

Braithwaite, R.J. and Y. Zhang. 1999. Modeling changes in glacier mass balance that may occur as a result of climate changes. Geografiska Annaler 81 A: 489-496.

Chander, G., B. Markham. 2003. Revised Landsat 5 TM radiometric calibration procedures and post-calibration dynamic ranges. White Paper, USGS Landsat Project Website. Online at http://landsat7.usgs.gov/documents/L5TMCal2003.pdf.

Chavez, P.S. 1988. An improved dark-object subtraction technique for atmospheric scattering correction of multispectral data. Remote Sensing of Environment 24: 459-479.

Chavez, P.S. 1989. Radiometric calibration of Landsat Thematic Mapper multispectral images. Photogrammetric Engineering and Remote Sensing 55(9): 1285-1294.

Chavez, P.S. 1996. Image-based atmospheric corrections – Revisited and improved. Photogrammetric Engineering and Remote Sensing 62(9): 1025-1036.

125

Colby, J.D. 1991. Topographic normalization in rugged terrain. Photogrammetric Engineering and Remote Sensing 57(5): 531-537.

Colby, J.D. and P.L Keating. 1998. Land cover classification using Landsat TM imagery in the tropical highlands: the influence of anisotropic reflectance. International Journal of Remote Sensing 19(8): 1479-1500.

Davis, R.E., A.W. Nolin, R. Jordan, and J. Dozier. 1993. Toward predicting temporal changes of the spectral signature of snow in visible and near-infrared wavelengths. Annals of Glaciology 17: 143-148.

d'Orbigny, A.D. 1835-1847. Voyage dans l'Amérique Méridionale (le Brésil, la République orientale de l'Uruguay, la République Argentine, la Patagonie, la République du Chili, la République de Bolivia, la République du Pérou), executé pendant les années 1826, 1827, 1828, 1829, 1830, 1831, 1832 et 1833 [Voyage to South America (Brazil, Uruguay, Argentina, Patagonia, Chile, Bolivia, Perú), carried out during the years 1826-1833]. Paris, Pitois-Levrault and Company 9v.

Dozier, J. 1989. Spectral signature of alpine snow cover from Landsat Thematic Mapper. Remote Sensing of Environment 28: 9-22.

Dozier, J., and D. Marks. 1987. Snow Mapping and Classification from Landsat Thematic Mapper data. Annals of Glaciology 9: 97-103.

EDC. 2003. Landsat Pathfinder Program. Land Processes Distributed Active Archive Center (LP DAAC). Earth Observatory System Data Center (EDC). Online documents version at http://edcdaac.usgs.gov/pathfinder/pathpage.html

ERDAS. 1997. ERDAS Imagine, ERDAS Field Guide. Leica Geosystems GIS & Mapping, Atlanta, GA.

ESRI. 2003. ArcGIS 8.3 Online Help Documentation. ESRI Redlands, CA.

Fily, M., B. Bourdelles, J.P. Dedieu, and C. Sergent. 1997. Comaprison of In Situ and Landsat Thematic Mapper derived snow grain characteristics in the Alps. Remote Sensing of Environment 59: 452-460:

Finsterwalder, R. 1987. Map of the Cordillera Real Nord (Illampu). München, Deutsche Alpenverein, scale 1:50,000.

Finsterwalder, R. 1990. Map of the Cordillera Real Süd (Illimani). München, Deutsche Alpenverein, scale 1:50,000.

Francou, B. 1993. Une représentation factorielle des cryosphères d'altitude dans le monde in Géomorphologie et aménagement de la montagne--melanages in hommage à P. Gabert [A quantitative description of the high-altitude cold regions in the world in Geomorphology and development of the mountain--a series of

126

papers dedicated to Pierre Gabert]: Centre National de la Recherche Scientifique, Centre de Geomorphologie de Caen, Bulletin 42: 75-86.

Francou, B., M. Vuille, P. Wagnon, J. Mendoza and J. Sicart. 2003. Tropical climate change recorded by a glacier in the central Andes during the last decade of the 20th century: Chacaltaya, Bolivia, 16ºS. Journal of Geophysical Research, 108(0).

Gao, J. and Y. Liu. 2001. Application of remote sensing, GIS and GPS in glaciology: A review. Progress in Physical Geography 25 (4): 520-540.

