.. (I: ‘ 1 Elfin. . 4.1: 5v 514:“; :v 3 . ffam .6): .15.: H. 855.... z 4.31:?! in: . » tn..mm...r...3§ . EH 0’? «u’VIvthAn... ;- lil- u 31.”... 5 “v3... fiflwhfinwmma 5.» M; ‘,.. THESYS Y'x LIBRARY ”when State University This is to certify that the thesis entitled MICHIGAN FOREST DEFOLIATION DETECTION USING MULTITEMPORAL' NDVI SIGNATURES i DERIVED FROM AVHRR SATELLITE IMAGERY presented by Kenneth A. Duda has been accepted towards fulfillment of the requirements for M. A. degree in Geography eel/WM Major professor Date October 2, 1998 0.7639 MS U is an Affirmative Action/Equal Opportunity Institution PLACE IN REFURN BOX to rem TO AVOID FINES return on or - “I. ove this checkout from your record. before date due. MAY BE RECAILED with earlier due date if requested. DATE DUE DATE DUE DATE DUE MICHIGAN FOREST DEFOLIATION DETECTION USING MULTITEMPORAL NDVI SIGNATURES DERIVED FROM AVI'IRR SATELLITE IMAGERY by Kenneth A. Dude A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF ARTS Department of Geography 1998 ABSTRACT MICHIGAN FOREST DEFOLIATION DETECTION USING MULTITEMPORAL NDVI SIGNATURES DERIVED FROM AVI-IRR SATELLITE IMAGERY by Kenneth A. Dude Multitemporal Advanced Very High Resolution Radiometer (AVHFIR) satellite data and ancillary information stored in a geographic information system (GIS) were used to assess the effectiveness of employing maximum likelihood (ML) and backpropagation (BP) artificial neural network (ANN) classifiers to discriminate areas of forest defoliation caused by gypsy moths in the Lower Peninsula of Michigan and in Roscommon County during a single year. The AVHFIFI sensor offered comparatively high temporal resolution. The Normalized Difference Vegetation Index (NDVI) was calculated for each of three temporal composites corresponding to pre-, peak and post defoliation stages, then sampled image data were input to the classifiers. Overall accuracies ranged from 19.7% to 78.8%. Based on the Kappa statistic, results were similar for the ML and ANN methods in the Lower Peninsula, however the ML method markedly outperformed the BP ANN technique at the county level. and compared favorably with other reported results. The generalizing nature of the ML approach yielded greater accuracy than did the attempt by the ANN to seek detail while limited by the coarse spatial resolution of the imagery, which resulted in heterogeneous pixel content and limited availability of training samples. Copyright by KENNETH ALAN DUDA 1998 My parents. Clifford A. and Ellzabeth J. Dude, provided constant early encouragement as I sought answers to questions about our world, and they instilled in me the desire to make the best possible use of those abilities which I have been given. Always placing a high value on the pursuit and application of knowledge and the exercise of reason, their unwavering support throughout the years has contributed to all of my accomplishments, including this work, which I dedicate to them. ACKNOWLEDGEMENTS Foremost. I must acknowledge the insightful guidance of my graduate committee, in particular my advisor Daniel G. Brown. Ph.D., and also additional readers Bruce w. Pigozzl, Ph.D., David P. Luech. Ph.D. and Jon B. Burley, Ph.D. Their ready availability. timely suggestions and continual patience, especially during the latter phase of the project, helped bring this work to a rapid conclusion. In addition. I received recommendations concerning the application of artificial neural networks to satellite imagery classification during visits with Anil P. Jain, Ph.D., Computer Science Department, and Fathi M. A. Salem, Ph.D., Electrical Engineering Department, both at Michigan State University. I also received similar early counsel from Tom Vogl, Ph.D. and Tom German, both of the Environmental Research institute of Michigan (ERIM), as well as from Sucharita Gopal, Ph.D., Department of Geography, Boston University. The basic data were directly provided by the Center for Remote Sensing (CRS) at Michigan State University, and I completed some data preprocessing while employed there. Funding for the CRS project and additional support were provided by Frank Sapio of the Michigan Department of Natural Resources. Forest Management Division. I am ewedaIIy thankful for strong bonds of kinship and friendship. capable of weathering the extended periods of separation necessitated by the demands of this work. To new friends and old, and especially to my family, i extend my heartfelt thanks for your valued support, which was generously offered in many ways. TABLE OF CONTENTS LIST OF TABLES ................................................................................. viii LIST OF FIGURES ............................................................................... ix LIST OF ABBREVIATIONS ................................................................... xii CHAPTER 1.0 INTRODUCTION ......................................................... 1 1.1 Overview .............................................................................. 1 1 .2 Research Questions .............................................................. 4 CHAPTER 2.0 BACKGROUND .......................................................... 7 2.1 Gypsy Moth Defoliation .......................................................... 7 2.2 Satellite Imagery .................................................................... 15 2.2.1 Spectral Responses ........................................................ 17 2.2.2 AVHRR Characteristics .................................................. 18 2.3 Classification Techniques ....................................................... 24 2.3.1 Maximum Likelihood .................................................. 27 2.3.2 Artificial Neural Networks ........................................... 28 CHAPTER 3.0 DATA ......................................................................... 35 3.1 Study Area ........................................................................... 37 3.2 AVHRR Imagery ................................................................... 47 3.3 Land Cover ........................................................................... 53 3.4 Defoliated Areas .................................................................... 54 CHAPTER 4.0 METHODS .................................................................... 58 4.1 Data Processing ................................................................... 59 4.1.1 Preprocessing ............................................................. 59 4.1.1.1 Raster AVHRR Imagery and Land Cover Data..... 59 4.1.1.2 Defoliated Areas ................................................ 82 4.1.1.3 Sampling ........................................................... 65 4.1.1.4 Roscommon County Scene Extraction ................. 69 4.1.2 Maximum Likelihood Classification .............................. 70 4.1.3 Artificial Neural Network Classification ......................... 73 4.2 Accuracy Assessment ............................................................ 78 CHAPTER 5.0 RESULTS, DISCUSSION AND CONCLUSIONS ............. 81 5.1 Results ............................................................................... 81 5.1.1 Lower Peninsula Classification ................................... 82 5.1.2 Roscommon County Classification .............................. 84 5.2 Answers to Research Questions and Discussion ...................... 88 5.3 Conclusions and Recommendations for Future Research .......... 105 vi APPENDIX A Selected World Wide Web Resources ................................................. 112 REFERENCES CITED .......................................................................... 1 16 vii Table 4.1 . Table 4.2. Table 4.3. Table 5.1 . Table 5.2. Table 5.3. Table 5.4. Table 5.5. Table 5.6. Table 5.7. Table 5.8. LIST OF TABLES Computed linear transformation coefficients of the back transfon'nation....61 Lower Peninsula NDVI summary statistics .......................................... 72 Roscommon County NDVI summary statistics ..................................... 72 Error matrix for maximum likelihood classification of Lower Peninsula data ........................................................................................ 82 Error matrix for neural network classification of Lower Peninsula data ....... 83 Error matrix for maximum likelihood classification of Roscommon County data .............................................................................. 85 Error matrix for neural network classification of Roscommon County data ........................................................................................ 87 Lower Peninsula and Roscommon County defoliation by actual percent defoliated quartiles .................................................................... 93 Roscommon County maximum likelihood results for training sample range and for category outside training range .................................. 98 Roscommon County maximum likelihood defoliated accuracy and class homogeneity ............................................................................. 99 Areal extent and correct/Incorrect classification ratio for key values of percent defoliation .................................................................... 