Haeberli, W. 2003. Systematic long-term mass-balance observations in the Andes as a high-priority need in worldwide climate-related glacier observation programs. Symposium on mass balance of Andean glaciers and 1st mass balance workshop on Andean glaciers. Centro de Estudios Cientificos, Valdivia (Chile).

Haeberli, W., R. Frauenfelder, M. Hoelzle, and M. Maisch. (1999). On rates and acceleration trends of global glacier mass changes. Geografiska Annaler 81 A:585-591.

Haeberli, W. and M. Hoelzle. 1994. Application of inventory data for estimating characteristics of and regional climate change effects on mountain glaciers – A pilot study with the European Alps. IGS symposium on the role of the Cryosphere on global change, Columbus Ohio.

Hall, D.K., A.T.C. Chang, J.L. Foster, C.S. Benson, and W.M. Kovalick. 1989. Comparin of In Situ and Landsat Derived Reflectance of Alaskan Glaciers. Remote Sensing of Environemnt 28: 23-31.

Hall, D.K., A.T.C. Chang, and H. Siddalingaiah. 1988. Reflectance of glaciers as calculated using Landsat-5 Thematic Mapper data. Remote Sensing of Environment 25: 311-321.

Hall, D.K., J.L. Foster, D.L. Verbyla, and A.G. Klein. 1998. Assessment of snow-cover mapping accuracy in a variety of vegetation-cover densities in central Alaska. Remote Sensing of Environment 66: 129-137.

Hall, D.K., J.P. Ormsby, R.A. Bindschadler, and H. Siddalingaiah. 1987. Characterization of snow and ice reflectance zones on glaciers using Landsat Thematic Mapper data. Annals of Glaciology, 9: 104-108.

Hall, D.K., G.A. Riggs, and V.V. Salomonson. 1995. Development of methods for mapping global snow cover using moderate resolution imaging spectroradiometer data. Remote Sensing of Environment 54: 127-140.

Herrmann, A. 1993. Review of Jordan, Ekkehard, 1991. The glaciers of the Bolivian Andes, a photogrammetric-cartographical inventory of the Bolivian glaciers as a

127

basis for climatic interpretation and potential economic use. Catena 20(3): 351-353.

Hooker, B.L. and B.B. Fitzharris. 1999. The correlation between climatic parameters and the retreat and advance of Franz Josef Glacier, New Zealand. Global and Planetary Change 22: 39-48.

Huang, C., L. Yang, C. Homer, B. Wylie, J. Vogelmann and T. DeFelice. 2003. At-satellite reflectance: a first order normalization of Landsat 7 ETM+ images. Raytheon ITSS EROS Data Center, Sioux Falls, SD. U.S. Geological Survey online document at http://landcover.usgs.gov/pdf/huang2.pdf .

IPCC, 2001. Climate Change 2001: The Scientific Basis. Published for the Intergovernmental Panel on Climate Change. Cambridge University Press, NY.

Irish, R. 1998. Landsat 7: Science Data Users Handbook. Goddard Space Center. Greenbelt, MD. Online at http://ltpwww.gsfc.nasa.gov/IAS/handbook/handbook_toc.html.

Jordan, E. 1991. Die Gletscher der bolivianischen Anden, eine photogrammetrisch-kartographische Bestandsaufnahme der Gletscher Boliviens als Grundlage für klimatise Deutungen und Potential für die wirtschaftliche Nutzung [The glaciers of the Bolivian Andes, a photogrammetric-cartographical inventory of the Bolivian glaciers as a basis for climatic interpretation and potential for economic use]. Stuttgart, Franz Steiner Verlag, 401p. maps.

Jordan, E. 1999. Glaciers of Bolivia. Satellite Image Atlas of the World. USGS professional paper 1386-I-5. http://pubs.usgs.gov/prof/p1386i/index.html. Accessed on 12/12/03.

Jordan, E., C. Brockman, A. Fernandez, R. Alvarez and K. Jacobsen. 1980, The glacier inventory of Bolivia in world glacier inventory, proceedings of the Riederalp (Switzerland) workshop. International Association of Hydrological Sciences, IAHS-AISH 126: 25-32.

Jordan, E. and R. Finsterwalder. 1992. Observaciones respecto al mapa Cordillera Real Norte (Illampú) 1:50,000--una contribución a la representación cartográfica y a la glaciología e historia de los glaciares de los Andes bolivianos [Explanatory notes to the map Cordillera Real North (Illampu) 1:50,000--A contribution to the cartographic representation, glaciology, and glacial history of the Bolivian Andes]. La Paz, Instituto Geográfico Militar: 178p. [In Spanish, German, English, and French.]