104 viii LIST OF FIGURES WW figure 1.1 Comparison of pattern recognition and 'inferencing capabilities for selected artificial intelligence techniques ................................... 3 CW Figure 2.1 Gypsy moth larva .................................................................... 8 figure 2.2 Global distribution of the gypsy moth........ .................................. 8 figure 2.3 Etienne Leopold Trouvelot ........................................................ 9 Figure 2.4 1990 distribution of the gypsy moth in North America ..................... 11 figure 2.5 25-year projection of gypsy moth spreading .................................. 12 figure 2.6 Annual acreage of Michigan gypsy moth defoliation ....................... 13 figure 2.7 Annual number of counties with Michigan gypsy moth defoliation ...... 14 figure 2.8 NOAA Polar Orbiter TiROS-N .................................................... 16 figure 2.9 2,399 kilometer swath of AVHRR coverage over North America ........ 20 figure 2.10 Artificial neural network architecture employed in this research ....... 29 mm Figure 3.1 Roscommon County in the State of Michigan ................................ 38 figure 3.2 Monme temperature profile for Houghton Lake, Michigan .............. 40 figure 3.3 Monme heating degree day profile for Houghton Lake, Michigan ...... 41 figure 3.4 Monthly cooling degree day profile for Houghton Lake, Michigan ....... 42 figure 3.5 Monme precipitation profile for Houghton Lake, Michigan ................ 42 figure 3.6 Monthly snowfall profile for Houghton Lake, Michigan ..................... 43 Figure 3.7 1993 timberland ownership in the Northern Lower Peninsula Unit..... 45 figure 3.8 1980 and 1993 areal extent of major foresttypes in the Northern Lower Peninsula Unit ........................................................... 47 figure 3.9 Lower Peninsula pro-defoliation NDVI ........................................ 50 figure 3.10 Lower Peninsula peak defoliation NDVI ...................................... 50 figure 3.11 Lower Peninsula post defoliation NDVI ...................................... 51 figure 3.12 Roscommon County pre-defcliaticn NDVI ................................... 51 figure 3.13 Roscommon County peak defoliation NDVI ................................ 52 figure 3.14 Roscommon County post defoliation NDVI ................................. 52 figure 3.15 Lower Peninsula forest mask ................................................. 54 figure 3.16 Michigan counties with observed defoliation in 1993 ..................... 56 W90: figure 4.1 Control points used in resampling ............................................... 62 figure 4.2 Reference sketch map for Roscommon County defoliation cluster ..... 64 figure 4.3 Complete Lower Peninsula reference data ................................... 68 figure 4.4 Sampled Lower Peninsula data .................................................. 68 figure 4.5 Roscommon County complete and sampled reference data ............. 69 WW figure 5.1 Frequency distribution of proportion defoliated for all Lower Peninsula defoliated pixels .................................................... 94 figure 5.2 Frequency distribution of proportion defoliated for all Roscommon County defoliated pixels ........................................................ 95 figure 5.3 Spatial distribution of proportion defoliated for all Roscommon County defoliated pixels ........................................................ 95 Figure 5.4 Frequency distribution of proportion defoliated for all Rommmon County defoliated pixels correctly classed as defoliated by ML classifier ............................................................................ 96 figure 5.5 Subpixel percent of range versus percent defoliation for Roscommon County ML classification ....................................................... 99 figure 5.6 Correct/incorrect ratio versus percent defoliation for Roscommon County ML classification ....................................................... 102 figure 5.7 Correct/Incorrect ratio versus percent defoliation for Lower Peninsula and Roscommon County ML and ANN classifications ................. 104 xi Al ANN AVHRR 80 BP 000 CDA CIR CRS DNR FAQ FIA FTP GAC GIS GLERL HRV LAC MCC MDNR ML MSS MVC NCAAS NCDC NDVI NESDIS NIR NOAA NODC NPV PAN PCC RMSE SPOT TD TM LIST OF ABBREVIATIONS Artificial Intelligence Artificial Neural Network Advanced Very High Resolution Radiometer Bhattachanya (or Jeffries-Matusita) Distance Backpropagation Charge-Coupled Device Command and Data Acquisition Color infrared Center for Remote Sensing Department of Natural Resources Frequently Asked Questions Forest Inventory and Analysis file Transfer Protocol Global Area Coverage Geographic information System Great Lakes Environmental Research Laboratory High Resolution Visible Local Area Coverage Midwestern Climate Center Michigan Department of Natural Resources Maximum Likelihood Multispectral Scanner Maximum Value Compositing NOAA CoastWatch Active Access System National Climatic Data Center Normalized Difference Vegetation index National Environmental Satellite, Data and Information Service Near Infrared National Oceanic and Atmospheric Administration National Oceanographic Data Center Neucleopoiyhederosis virsus Panchromatic Percent Correctly Classified Root Mean Squared Error Systeme Pour I'Observation de la Terra Transformed Divergence Thematic Mapper xii Chapter 1 INTRODUCTION 1.0 introduction Through the use of emerging technologies, this work attempted to improve upon previous efforts to efficiently identify widely-distributed areas of undesirable deciduous forest leaf-loss, or defoliation, caused by insect feeding. Successful implementation of the approach proposed in this effort would enable land managers, entomologists, ecologists, geographers and landowners to rapidly identify areas of leaf-loss as well as confirm the locations of unaffected regions, reduce associated monitoring costs, assist in targeting available control funds to those areas mom affected, and base decisions on known quantitative levels of analytical accuracy. 1.1 Overview The Forest Management Division of the Department of Natural Resources (MDNR) of the State of Michigan, represented by Frank Sapio, and the Center for Remote Sensing (CRS) at Michigan State University. represented by Dr. David P. Lusch, attempted to develop a rapid and relatively inexpensive technique to delineate the extent of deciduous forest defoliation caused by gypsy moths (Lymantn’a dispar L.) in Michigan (Lusch and Sapio, 1994). Current detection methods are labor intensive, time- consuming, costly, and based on outdated equipment that is nearing the end of its life expectancy. Specific areas of defoliation are normally first visually identified on airborne imagery acquired for the entire region, and then hand-sketched onto topographic maps which are used for later analysis. Funding was provided to CRS by MDNR to acquire and prepare suitable data for analysis. The purpose was to investigate the viability of using Earth-orbiting satellite data in order to detect portions of forested areas in the State of Michigan which had been defoliated by gypsy moths during one growing season. This investigation by MDNR/CR8 utilized Advanced Very High Resolution Radiometer (AVHRR) satellite imagery in conjunction with maps produced by MDNR of areas known to be defoliated, which were based on the interpretation of airborne video reconnaissance. Initial attempts did not yield conclusive results, and no further work was planned by the partnership on this problem (Lusch and Sapio. 1994). In this report, such Earth-orbiting satellite data and data captured by sensors in aircraft are referred to as ‘remoteiy sensed,’ and the monitoring and data acquisition activities are described as ‘remote sensing.’ This terminology is consistent with the usage by practitioners in the field, though it is truly nonspecific and could equally apply to other techniques, for example various surface and subsurface geophysical exploration procedures. While working at the Center for Remote Sensing at Michigan State University, one of my assigned tasks involved the preprocessing of spatial data stored in a geographic information system (GIS). These data were being prepared as part of the 2 aforementioned project undertaken by CRS on behalf of MDNR. At about the same time period, I read an article by McConnack and Day (1993) ccnceming the application of various artificial intelligence (Al) techniques to Earth investigations. Of particular interest was the mention made of the superior pattern recognition capabilities of the artificial neural network (ANN) approach. The authors noted that the ability of human intelligence to infer and to recognize patterns is the essence of its. power. A plot of pattern recognition capability versus inferencing ability was presented for expert systems, fuzzy logic, genetic algorithms and neural networks, included here as figure 1.1. Genetic algorithms and especially neural networks were shown to have superior pattern recognition capabilities and weaker inferencing abilities. Fuzzy logic and especially expert systems excel at inferencing, but do not periorrn as well at pattern recognition. This article, and my general interest in Al applications, led to further reading on the subject, in particular concerning artificial neural networks. Neural 1 Network Pattern 1 Genetic 1 Recognition Algorithm Capability Fuzzy Expert Logic System Inferenclng Ability Figure 1.1 Compariaon of pattern recognition and inferencing capabllltiea for eelected artificial intelligence techniques. After McConnack and Day (1993). 