JPL, 2003. SAR Interferometry. Document, Courtesy of the Jet Propulsion Laboratory. http://www-radar.jpl.nasa.gov/sect323/InSar4crust/SarInterferometry.html Accessed on 06/2003.

128

Kaser, G. 1999. A review of modern fluctuation of tropical glaciers. Global and Planetary Change 22: 93-103.

Keller, K., G. Casassa, A. Rivera, R. Forsberg and N. Gundestrup. 2003. Mass balance of glaciers in Patagonia by means of airborne laser altimetry. Symposium on mass balance of Andean glaciers and 1st mass balance workshop on Andean glaciers. Centro de Estudios Cientificos, Valdivia (Chile).

Klein, A.G. and B.L. Isacks. 1998. Alpine glacial geomorphological studies in the central Andes using Landsat thematic mapper images. Glacial Geology and Geomorphology, rp01/1998. http://ggg.qub.ac.uk/ggg/papers/full/1998/rp011998/rp01.htm

Klein, A.G. and B.L. Isacks. 1999. Spectral mixture analysis of Landsat thematic mapper images applied to the detection of transient snowline on tropical Andean glaciers. Global and Planetary Change 22: 139-154.

Knap, W.H., C.H. Reijmer. 1998. Anisotropy of the reflected radiation field over melting glacier ice: measurements in Landsat TM bands 2 and 4. Remote Sensing of Environment 65: 93-104.

Lange, H., J. Araos and A. Rivera. 2003. Ice elevation changes and ice velocities of glaciar Juncal Norte, Chile. Symposium on mass balance of Andean glaciers and 1st mass balance workshop on Andean glaciers. Centro de Estudios Cientificos, Valdivia (Chile).

Markham, B.L., G. Chander. 2003. Revised Landsat 5 TM Radiometric Calibration Procedures and Post-Calibration Dynamic Ranges. http://landsat7.usgs.gov/documents/L5TMCal2003.pdf, White Paper, USGS Landsat Project Website.

Markham, B.L., J.L Barker. 1986. Landsat MSS and TM post-calibration dynamic ranges, exoatmospheric reflectances and at-satellite temperatures. EOSAT Landsat Data User Notes. Lanham, MD.

Martin, C. 2004. The phase diagram of water. London South-Bank University. Online document at http://www.lsbu.ac.uk/water/phase.html.

Mather, P.M. 1999. Computer processing of remotely-sensed images: An introduction. John Wiley & Sons Inc. New York, NY.

McKnight, T.L. 2000. Physical Geography: a landscape appreciation. Prentice-Hall Inc. Upper Saddle River, NJ.

Mercer, J.H. 1967. Glaciers of Bolivia and of Chile north of latitude 23ºS; in Southern Hemisphere glacier atlas. U.S. Army Natick Laboratories, Earth Sciences Laboratory, Series ES-33, Technical Report 67-76-ES: 65-81.

129

Minnaert, J.L. 1941. The reciprocity principle in lunar photometry. Astrophysics Journal 93:403-410.

Moran, M.S., R.D. Jackson, P.N. Slater and P.M. Teillet. 1992. Evaluation of simplified procedures for retrieval of land surface reflectance factors from satellite sensor output. Remote Sensing of Environment 41:169-184:

Nave, C.R. 2000. Mie Scattering. Georgia State University. Online document at http://hyperphysics.phy-astr.gsu.edu/hbase/atmos/blusky.html.

NASA. 2003. Landsat 7: Documentation. National Aeronautics and Space Administration. Online version at http://landsat.gsfc.nasa.gov/main/documentation.html.

Nolin, A.W., J. Dozier. 2000. A hyperspectral method for remotely sensing the grain size of snow. Remote Sensing of Environment 74: 207-216.

Nolin, A.W., J. Dozier, and L.A.K Mertes. 1993. Mapping alpine snow using a spectral mixture modeling technique. Annals of Glaciology 17: 121-124.

Painter, T.H., J. Dozier, D.A. Roberts, E.D. Robert, and R.O. Green. 2003. Retrieval of subpixel snow-covered area and grain size from imaging spectrometer data. Remote Sensing of Environment 85: 64-77.