3 The research described in this thesis evolved from my desire to apply both conventional remote sensing image classification procedures and the less commonly utilized ANN approach to a multitemporal change detection problem, in this case involving a forest management issue affecting many states in the northeastem United States. It was hoped that insight would be gained concerning the relative strengths and weaknesses of each method, and also that a practical surveillance solution could be developed for the detection of affected areas. I further wished to gain familiarity with methods that might enable the integration of multisensor ‘remoteiy sensed’ data (not limited in this case only to satellite and aircraft observations) in order to develop characteristic feature classification signatures. 1.2 Research Questions The work conducted in this study quantitatively evaluated the ability of two supervised classifiers to discriminate areas of forest defoliation caused by gypsy moths through the integration of multitemporal and multispectral AVHRR imagery and ancillary spatial in- formation stored within a geographic inforrnaticn system. The problem was one involving two opposite forest states: leaf-on (nondefoliated) or leaf-off (defoliated), with the possible complicating existence of varying combinations of these two conditions. Supervised classification employs known outcomes for several representative input vectors to develop classification solutions for vectors with initially unknown class assignments. As explained in greater detail in section 2.2, the selection of the sensor involved a compromise between temporal and spatial resolution. with questions primarily arising concerning the adequacy of the spatial component in measuring the targeted features, given the scale of the defoliation process and the extent during the monitoring period. in addition, it was necessary for the temporal component to enable image acquisition at sufficient frequency to adequately sample the process being monitored. Satellite imagery has been frequently used in land cover studies. The maximum likelihood (ML) classification technique is one of the most commonly used methods (Foody, McCulloch and Yates, 1995a; Jensen, 1996). ANNs have been successfully utilized in several image classification tasks, and have often, but not always, produced higher levels of classification accuracy than conventional approaches. It was anticipated that the neural network method adopted for use in this research might yield higher classification accuracies than the conventional algorithm, by exploiting the widely cited robust nature of the technique in processing the temporally-composited, band-ratioed, coarse-resolution satellite data available for this study. Specific questions addressed through this research include the following: 1) Can the maximum likelihood supervised classification technique be successfully applied to the problem of detecting defoliated areas? 2) Can the ANN classification technique be successfully applied to the problem of detecting defoliated areas? 3) How do the accuracy levels of the maximum likelihood and the ANN techniques compare in this application, given the final parameters utilized? 4) Are the classification accuracy levels for defoliated and nondefoliated pixels similar? 5) At a subpixel level of detail, what insight may be gained from an analysis of pixel homogeneity? Chapter 2 BACKGROUND 2.0 Background To explore the possibility of differentiating between gypsy moth defoliated and nondefoli- ated forest land cover in Michigan, multichannel and band-ratioed multitemporal AVHRR satellite data were utilized. Analysis involved the use of commme available remote sensing, GIS, statistical and ANN software, with data spanning a period of approximately 3.5 months during the 1993 growing season. Forest defoliation by gypsy moths, satellite imagery characteristics and classification techniques are summarized in this Chapter. 2.1 Gypsy Moth Defoliation Defoliation can cause tree mortality and dramatically alter the character of the affected woodlot or forest. With some 300 million acres of susceptible forest cover throughout the United States, the impact on ecological balances, local economies. wildlife habitat. water quality and flow, and available shade can be substantial. The natural range of the gypsy moth is in Europe, North Africa and Asia, where it has a widespread distribution. Introduction to the United States occurred during the 19th century. The larva is shown in figure 2.1, while a map of the global distribution is provided In Figure 2.2. Figure 2.1 Gypsy moth larva. Source: USDA Forest Service, Northeastern Forest Experiment Station. Gypsy Moth in North America http://gypsy.fsl.wvnet.edu:80/gmoth/ 3/31/98 Figure 2.2 Global distribution of the gypsy moth. The natural range is shown as dark gray, the region of recent introduction is shown in light gray. Source: USDA Forest Service. Northeastern Forest Experiment Station. Modified from Gypsy Moth in the World http: //gypsy. fsl. wvnn 4 'g .... ..... ‘html 3/22/98 The gypsy moth was accidentally introduced to the United States in 1868 or 1869 by French artist and amateur scientist Etienne Leopold Trouvelot (1827-1895). after he brought specimens to Massachusetts in 1868 for cross-breeding purposes. His unsuccessful experiment was an attempt to develop a new strain of ‘silkworm.’ He is reported to have realized the significance of the initial release from his home in Medford. near Boston. and advised authorities, but no action was taken at that time (USDA Forest Service, 1998a). Refer to figure 2.3 for an image of Trouvelot. s. “\- X '2‘. \\. \ \- \\\\.\.\\ R‘t" ‘. \xk \'.\\ “\ “\‘.\‘. \:::.I\'I.\\t& u.“ aims... ““- \ .\ \\ I. \ ‘- \K‘. x 4.\ \I. \ \\\1§ \. ‘Y‘ s. s \ ~ are. I, \;~% '- \ \\\ \ \\ z‘. \ \\ Ix; \ ‘24- “R i \ x \\\ \k\ \ \\\\ \\ \\.\\. \ A\\\ -.-.- “R -.k “shun. ,./ ' r \\ '- $ \ \- \).t'_\\ i \ Figure 2.3 Etienne Leopold Trouvelot. The individual responsible for introducing the gypsy moth to the North American continent. Source: USDA Forest Service, Northeastern Forest Experiment Station. E. Leopold Trouvelot, Perpetraior of Our Problem http'J/gypsy.fsl.wvnet.edu/gmoth/worldlworld.html 3/22/98 Courtesy of Mary Lea Shane Archives, Lick Observatory. University of California. Santa Cruz. The gypsy moth is considered to be the primary defoliator of hardwoods in the northeastem portion of the United States, where its distribution is widespread, as shown in figure 2.4. A map of the projected spread of gypsy moths over the next 25 years is presented in Figure 2.5, and this extension is expected to cover a wide band adjoining the outer boundary of the present distribution. A thorough study assessing United States forest susceptibility to gypsy moth defoliation was completed by Liebhold, Gottschalk and Mason (1998), and an excerpt from their summary‘document is included below. Over-all forest susceptibility was quantified using five measures: average basal area per acre of preferred species, proportion of stand basal area in preferred species, proportion of land area covered by susceptible stands (> 20% of basal area in preferred species), proportion of land area covered by highly susceptible stands (> 50% of basal area in preferred species), and proportion of land area covered by extremely susceptible stands (> 80% of basal area in preferred species) for each county. All three [sic] measures yielded maps that indicated similar distributions of susceptible forests over the cotenninous US. The areas with the highest concentration of susceptible forests were in the central and southem Appalachians, the Cumberland Plateau, the Ozark Mountains, and the northwestern lake states. Comparison of these maps with the know [sic] distribution of individual susceptible species indicates that oaks are the major component of susceptible forests in the Appalachian, Cumberland, and Ozark areas but quaking aspen is the major susceptible species in the northwestem lake states (Liebhold, Gottschalk and Mason, 1998). The ten top-rated species, In terms of susceptibility to gypsy moth defoliation, were determined to be (highest susceptibility first) white oak (Quercus alba), sweetgum (Liquidambar styreclflua), quaking aspen (Populus tremuloides), northern red oak (Quercus rubra). black oak (Quercus velutina). chestnut oak (Quercus prinus), post oak (Quercus stellata). water oak (Quercus nigra), paper birch (Betuia papyrr’fera) and southern red oak (Quercus falcata). These ten species have a combined total basal area of 7.714 billion square feet per acre in the 48 cotenninous states in the United States. A complete list of the top 20 species identified, as well as maps depicting the 10 distribution, in basal square feet per acre. for all of these are provided by Liebhold, Gottschalk and Mason (1998). In the same work. these authors concluded that, on a national level, 'the impacts of defoliation and costs of gypsy moth management are likely to increase in the future.