Paul, F., A. Kaab, M. Maisch, T. Kellenberger and W. Haeberli. 2002. The new remote-sensing-derived Swiss glacier inventory: I. Methods. Annals of Glaciology 34: 355-361.

Riaño, D., E. Chuvieco, J. Salas, and I. Aguado. 2003. Assessment of different topographic corrections in Landsat TM data for mapping vegetation types. IEEE Transactions on Geoscience and Remote Sensing 41(5):1056-1061

Rignot, E., A. Rivera and G. Casassa. 2003. Volume changes and ice dynamics of the Patagonian icefields from SAR interferometry. Symposium on mass balance of Andean glaciers and 1st mass balance workshop on Andean glaciers. Centro de Estudios Cientificos, Valdivia (Chile).

Rivera, A. and G. Casassa. 1999. Volume changes on Pio XI glacier, Patagonia: 1975-1995. Global and Planetary Change 22: 233-244.

Rivera, A., T. Benham, G. Casassa, J. Bamber and J. Dowdeswell. 2003. Glaciological application of ASTER DEM from Campo de Hielo Norte, Patagonia, Chile. Symposium on mass balance of Andean glaciers and 1st mass balance workshop on Andean glaciers. Centro de Estudios Cientificos, Valdivia (Chile).

Rosenthal, W, J. Dozier. 1996. Automated mapping of montane snow cover at subpixel resolution from Landsat Thematic Mapper. Water Resources Research, 32(1): 115-130.

130

RSI. 2001. ENVI User’s Guide. Research System Inc. Boulder, CO.

Schnirch, M., C. Schneider, G. Casassa and R. Kilian. 2003. Glacier inventory of Peninsula Munoz Gamero, Patagonia, and glacier change during recent decades. Symposium on mass balance of Andean glaciers and 1st mass balance workshop on Andean glaciers. Centro de Estudios Cientificos, Valdivia (Chile).

Schott, J. 1997. Remote Sensing: The image chain approach. Oxford University Press. New York.

Skvarca, P., M. Stuefer and H. Rott. 1999. Temporal changes of Glacier Mayo and Laguna Escondida, southern Patagonia, detected by remote sensing data. Global and Planetary Change 22: 245-253.

Smith, J., T. Lin, and K. Ranson. 1980. The Lambertian assumption and Landsat data. Photogrammetric Engineering and Remote Sensing 46:1183-1189.

Teillet, P.M., B. Guidon, and D.G. Goodenough. 1982. On the slope-aspect correction of multispectral data. Canadian Journal of Remote Sensing 51:229-235.

Teillet, P.M., J.L. Barker, B.L. Markham, R.R. Irish, G. Fedosejevs, and J.C. Storey. 2001. Radiometric cross-calibration of Landsat-7 ETM+ and Landsat-5 TM sensors based on tandem data sets. Remote Sensing of Environment 78: 39-54.

Thompson, L.G. 2000. Ice core evidence for climate change in the Tropics: Implications for our future. Quaternary Science Review 19: 19-35.

Thompson, L.G. 2003. Personal interview at Appalachian State University, NC.

Thompson, L.G. and M.E. Davis. 1998. A 25,000 year tropical climate history from Bolivian ice cores. Science 282 (5395): 1858-1865.

Troll, C. and R. Finsterwalder. 1935. Die Karten der Cordillera Real und des Talkessels von La Paz and die Diluvialgeschichte der zentralen Anden [The maps of the Cordillera Real and the valley of La Paz and the Pleistocene history of the Central Andes]. Petermanns Geographische Mitteilungen 81(11): 393-399; 81(12): 445-455.

USGS. 1999. Satellite Image Atlas of Glaciers of the World. U.S. Geological Survey Professional Paper 1386-I. Online at http://pubs.usgs.gov/prof/p1386i/

USGS. 2003. Landsat-7 Level-0 and Level-1 Data Sets Document. U.S. Geological Survey. Online at http://eosims.cr.usgs.gov:5725/DATASET_DOCS/landsat7_dataset.html.

Warren, S. G. 1982. Optical properties of snow. Review of Geophysics and Space Physics, 20: 67-89.

131

Xiao, X., Z. Shen and Z. Qin. 2001. Assessing the potential of Vegetation sensor data for mapping snow and ice cover: a Normalized Difference Snow and Ice Index. International Journal of Remote Sensing 22(13):2479-2487.

Zheng, Q., C.M. Cao, X. Feng, F. Liang, X. Chen, W. Sheng. 1984. Study on spectral reflection characteristics of snow, ice and water of northwest China. Scienctia Sinica 27:647-656.

References Consulted

Bishop, M.P., J.F. Shroder, and J.D. Colby. 2003. Remote sensing and geomorphometry for studying relief production in high mountains. Geomorphology 55: 345-361.

Demuth, M. and A. Pietroniro. 1999. Inferring glacier mass balance using Radarsat: Results from Peyto Glacier, Canada. Geografiska Annaler 81 A: 521-540.

Hall, D.K., K.J. Bayr, W. Schoner, R.A. Bindschadler, and J.Y.L. Chien. 2003. Consideration of the error inherent in mapping historical glacier positions in Austria from the ground and space (1893-2001). Remote Sensing of Environment 86: 566-577.

Kaab, A., F. Paul, M. Maisch, T. Kellenberger and W. Haeberli. 2002. The new remote-sensing-derived Swiss glacier inventory: II. First Results. Annals of Glaciology 34: 362-366.

Li, Z, W. Sun and Q. Zeng. 1998. Measurements of glacier variation in Tibetan Plateau using Landsat data. Remote Sensing of Environment 63:258-264.

McGuffie, K. and A. Henderson-Sellers. 1997. A climate modeling primer. John Wiley & Sons, New York.

Meier, M.F. 1973. Evaluation of ERTS imagery for mapping and detection of changes in snow cover on land and on glaciers. Symposium on significant results obtained from earth resources technology satellite-1, NASA SP-327: 863-875.

Ribstein, P., E. Tirau, B. Francou and R. Saravia. 1993. Tropical climate and glacier hydrology: A case study in Bolivia. Journal of Hydrology 165: 221-234.

Thompson, L.G. 1997. Tropics on Ice. Earth 6(5):38-46.

Winther, J.G., D.K. Hall. 1999. Satellite-derived snow coverage related to hydropower production in Norway: present to future.

132

Appendix A.

Anisotropic Reflectance Correction and Topographic Normalization of

rugged terrains and snow covered surfaces.

Concepts of Anisotropic Reflectance and Topographic normalization

One of the most complex questions remote sensing research needs to address is

the question of how the energy leaving a surface is angularly distributed along the

hemisphere centered on the point of reflection (Schott 1997). Depending on the physical

properties of the illuminated surface, the incoming energy could be perfectly reflected in

one direction (specularly, such as in a mirror), perfectly diffused in all directions, or as in

nature, vary between nearly diffuse to nearly specular.

An ideal flat surface that diffuses electromagnetic radiation perfectly in all

directions is defined as Lambertian. As a result, the radiation measured from any

observation angle is the same, and the surface is said to present isotropic reflectance. In

this situation, the amount of the reflected energy is only a function of the angle of

incidence and of the magnitude of the incoming radiation. In remote sensing research,

surfaces are often assumed to be Lambertian in order to simplify models and reduce

computation time. A flat surface that presents nearly diffused to nearly specular reflection

will reflect energy unevenly. Measurements of reflected radiation will vary with the

observation angle by a complex function. In this case, the surface is defined non-

Lambertian and it presents anisotropic reflectance. The reflected energy measured by an

133

observer (such as a satellite sensor) will depend on the geometry of the source (incidence

angle, azimuth and magnitude), and by the geometry of the observer (exitance angle and

azimuth) (Colby and Keating 1998; Riaño et al. 2003). The bidirectional reflectance

distribution function (BRDF), in conjunction with anisotropic reflection models, is used

to describe bidirectional reflectance values for all combinations of input-output angles

(Schott 1997).

Most of the targets and landcovers on the surface of the Earth present a certain

degree of anisotropic reflectance, which could more easily be corrected if all the surfaces

were ideally flat, however, the irregular topography of the Earth’s surface introduces a

further complication. In fact, even if source and observer maintain a constant geometry,

the orientation of the targets may vary depending on local slope and aspect. As a result,

the amount of solar energy reflected toward the sensor will vary as a function of (1)

source and observer geometry, and (2) slope and aspect of the local terrain (Colby 1991;

Colby and Keating 1998). This phenomenon, known as topographic effect, make surfaces

appear brighter when exposed to direct illumination and darker otherwise. This

differential illumination is perceived visually as relief in photographic pictures and

optical satellite images (Colby and Keating 1998). Topographic normalization techniques

aim to eliminate, or reduce, the topographic effect to obtain surface brightness values as

if the surface was flat.