‘ \ . G d eh". I... _ .- . setteifi’tfi“ “'e'?t‘ , v.5. r 2:33W" AF-anr ."' n‘.‘ .. figure 2.4 1990 Distribution of the gypsy moth In North America. Counties with gypsy moth presence are shaded. The black circle has a radius 01850 kilometers and is centered at the point where the gypsy moth was introduced to the United States in 1868 or 1869. This spreading has thus occurred over a period of approximately 122 years. Source: USDA Forest Service, Northeastern Forest Experiment Station Gypsy Moth Spread in North America http://gypsy.fsl.wvnet.edu/gmothlspread/spread.html 3/22/98 Figure 2.5 25-year projection of gypsy moth spreading. Source: USDA Forest Service. Northeastern Forest Experiment Station Gypsy Moth Spread in North America http://gypsy.fsl.wvnet.edu/gmothlspread/spread.html 3/22/98 The distribution of gypsy moths within the State of Michigan now includes every county in the Lower Peninsula and also portions of the Upper Peninsula. Joria, Aheam, and Connor (1991) state that there are approximately 2.02 million hectares (just under five million acres) of forested land in the State of Michigan that are susceptible to defoliation by the gypsy moth, though USDA forest inventory figures suggest the actual extent is somewhat larger. as discussed in section 3.1 of this report (Leatherberry, 1993). Oak and aspen are particularly affected, though over 600 species of trees are subject to leaf consumption. As of the late 1980’s, the projected extent of defoliation was expected to reach hundreds of thousands of hectares per year (Montgomery, 1988). Defoliation in Michigan has been inconsistent, both in the areal extent and the number of counties affected, as shown in Figures 2.6 and 2.7, respectively. The peak over the last 19 years occurred in 1992. There has been decreased extent since then, though there was a large increase in the number of counties affected in 1997 compared to 1996. Although recent Michigan defoliation acreage has been less than that of the 1992 peak, a new strain of gypsy moth was discovered in North Carolina that has the potential for dissemination throughout the Midwest, suggesting the need for enhanced detection capabilities and continued monitoring efforts. The 1993 study period was the third highest in both areal extent of defoliation and number of counties affected for the period of record from 1979 to 1997. 800000 709000.- 600000-- gm... 5400000- gawooo- 32.0.0... 100000- c- YODI' Figure 2.6 Annual acreage of Michigan gypsy moth defoliation. Data source: Sapio (1998b). Bearer Counties with Defoliation figure 2.7 Annual number of counties with Michigan gypsy moth defoliation. Data source: Sapio (1998b). Decreased activity levels in recent years are attributed to the viral disease Neucleopoiyhederosis virsus (NPV), to the introduced Japanese Entomophaga maimaiga fungus, to parasitoids which feed in the body of the various life stages of the gypsy moth, to predators that consume the gypsy moth, as well as to very cold winters that result in high mortality of the egg masses (lngeils, 1995; Michigan State University Extension, 1998; Sapio, 1998a). The effectiveness of the fungus is dependent upon the abundance of moisture and high humidity during early June, with high mortality occurring when these conditions exist (lngeils, 1998). Federal and State government gypsy moth suppression costs in Michigan over the period 1986 to 1995 totalled approximately US$18 million, with a combined (Federal and State) peak annual expense of US$36 million in 1993 (USDA Forest Service, 1998b). In addition, some communities initiated locally-funded control programs. Given the magnitude and impact of the problem, it is desirable to closely monitor the extent of defoliation. The MDNR has traditionally obtained information through the use of a color infrared (CIR) video monitoring system employed in air flights over the State. The typical method of mapping widespread occurrence is by sketching the affected areas onto topographic maps based on low altitude aerial reconnaissance (Ahem at al., 1991; Talerlco, Walker and Skratt 1978). Limitations of this approach were summarized by Joria, Ahearn and Conner (1991 ), and include the subjective and variable skill of interpreters, the changing condition of monitored areas during the length of time re- quired to obtain a complete data set over the millions of hectares involved, and the brief two to three week temporal window of peak defoliation during which field data must be acquired. This typically occurs during the first three weeks in June (Muchoney and Haack, 1994). The major source of interpretation error has been tree mortality caused by defoliation in previous years. 2.2 Satellite Imagery Satellite imagery has been utilized in an attempt to overcome the limitations mentioned in the preceding section concerning the traditional method of identifying widespread areas of leaf loss (Dottavic and WIIliams, 1983). Landsat Thematic Mapper (TM), Multispectral Scanner (M88), and Systeme Pour l'Observation de la Terra -1 (SPOT-1) scenes have enabled identification of defoliated regions, as summarized by Muchoney and Rack (1994). AVHRR data acquired by National Oceanic and Atmospheric Administration (NOAA) satellites have also been proposed for use In this application (Lusch and Sapio, 1994). AVHRR acquisition commenced with the launch of TlROS-N on October 13, 1978, and continues to the present with other operational satellites in the 15 numbered NOAA series (Figure 2.8). These systems are passive, in that, in the visible and near-infrared (NIR) bands, they rely on reflected solar energy for illumination of the target, rather than actively emitting and then receiving a pulse as in radar systems. Figure 2.8 NOAA Polar Orbiter TIROS-N. Precursor to the NOAA AVHRR satellite series. Source: NOAA http://psbsgi1.nesdis.noaa.gov:8080./NSORS/ML/nsorsz.html 3/16/97 Earth-orbiting sensors have been widely used in land cover studies because of the large surface areas that can be monitored at the same time, the ability to use several electromagnetic wavelength bands in classifications, and the ability offered to revisit locations for temporal comparisons. Sensors acquire information for a narrow range, or band, of wavelengths, rather than for one specific value. The general concept of spectral response, which is the fundamental issue concerning target separability, and information concerning AVHRR data, which are the type used in this analysis, are briefly reviewed in the next two subsections. 2.2.1 Spectral Responses Surface materials have differing spectral responses, or degrees of reflectivity, at different wavelengths, and it is this characteristic that enables multi-band satellite sensor data to be used to discriminate between features. For example, there is comparatively little difference between vegetation and sandy loam soil In the visible (blue. green and red) wavelengths, which are approximately between 0.4 and 0.7 micrometers. However, there is a broad separation between the percent reflectance curves for vegetation and sandy loam soil in the near-infrared (NIR) wavelengths between 0.7 and 1.3 micrometers. At these NIR wavelengths, vegetation has approximately 2.5 times the response and in excess of 60% reflectance. Consequently, these materials would be easily separable using a NIR wavelength band. Similany, different types of vegetation have more widely varying levels of reflectance in the near-infrared, while in the visible wavelengths the values are much less separable. The variation in average percent reflectance over the visible and NIR wavelengths for certain vegetation classes is distinctive. The reflectance separation between classes enables the discrimination be- tween various types of vegetation. For example, deciduous trees characteristically pro- vide almost twice the average percent reflectance of coniferous trees (Avery and Berlin, 1992). In this manner, reflectance information acquired by the AVHRR sensor was used in this research in an attempt to discriminate between defoliated and nondefoliated deciduous trees, since the spectral response is different for healthy versus damaged leaves, for leaves at various stages of senescence, and for leaf-on versus leaf-off conditions (Avery and Berlin, 1992). Note, however, that removal of the canopy foliage allows undergrowth I7 and soil to contribute more to the reflected signals captured by the satellite sensor. All incoming solar radiation is either absorbed, transmitted or reflected. At visible wavelengths (AVHRR channel 1 is red) where high absorptance is typical for vegetation, plant pigmentation controls the spectral response. At longer wavelengths in the near- infrared (AVHRR channel 2), absorptance is much reduced and transmittance and reflectance assume much greater and almost equal importance. The response from vegetation in the near-infrared is controlled by the internal leaf structures, with the palisade mesophyll of the dorsiventral structure serving as light ‘plpes’ to the inner air voids where refraction occurs prior to emittance. Younger leaves have fewer air voids and a corresponding lesser level of reflectance in the near-infrared than leaves of greater maturity. The compact leaf structure has fewer air voids and lesser levels of reflectance. Finally, it should be noted that the NiR reflectance from a canopy of leaves is greater than from just one leaf due to partial transmittance through leaves to others, and then reflectance from all (Lusch, 1993). 