Topographic Normalization: Methods and Techniques

Three techniques available to perform topographic normalization are band

ratioing, Lambertian models, and non-Lambertian models. While band ratioing relies on

134

in-scene information, Lambertian and non-Lambertian methods require additional data in

the form of a digital elevation model (DEM) to extract slope and aspect of the local relief.

Also, local knowledge about the landcover may improve the efficiency and accuracy of

non-Lambertian models.

Topographic normalization techniques based on band ratioing assume that the

relative reflectance of spectral bands for a specific landcover is independent of

illumination, thus the ratio of two spectral bands will be a constant not affected by the

topographic effect (Schott 1997; Mather 1999). Band ratioing has been considered to be a

suitable methodology to normalize the terrain effect because it presents low

computational cost and does not require additional data; nevertheless, researchers argued

that variations of the reflectance of a surface may be misinterpreted since different

surfaces may present similar reflective properties, thus leading to misclassification

(Colby 1991). Also, it is argued that the assumption of constant proportionality between

spectral bands is not supported by physical models. In fact, in case of diffuse radiance,

the pattern of the reflected energy depends on wavelength and incident angle (Riaño et al.

2003).

The Lambertian model implemented to perform topographic normalization

assumes that the surface investigated diffuse the reflected energy uniformly across an

ideal hemisphere centered on the reflection point. Also, Lambertian models state that the

intensity of the energy Ln measured from any observation angle is proportional to the

incoming incident radiation L, and inversely proportional to the cosine of the incident

angle (equation A.1) (Schott 1997; Colby and Keating 1998; Mather 1999; Riaño et al.

2003).

135

iLLn cos

= (equation A.1)

Three main problems arise when considering natural surfaces to be Lambertian.

First, it can be argued that while there are natural landcovers that present Lambertian

reflection, such as fresh snow (Knap and Reijmer 1998), most of the Earth’s landcovers

are characterized by anisotropic reflectance, such as vegetated areas. Second, Earth’s

surface can be characterized by complex topography, thus the incident angle cannot be

simply represented by the source geometry (sun elevation), but must be corrected for the

observer position considering slope and aspect at the incident point. The most accurate

Lambertian topographic normalization methodologies take into account both source and

observer geometries. Nevertheless, a third flaws affects the Lambertian models. Being

inversely proportional to the cosine of the incident angle, the Lambertian model over

estimates radiance calculated for incident angles proximate to vertical (reflection plane

close to perpendicularity). This phenomenon is due to the fact that the cosine approaches

zero when the incident angle is close to vertical, thus amplifying the estimated reflectance

radiance toward infinity. This effect is particularly visible for sun-facing slopes, or for

flat surfaces recorded by nadir sensors and high solar elevation (Schott 1997; Colby and

Keating 1998).

Non-Lambertian models are based on semi-empirical models, and include the

Minnaert correction (Minnaert 1941), the Teillet correction (Teillet et al. 1982), and the C

correction (Teillet et al. 1982). Riaño et al. (2003) evaluated the application and response

of different non-Lambertian methods. Colby (1991) assessed the efficiency of the

Minnaert correction for topographic normalization of rugged terrain. Later, Colby and

136

Keating (1998) and Bishop and Colby (2002) further discussed and successfully applied

the Minnaert correction to normalize topography in tropical montane regions, and in the

Himalayas. The Minnaert correction can be derived using the backward radiance

correction transformation (BRCT) (Smith et al. 1980), and is based on the estimation of

semi-empirical constants, called Minnaert coefficients, that are dependent on the spectral

bands (wavelength) and on the physical properties of the landcover (Colby 1991). For a

fast, less accurate result, the Minnaert coefficients necessary to normalize each spectral

band could be estimated on a scene-wide scale. A more accurate approach, which takes

into account local knowledge about the physical properties of the landcovers investigated,

was used to estimate Minnaert coefficients that correct spectral bands on a landcover

basis, thus increasing considerably the accuracy and performance of the methods.

Case Study: Image Processing

The Minnaert correction was found to be particularly

suitable to normalize Landsat imagery of rugged terrains in

tropical environments (Colby and Keating 1998). The Minnaert

correction was chosen to perform the topographic normalization

of two subset areas of a 1987 Landsat TM scene of the Cordillera

Real. The first subset (about 59 km2) represents a portion of the

Ancohuma-Illampu massif and contains mainly snow and ice

covers (figure A.1), while the second subset (about 43 km2)

Figure A.1. Snow and ice cover sample of approximately 59 km2.

Snow and Ice

137

represented a portion of rugged terrains

characterized by barren soils and low sparse

vegetation (figure A.2). A digital elevation model

(DEM) of the year 2003 was obtained from the

available free dataset of the satellite ASTER.

The Minnaert correction theory assumes

that the pattern of scattered radiation can be

described by the bidirectional reflection

distribution function (BRDF) which requires the calculation of incidence and exitance

angles. The Minnaert constants k can be used as an approximate representation of the

BRDF.

Equation A.2 represents a non-Lambertian reflection model based on Minnaert

correction; L is the incoming radiance, i is the angle of incidence, e is the angle of

exitance, and k is a Minnaert constant.

eieLL kkn coscos

cos= (equation A.2)

For each spectral band n, a Minnaert constant kn can be estimated by regression of pre-

processed layers obtained from the DEM and the Landsat scene (Colby 1991). To begin,

it is necessary to rewrite Equation A.2 as follow:

eiLeL kkn coscoscos = . (equation A.3)

Then a mathematical transformation enable k to be represented as a linear coefficient:

eikLeL n coscosloglogcoslog += . (equation A.4)

Figure A.2. Soils and vegetation sample of approximately 43 km2.

Soils and low vegetation

138

Equation A.4 is in the form of a linear equation y=kx+b where y=logLcose,

x=logcosicose, and b=logLn. The x term (independent variable) and y term (dependent

variable) can be used in a regression model to estimate the k constant for each spectral

band.

The available DEM was rectified, co-registered, and re-sampled to 28.5 m to

match the spatial characteristics of the Landsat scene. The DEM was then processed to

obtain slope and aspect layers. The calculation of slope and aspect represent a necessary

step to compute the correct angle of incidence i and angle of exitance e for each pixel of

the scene. In fact, the angles of incidence and exitance depend on the geometry of the

source (solar illumination), the position of the observer (nadir satellite sensor in this

case), and the local topography. The next step involved the calculation of the cosine of

the angle of incident (cosi) and of the angle of exitance (cose). The product of the cosines

(cosicose) and the logarithm of the product (logcosicose) were also calculated.

The Landsat scene had been previously pre-processed to obtain radiance layers L

and calculate the logarithm of the product between the cosine of the angle of exitance and

the radiance for each spectral band (logLcose).

Finally, the Minnaert coefficients kn (table A.1 and table A.2 ) were obtained by

regression using the member logcosicose as the independent variable (x term), and the

member logLcose as the dependent variable (y term). This procedure was performed on

both subsets, and for all the available spectral bands (visible, near infrared, and mid-

infrared).

139

k 1 0.670794658955001 k 1 0.725530599314385k 2 0.841102853169116 k 2 0.932826820287288k 3 0.800794183002669 k 3 0.897134115851479k 4 0.766390160049200 k 4 0.963814659244721k 5 0.811673282184080 k 5 0.704773225411237k 7 0.847638785533782 k 7 0.355635525943596

Table A.1. Rugged terrain subset Table A.2. Ancohuma-Illampu glaciers

Soils and low vegetationMinnaert coefficientsSnow and ice cover

Minnaert coeffcients

Case Study: Results and Discussion

Figures A.3 and A.4 present the subsets before and after topographic

normalization. A first order analysis of the results based on visual inspection reveals that

the soil and low vegetation normalized sample is characterized by a general optical

flattening of the topography. This is in agreement with the concept of topographic

normalization which aims to reduce topography to a flat horizontal surface, and

homogenize brightness variations induced by slope and aspect.

Figure A.3. Soils and vegetation sample before and after topographic normalization

Before After

140

The visual difference between corrected and uncorrected snow and ice cover

samples, instead, appear to be less pronounced. This is mostly attributed to the saturation

of the Landsat sensors in the visible and near-infrared bands which induces brightness

cut-off, and thus the optical flattening of the topography in the original sample. Also,

fresh and non-diagenized snow present a strong Lambertian reflection (Knap and Reijmer

1998), which partly masks the terrain effect in optical imagery.