2.2.2 AVHRR Characteristics AVHRR data are widely used for environmental monitoring, their primary advantage being useful spectral bands with high temporal resolution and single scene coverage of large areas. The most significant shortcomings of this sensor are '(1) the lack of radiometric calibration coefficients, and (2) the high slant angle which distorts radio- metric readings at extreme off-nadir angles' (Ehrlich, Estes and Singh, 1994, p. 147). The lack of atmospheric corrections is another serious shortfall. 18 The AVHRR sensor acquires information for a 2,399 kilometer swath over the Earth’s surface with each pass, and every day 14.1 satellite orbits are completed around the Earth at the nominal 833 kilometer altitude. Data are acquired at a 1.1 kilometer ground resolution at nadir. A diagram of the swath coverage is depicted in Figure 2.9. The period of each orbit is 102 minutes, and there is a 255° distance between orbits. The orbit is inclined at 98.9° and there is a 155° scan angle from nadir (Ullesand and Kiefer, 1994). Full-resolution Local Area Coverage (LAC), and sampled Global Area Coverage (GAC) data at a nominal spatial resolution of 4 kilometers, are recorded onboard as acquired and then later transmitted to a ground station during overpass. Command and Data Acquisition (CDA) stations at Gilmore Creek, Alaska and at Wallops island, Virginia receive sensor data and use communications satellites to rebroadcast the information to the Suitland, Maryland facilities of the NOAA National Environmental Satellite, Data and lnforrnation Service (NESDIS; USGS EROS Data Center, 1998). AVHRR data are available for the period from 1979 to the present. Current imagery for Michigan is readily available through the NOAA CoastWatch program, as summarized by Leshkevich, Schwab and Muhr (1993). In contrast to other imagery, the AVHRR data were available at no cost for this study, a significant factor in this investigation and also worthy of consideration in the potential application. The specific products used in this research are discussed in section 3.2. 19 figure 2.9 2,399 kilometer swath of AVHRR coverage over North America. Source: USGS EROS Data Center. lrnage modified from httpzlledcwwwcr.usgs.gov/glis/graphicslguidelavhrr/figure1 .gif 3/19/98 Data products of the CoastWatch program are available for four windows covering the Great Lakes region. The imagery used in this study was from the Michigan-Huron window, a Local Area Coverage (LAC) which encompasses 40.76 to 46.73 degrees north latitude and 79.78 to 88.05 degrees west longitude, with a mid-latitude pixel size of 1.3 kilometers. 'Actual resolution is determined by dcose, where d is the spatial resolution at the equator and 6 is the Iatitude...level 1b data are mapped to a Mercator projection and resampled to a 512 by 512 pixel grid“ (Leshkevich, Schwab and Muhr, 20 1993, p. 372). A subset was utilized which more closely corresponded to the specific area of interest. The AVHRR sensors acquire information in five radiometric bands. For this work, data from channels 1 and 2, representing 0.58 to 0.68 micrometers in the visible red, and 0.725 to 1.0 micrometers in the reflected infrared, respectively, were combined by CRS in the following manner to yield the Normalized Difference Vegetation Index (NDVI): NDVI = W Band 2 (near IR) + Band 1 (visible red) NDVI is a commonly used indicator of both the presence as well as the condition of crops and natural vegetation, and values typically fall between approximately 0.1 to 0.6 for vegetation. Damaged leaves progress through characteristic changes in spectral re- sponse, and a defoliated tree provides a different response than does a healthy tree in a normal leaf-on state, as outlined in section 2.2.1. NDVI is thus considered an ap- propriate indicator of defoliation (Avery and Berlin, 1992). In addition, NDVI provides some compensation for variations in illumination, the slope of the ground surface, aspect, and other characteristics affecting the analysis (Llllesand and Kiefer, 1994). Goward et al. (1991) review AVHRR NDVI measurements and related considerations. Cihiar, St.-Laurent and Dyer (1991) found that AVHRR NDVI growing season trajectories offer useful insight on vegetation development. The use of ratios in classification has often provided better results than when the channel data were presented separately. since the relationship was provided to the classifier explicitly and thus this information did not have to be discovered during 21 processing (Bailey and Thompson, 1990; Hammerstrom, 1993b). Green and Cosentino (1996) also cite as benefits the reduction of topographic effects and the augmentation of spectral trends. As the biomass increases through leaf generation during a typical growing season, and as the maturing leaves develop more air cavities in the mesophyll, the NDVI ratio increases. Later in the season, chloroplast deterioration occurs and eventually the cell structure collapses. Visible leaf colors shift from green to yellow to red-brown, and NDVI decreases. By the end of the growing season, the NIR reflectance has decreased and the reflectance at red wavelengths has increased, thus reducing NDVI (Avery and Berlin, 1992; Jensen, 1996). The reduction in biomass at the peak defoliation stage, caused by gypsy moth feeding, was expected to result in a reduction in measured NDVI. It was anticipated that these changes would be evident in the observed spectral- temporal signature, and that discrimination of affected areas would thus be possible. As described in detail in Chapters 3 and 4, three time periods were selected at which spectral responses were measured and then employed in data classification. The 1,300-meter-square spatial resolution of AVHRR imagery in Michigan is profoundly inferior to either 30-meter Landsat TM data or 20-meter SPOT multispectral High Resolution Visible (HRV) scenes, but satellites canylng AVHRR sensors offer a daily revisit frequency for a given location. This is vastly superior to the 16-day revisit period available for Landsat or the variable 5-day repeat coverage cycle for SPOT sensors in the mid-latitudes. For Landsat, this time refers to acquisition by the same satellite, and more frequent revisits are possible using the second sensor platform. For SPOT, which 22 has an orbital revisit period of 26 days, the more frequent acquisition is possible using the pointable optics for off-nadir imaging. In addition to the increased temporal resolution, Lillesand and Kiefer (1994) note that the coarser spatial resolution of the AVHRR sensor yields an increased radiometric resolution due to the greater amount of total energy arriving at the detectors of the charge-coupled device (000), which results in higher signal levels and an increased ability to detect very small differences in incoming radiation. Roesch, Van Deusen and Zhu (1995) reported, however, that scale effects of the coarse AVHRR spatial resolution contributed to a bias in the estimation of percent forest cover when target features are substantially smaller than the pixel size. Small regions of all classes may go undetected, and smoothing of edges would occur. The high temporal resolution of AVHRR data provides a solution to the problem of ob- taining synoptic information over large areas with easily manageable data volumes. This enables maximum value compositing (MVC) techniques to be applied in order to generate cloud-free scenes which depict the entire study area at approximately the same time. This is necessary because there is only a 5 to 10 percent probability of any overflight of the State of Michigan having minimal (0 to 10 percent) cloud cover (Lusch, 1994). With MVC, a balance must be established between the conflicting demands of using a compositing period long enough to minimize cloud presence yet short enough to provide the temporal resolution required to isolate targeted changes in the landscape. It should be noted, however, that multitemporal AVHRR MVC may be sensitive to ”residual cloud 23 contamination, [variations in] preceding climatic events, temporal and spatial scales of analyses, and the composition and spatial organization of the study area“ (Allen, Bara and Walsh, 1994, p. 52). 2.3 Classification Techniques Remote sensing image classification represents an attempt to categorize surface features based on data acquired by sensors positioned at some distance above the surface. Land cover classification of remotely sensed data usually leaves inaccuracies in the final product. These errors can be quantified when a land cover map exists that is considered to be completely correct, however such reference (or ‘truth’) information is not always readily available. There may be several concurrent contributing factors to the error. Temporal variation in the characteristics of the targets, spatial variation in responses from a given target, atmospheric influences upon the radiafion received at the sensor, off-nadir viewing angles and sun-target-sensor geometry, variance in sensor radiometric response and inadequate or variable detector calibration can all affect the data quality and the resulting classification. Research continues to address the normalization of these factors which degrade the accuracy of the final classification (Lillesand and Kiefer, 1994). Classification mchnlques provide varying levels of ability to accommodate such variance in remotely sensed imagery. The commonly applied Gaussian ML classifier considers both the variance and covariance of the class spectral vectors when assigning an unknown pixel to a category, but a normal distribution of the training data is assumed to exist. Violation of this assumption may result in an unacceptably high level of mlsclassification (Lillesand and Kiefer, 1994; Paola and Schowengerdt, 1995a). Chen and Takagi (1993) investigated the use of an ANN to classify NOAA AVHRR data, and pointed out that it is unrealistic to expect traditional statistical classifiers to perform well when classes overlap and when information is distributed across many often very highly correlated bands along only a limited number of major components. Here at al. (1994, p. 102) note that 'the most severe limitation arises when conventional classifiers implicitly presume a particular input metric with which to measure likeness in the multidimensional space of the feature vector.’ Further, these authors indicate that the absence of these assumptions in ANN classifiers which, in training, adaptively determine 'a metric specific to the problem,‘ allows the method to avoid 'this prejudicing of the classification process” (Hara etal.,1994, p. 102). in a detailed comparison of BP ANN and ML classifiers in an urban application, Paola and Schowengerdt (1995a, p. 995) concluded that, due to the nonparametric nature of the ANN approach, it “is more robust to training site selection.” This is contrasted with the maximum likelihood method, which they observe to be “sensitive to the purity of the class signatures...and performs poorly if they are not pure” (Paola and Schowengerdt, 1995a, p. 995). They also note that the lack of a statistical assumption in the neural network allows it to produce more mutually exclusive class discriminant functions, resulting in more detail at the pixel level and possibly at subpixel levels. The maximum-likelihood classifier, because it depends on to [sic] the assumption of normal (and potentially high variance) distributions, can mask this fine detail in the classification (Paola and Schowengerdt, 1995a, p. 994). Brown, Lusch and Duda (1998, p. 248) confirmed this greater degree of ‘smoothing’ produced by the ML classifier, noting that the ANN technique was better able to identify the classes which were more difficult to discern, and that “ANNs can be much more sensitive to slight variations in the inputs.” Paola and Schowengerdt (1995a) further elaborate by stating that The neural network, because of its fully-interconnected nature, produces a fundamentally different type of classification. The warped appearance of the network class output functions is due to the competitive learning that takes place during backpropagation training. As one class is trained it suppresses the output value of all the other classes and, in this way, the classes compete with one another for ‘territory" in the feature space. Thus, the network produces “probability densities“ that have sharper inter-class transitions and do not overlap nearly as much as those of maximum likelihood (Paola and Schowengerdt,1995a, p. 989). Because of these reports from current researchers, and other similar observations which have appeared in the literature, this work employed the ANN classifier in addition to the conventional maximum likelihood approach. The intent in doing so was to determine whether the performance characteristics of either would yield greater classification accuracy, given the data described in Chapter 3 and the methods described in Chapter 4. Openshaw (1994) has argued that this is in fact the only valid jusfification for experimenting with what he refers to as ‘neuroclasslfication.’ The approach used in this research is consistent with the integrative role suggested by McKeown (1987) and Foody (1995) for the use of artificial intelligence techniques in combining a variety of remote sensing, GIS and other input information to produce a greater understanding of land-related issues. 26 2.3.1 Maximum Ukellhood The maximum likelihood classification method is a supervised approach, in that known per-pixel class information is used to develop a characteristic signature, based on the corresponding input band data, that is then used to classify unknown pixels. The technique is commme employed in remote sensing applications and has been thoroughly described in related literature (Dude and Hart, 1973; Jensen, 1996; Lillesand and Kiefer, 1994; Schowengerdt, 1983). Maximum likelihood is considered the statistically optimal classifer for such applications (Bischof, Schneider and Pinz, 1992; Wan, 1990). Both the variance and the covariance of the input band spectral information are utilized as the probability, or likelihood, of class membership is assessed for each unknown pixel to be classed and for each possible assignment category. Probability density functions are generated for each category and are used in determining which category has the maximum probability, or is most likely, to be correct for the input vector of each unknown pixel. Assignment is made to the category of highest probability, with the underlying assumption that a Gaussian, or normal, distribution of single-band class training data exists. During implementation, the minimum number of pixels required for training purposes varies from (ion) to (100-n), where n is the number of the spectral bands used as input variables (Clark Labs, 1995; Lillesand and Kiefer, 1994). Details ccnceming the maximum likelihood methods used in this research may be found in section 4.1.2. 27 2.3.2 Artificial Neural Networks Within the field commonly described as artificial intelligence, many neural network algo- rithms have been developed, many versions of software implementations are available and research interest persists (Sarle, 1998a). Comprehensive summaries of developments in the field have been prepared by several researchers in recent years (Hush and Home, 1993; Lippman, 1987; Rumelhart and McClelland, 1986). Introductory information on the implementation of ANNs has been provided by Hammerstrom (1993a and 1993b). ANNs have been successfully utilized in numerous classification applications involving remote sensing data (Paola and Schowengerdt, 1993; 1995b). The method with the widest reported application in image classification in recent years is backpropagation (BP), with several variants available. These methods employ many processing iterations to arrive at a solution. Feedforward BP ANNs offer the ability to perform supervised classification. They consist of a variable number of layered, weighted, and interconnected processing elements which accept input data, sum and transfer these data using pro-defined functions, compare output values with reference data, and readjust the weights on the connections based on a specified ieaming rule. The process is repeated until the network has been trained, by adjusting the weights, to classify training data such that the output meets a target error threshold. Variations of feedforward BP algorithms primarily involve 28 modification of the transfer functions and ieaming rules. A diagram depicting the ANN architecture used in this research is presented in Figure 2.10. The activation, or transfer, functions are utilized within the processing elements of the hidden layer to introduce nonlinearity into the ANN computations (Sarle, 1998b). One advantage of the method is that highly irregular spikes or blades in the error surface are less likely to occur, thus facilitating the desired descent to the global minimum error. Input Layer (3 variables) Hidden Layer (3 nodes) Output Layer (2 classes) [om Classification] Figure 2.10 Artificial neural network architecture employed In this research. Full interconnection exists between elements in adjacent layers. The input data and the desired classification output are used in determining the structure of a BP ANN. The input layer shown in figure 2.10 simply accepts the input data vectors, and is determined by the selection of variables considered cmcial to the task, with one ‘node’ or element for each variable. The internal stnicture must be 29 established by trial and error to optimize the classification accuracy (Paola and Schowengerdt, 1997). The input layer distributes the data to the ‘hidden’ or internal nodes where a weighted sum of the inputs is calculated, passed through a transfer function, and then distributed to the elements in the output layer. The output layer consists of one element for each of the desired categories in the final solution. Here, a weighted sum of the incoming values is computed, passed through a transfer function, and a result is generated. in supervised training, the actual result is compared with the desired result, and the derivatives of the error with respect to the weights are sent back (back-propagated) to the hidden nodes where the individual