To achieve a better understanding of the results, it is necessary to proceed with a

second order analysis to compare the data characteristics before and after the

normalization process. The hypothesis is that the original subset should present some

correlation between the brightness values and the incidence angle, in particular, the

reflectance registered by the sensor should be a function of the incidence angle. On the

other hand, in the normalized subsets, the brightness value should be largely uncorrelated

to the incident angle, since the normalization process should remove variation of

reflectance due to topography. To validate this test, it is possible to perform a regression

between the cosine of the incident angle (cosi) (independent variable) and the reflectance

Figure A.4. Snow and ice cover sample before and after topographic normalization.

Before After

141

(dependent variable) before and after normalization. The regression analysis was

achieved with IDIRISI Kilimanjaro (Clark Labs at Clark University). and the results for

the soil and low vegetation covers (band 3) are presented in figure A.5 and figure A.6.

It is possible to observe that, in fact, the coefficient of correlation between the

cosine of the incident angle (cosi) and the reflectance drops from 34.12% (figure A.5) to

6.19% after the topographic normalization (figure A.6). This change is significant and

indicates that the topographic normalization has reduced the geometric correlation

between incident angle and brightness values.

In the case of the snow and ice cover subset (figure A.7 and figure A.8 showing

band TM 3), the results needs to be interpreted in lieu of the characteristics of the

landcover. It can be observed that the normalized dataset presents a low degree of

correlation (4.33%), but it could be argued that the natural Lambertian properties of the

snow already reduces the correlation between incidence angle and reflectance (19.90% in

the source subset), and also, that due to the effect of sensor saturation over snow and ice

covers, the radiance values returned by the normalization process are based on cut-off

brightness values and thus, the normalized dataset do not represent the real physical

properties of snow and ice.

142

Figure A.5. Regression analysis for the soil and low vegetation subset (band TM 3) before normalization.

Figure A.6. Regression analysis for the soil and low vegetation subset (band TM 3) after normalization.

143

A further example (figure A.9) showing the regression of the cosine of the

incident angle and reflectance of TM band 5 seems to support the near Lambertian

reflection of snow covers. The TM Band 5 does not saturate over snow fields, thus the

reflectance values are in good agreement with the optical properties of snow. It can be

Figure A.8. Regression analysis for the snow and ice cover subset (band TM 3) after normalization.

Figure A.7. Regression analysis for the snow and ice cover subset (band TM 3) before normalization

144

observed that for TM band 5 the coefficient of correlation in the source subset is only

9.56%, showing a very low correlation between angle of incidence and reflectance.

After normalization, the coefficient of correlation drops to 4.11%, a value not

significantly different from the uncorrected dataset.

Limitations and Final Consideration

Topographic normalization and anisotropic reflectance correction techniques are

of utmost importance to study rugged terrain and alpine environments. The case study

highlighted that non-Lambertian normalization based on Minnaert coefficient is effective

in reducing the terrain effect over soils and vegetation in Landsat imagery, but did not

significantly improve the physical information over snow and ice covers. This was

mainly due to (1) the Lambertian properties of fresh snow which naturally reduce the

optical terrain effect, and (2) the loss of information in the visible and near infrared

Figure A.9. Regression analysis for the snow and ice cover subset (band TM 5) before normalization.

145

Landsat bands caused by sensor saturation. Also, studies of the optical properties of snow

and ice have shown that aging snow cover presents a strong forward scattering that can

be efficiently corrected with empirical expressions (or C corrections) (Knap and Reijmer

1998). Rosenthal and Dozier (1996) argued that the calculation of slope and aspect

amplifies the digital noise present in digital elevation models and introduces considerable

error in the classification of snow facies, instead they suggested the use of the

information obtained from elevation models to specify boundary conditions for radiative

transfer calculation by defining facies ranges from bi-spectral plots.

An important limitation of the non-Lambertian methodology resides in the use of

reliable and temporally consistent DEM datasets which are not always available for

remote areas such as in the case of the Cordillera Real. Future airborne and satellite

missions, and the release of classified information such as the Shuttle Radar Topography

Mission (SRTM) dataset, may improve the quality, availability, and worldwide coverage

of high resolution DEM to be implemented for topographic normalization.

146

Biographical Information

Christian Degrassi was born in Grado, Italy, on December 7th, 1969. He was

awarded the Bachelor of Science Degree in Geology from the University of Trieste, Italy

in 1994. In 1996 he was hired by Schlumberger Oilfield Services, Texas, where he had

the opportunity to travel and work worldwide in a high technology oriented environment.

In 2001 he began study toward a Master of Arts in Geography. The M.A. was awarded in

2004.