contribution to the error is calculated at the nodes, and weights are then changed in accordance with a pre- established criterion (Hammerstrom, 1993b). The values of the weights ‘leamed’ in this manner thus facilitate discrimination between classes. Network architecture optimization can be approached by either ‘pruning’ an overly large structure or by gradually building a minimal structure. One might visualize a bowl-shaped object in three dimensions to represent a highly simplified error surface which is to be traversed by gradient descent from maximum error somewhere near the upper lip of the bowl to minimum error at the base. The starting point is established by the weight initialization. The ‘learning rate’ would control the vertical descent from initial maximum error to the final minimum error, while the ‘momentum’ term would affect horizontal oscillatory movement at roughly similar magnitudes of error (vertical axis) along the surface of the bowl. To elaborate, the ieaming parameter in the BP approach “is analogous to the distance along the error surface traveled in a single epoch, so that the smaller the learning rate, 30 the smaller the changes in the weights of the network at each epoch“ (Skidmore et al., 1997, p. 508). One version of a feedforward BP ANN includes a momentum term that is “added to the learning rate to incorporate the previous changes in weight with the current direction of movement in the weight space. In other words, inclusion of the momentum term avoids wild swings across the error surface, while allowing the system to learn faster" (Skidmore etal., 1997, p. 508). Momentum controls the magnitude of weight change between epochs, and thus affects the speed of learning, but not the accuracy. An attempt is made during the processing routine to limit the number of iterations utilized such that there is a sufficient number for the network to adaptively ‘learn’ the discrimination functions through the weight update procedure, while at the same time retaining the ability to generalize on nontraining data. Because the learning rate is analogous to the incremental decrease in error, small values for this parameter will require a larger number of epochs, or iterations. in addition to ieaming rate and momentum, the rate of convergence to minimum error is also affected by the order in which the training vectors are presented, the complexity of the regions for which decision boundaries are generated and the network structure (Huang and Lippman, 1988). One of the key advantages of ANNs is their distribution-free nature. Prior knowledge about a class statistical distribution is not required. Further, the ANN automatically determines the relative reliability and weight associated with each data source input to the network (Benediktsson, Swain and Ersoy, 1990; Here at al., 1994). Hepner et al. (1990) found ANNs to be somewhat tolerant of noisy data and also of missing data. 31 Lippman (1987) noted that a multilayer system can generate arbitrarily complex decision boundaries and can discriminate overlapping classes. The degree of class boundary complexity. or the ability of the ANN to adapt to the n-dimensional distribution of input data, is limited by the number of processing elements, or nodes, In the network. A neural net classifier implemented by Huang and Lippman (1988) was found to be robust because it discounted outliers. Klimasauskas (1992, p. 51) noted that the effective com- plexity of BP ANNs varies depending on the complexity of the input data, and that 'the complexity of the functional relationships and the stability of the resulting solutions have given neural networks an edge in modeling dynamic systems.“ It has also been reported that an ANN approach was successful in training on only mixed pixels when attempting to classify woody vegetation in southern England (Foschi and Smith, 1997). Perhaps of greatest value, neural networks are better able than conventional classifiers to detect subtle pattems in the data being classed. Also, the greater flexibility afforded by their nonlinear nature, the ability to generalize and develop the required model, and the ability to accept various types of data without violating assumptions are positive characteristics (Skidmore er al., 1997). ANNs require many training samples (Benediktsson, Swain and Ersoy, 1990), however, and these must be of good quality (Skidmore at al. 1997). in addition, the number of training samples selected from each class establishes the a priori probabilities, which cannot be avoided (Decatur 1989; Foody etal. 1995b). The varying response to weight initialization can yield substantial differences in classifications, in parficular when a limited number of processing elements is specified in the hidden layer. The use of a greater number of nodes can provide more consistent results (Paola and Schowengerdt, 1997). While Skidmore er al. (1997) offer the opinion that BP ANNs will not become the dominant method used in remote sensing 32 and GIS classification due to the limitah'ons, they suggest that a hybrid approach Involving expert systems and rule-based techniques offers promise. While ML classification is considered the statistically optimal approach (Bischof, Schneider and Pinz, 1992; Wan, 1990), ANNs have nonetheless outperformed the ML method. Bischof, Schneider and Pinz (1992) attribute this to the often only minor devia- tion of the image spectral information from the normal distribution required by ML theory. Paola and Schowengerdt (1995a) concur, and add that the competitive nature of ANN training, combined with the ability to consider the training data for all classes while generating class boundaries, results in more precisely defined boundaries with less overlap than provided by ML. The neural network, because of its fully-interconnected nature, produces a fundamentally different type of classification. The warped appearance of the network class output functions is due to the competitive learning that takes place during backpropagation training. As one class is trained it suppresses the output value of all the other classes, and, in this way, the classes compete with one another for “territory' in the feature space. Thus, the network produces “probability densities” that have sharper inter-class transitions and do not overlap nearly as much as those of maximum-likelihood. These “densities” are warped around one another and fit together in a jigsaw fashion to form the final decision regions of the classifier (Paola and Schowengerdt, 1995a, p. 989). Paola and Schowengerdt (1995a) suggest that the ANN makes use of the training data with greater efficiency than the ML classifier. The maximum-likelihood classifier models each class by its own statistics, independent of all other classes. The neural network, on the other hand, considers all class training data when defining the “probability density“ for each class. This is accomplished through the competitive training property discussed previously. When all the other classes are being trained, the output for a given class is set low (e.g., 0.1), thus, in effect, providing information about where that class does not exist in the feature space. This is used in addition to the training data for the class itself, which indicates where the class does exist. Thus the network is using the training data more efficiently than maximum-likelihood to 33 determine the natural distributions of the classes and is more robust when faced with a decrease in training samples (Paola and Schowengerdt, 1995a, p. 991 ). Previous work with artificial neural networks has been adversely affected by a lack of reliable procedures for developing optimum network architectures, training data and sampling considerations, difficulty in reaching the global minima of the error curve with gradient descent, variable output depending on weight initialization, and slow processing speeds. Recent research such as that by Foody, McCullough and Yates (1995a; 1995b), Blamire (1996) and, In particular, Paola and Schowengerdt (1993, 1995a, 1995b, 1997) addresses these issues. Genetic algorithms may be used for neural network parameter optimization (Murray, 1994). Details concerning the specific BP ANN methods employed in this work are presented in section 4.1.3. Chapter 3 DATA 3.0 Data The digital data used in this work consisted of AVHRR scenes, raster GIS land cover information, and reference defoliation polygons for the State of Michigan. Those portions of the dataset within the Lower Peninsula and within the Roscommon County political boundaries were used for classification. The direct source of all of this information was the Center For Remote Sensing at Michigan State University, where earlier work had been conducted using the data. For the 1993 growing season, multidate AVHRR NDVI composites consisting of two- band ratios for each of three separate narrow temporal windows, as well as GIS data on susceptibility to defoliation, were used in classification. Areas of observed defoliation, hand-sketched by MDNR from airborne video imagery and later digitized, constituted the reference (‘truth’) data, and was used for comparison against the classification output. During the 1993 study period, forest defoliation caused by gypsy moths in Michigan was primarily observed in the northern Lower Peninsula. The year 1993 was selected for the work reported here because during that time period the extent of defoliation in Michigan was among the highest levels for the 19 years included in the period of record (1979 through 1997), thus enabling larger sample sizes for training purposes. The period of study was thus representative of one of the times of 35 greatest concern to land managers in Michigan, but was not the somewhat anomalous peak year. The year 1993 immediately followed a season of much greater defoliation, and preceded several years of lower extent. In addition to these considerations, the reference information was already available in digital form for 1993, and selecting this period minimized the amount of preliminary data preparation required before classification tests could be conducted. The three temporal windows isolated from the AVHRR data available for the year 1993 were chosen in an attempt to coincide with, firstly, the period of peak defoliation, and then for times during the growing season immediately prior to and just after this period. it was anticipated that the spectral response would differ depending on the varying abundance and condition of deciduous tree leaves present at these three stages. As noted by Muchoney and Rack (1994), the time of peak defoliation typically occurs from June 1-21, but there is some latitude possible for later satellite observations. The composite AVHRR data selected by MDNR/CR8 to correspond to the period of peak defoliation was acquired in mid-July, somewhat later than the actual peak. Because the coarse resolution of the AVHRR sensor resulted in many pixels having both defoliation and nondefoliation content, it was anticipated that a higher proportion of purer (more thoroughly defoliated) pixels could be obtained for this exploratory investigation by using a subset of the Michigan data. Roscommon County was selected, in addition to the Lower Peninsula scenes, for the classification experiments conducted in this work because the area within his political boundaries of the county included a large portion of one of the most dense concentrations of defoliated forest in the State. It was expected that this region would thus yield many of the purer pixels necessary for 36 classifier training, and would also provide some less homogenous pixels for comparative analysis (Jensen, 1996). in addition, there were relatively few nonforest pixels in this locale which would be excluded from classification. 3.1 Study Area As just mentioned, classification was performed on both Lower Peninsula and Roscommon County data which was extracted from the complete Michigan data set. As highlighted in figure 3.1, Roscommon County lies in the north-central portion of the Lower Peninsula. General information regarding the topography, geomorphology, soils, geology, climate, and forest cover of the State will now be reviewed, with an emphasis on that region in the north-central Lower Peninsula observed to be defoliated during the year 1993, and in particular the Roscommon County test area. The primary area of investigation is thus in the northern portion of the Lower Peninsula, but observed defoliation extends also to the southern portion of Michigan and to a county in the Upper Peninsula. Michigan’s two peninsulas have a varied topography, ranging from the more rugged Keeweenaw highland in the western portion of the Upper Peninsula to the level lake. border plains adjacent to lakes Michigan, Huron, Superior, Erie, and St. Clair. Other physiographic regions represented within the State include plains, roiling plains, hill- lands, upland plains, and hilly uplands. Most landforms are the result of continental glaciation in the region, with the last glacial recession occurring approximately 9,500 years before the present. Roscommon County falls in the Houghton-Higgins upland 37 Figure 3.1 Roscommon County in the State of Michigan. Classification trials used Lower Peninsula and Roscommon County data. plain and the West Branch hill-land found in the north-central Lower Peninsula, in a landscape ecosystem identified by Albert (1995) as the Grayling outwash plain. Landforms in this county include moraines, till plains, and outwash plains. Soil associations in Roscommon County are predominantly spodosols and histosols. Spodosols are commonly gray colored soils in which leaching has migrated minerals and humus from the upper A horizon to the lower B horizon. Histosols have a high organic content, have undecayed vegetation present, and are typically found in wet areas (Sommers, 1977; Sommers et al., 1984). There are two large lakes in west- central and northwest Roscommon County, named Houghton and Higgins, with areas of 38 8,113 hectares (20,032 acres) and 3,888 hectares (9.600 acres), respectively (Sommers et al., 1984). Elevations In the State range from 570 feet abover‘nean sea level along Lake Erie in the southeastern Lower Peninsula to 1,980 feet in Baraga County in the western Upper Peninsula. Elevations in Roscommon County range frbrn approximately 1,000 to 1,400 feet above mean sea level. The youngest bedrock is in the center of the Lower Peninsula, with older fonnatlons radiating out from this area. The youngest formations present in the center of the Lower Peninsula are Jurassic of the Mesozoic Era. The oldest underlying bedrock in the State includes Lower, Middle and Upper Precambrian in the western part of the Upper Peninsula. Predominantly Mississippian and Pennsylvanian rocks of the Mesozoic Era, which ranges from approximately 290 to 360 million years before the present, underlie Roscommon County. These sedimentary rocks are overlain by 200 to 800 feet of glacial drift (Sommers, 1977; Sommers at al., 1984). Michigan, located in the middle latitudes and having prevailing air movement from the west to the east, has a continental climate. Mean annual temperatures throughout the state range from less than 40 to in excess of 48 degrees Fahrenheit, with warmer readings in the south and along Lake Michigan, and the coldest temperatures in the western Upper Peninsula. The hills and valleys present the opportunity for relatively colder localized nighttime temperatures in surface depressions. The annual precipitation ranges from less than 28 to more than 36 inches. Highest rainfall occurs in the southwest portion of the Lower Peninsula and in the northwest and north-central areas of the Upper Peninsula (Sommers, 1977). 39 Average monme temperature, heating, cooling, precipitation and snowfall information is available for the period from 1961 through 1990 for the Houghton Lake Station located in west-central Roscommon County. The NOAA National Climatic Data Center (NCDC) Midwestern Climate Center (MCC) identification number for this station is 203932. The satellite data used in this research, described in detail in section 3.2, covered the periods mid-May to mid-June, mid-July, and late August. The 1961-1990 average monthly high and average monthly low Fahrenheit temperatures for these three periods, as indicated Figure 3.2, are approximately 70/45, 82/54, and 75/50 degrees, respectively. I‘DUGIfl‘IllLLAKEfiflSW, MI (203932) flown“: 181-199 Extrema: 1940-195 ,17 :r ij _Q. ‘ .. 1 ‘ _ T r , ft . .._-,~ .- . 0;“ 83:88:38.: r F—f 1 ° '* I l Trimaran-‘0 (F) l \ S \ l ’1‘ / 1. # -—_____r. If”- -' O .. - - .. -1. . -2. i , .3. r g .. Jan Han flay Jul Sep Nov Fob fipr Jun Flog Oct Dec North of Your - " ‘ ~ -* Jfi'r‘ ' .. ._ -- ‘ - -, fliduskrn I: Record High “Haw High—~80. Lou 2-- Record Lou] at...“ . ...... . JV. .. ., . - . 7 7 fl - ~. mm Figure 3.2 Monthly temperature profile for Houghton Lake, Michigan. Source: Midwestern Climate Center, Champaign, Illinois, a cooperative program of the NOAA NCDC and the Illinois State Water Survey Division of the lllinois DNR. htth/mccsws.uiuc.edu/Summary/Gif/203932.TempM.gif 3/23/98 40 Average monthly heating degree days at Houghton Lake, summarized in Figure 3.3, range from above 1400 in January to a low of about 50 in July. Average cooling degree days are graphically presented in Figure 3.4, and occur from April through October, with a peak in July. Figure 3.5 shows that the average monthly maximum precipitation at Houghton Lake ranges from a low of just over one inch in February to a high of over 3.5 inches in September. The months from April through November have higher average monthly precipitation levels than other months. Snowfall occurs at the Houghton Lake station from October through April, with an average monthly peak of over 12 inches in January and a low of less than one inch in October. Figure 3.6 depicts this distribution. HOUGHI‘0N_LAKE_6_HSW, MI (203932) averages: 131-1990 Figure 3.3 Monthly heating degree day profile for Houghton Lake, Michigan. Source: Midwestern Climate Center, Champaign, Illinois, a cooperative program of the NOAA NCDC and the Illinois State Water Survey Division of the lllinois DNR. hflp llmm‘ sws nim‘ J ,.'3if/203932. HddM. gif 3/23/98 41 iniKi‘flUUAKEfiJSH, ”1 (203932) Mega: 131-18!) Figure 3.4 Monthly cooling degree day profile for Houghton Lake, Michigan. Source: Midwestern Climate Center, Champaign, Illinois, a cooperative program of the NOAA NCDC and the lllinois State Water Survey Division of the Illinois DNFi. http-Ilm ewe nine ‘ "‘ ,.'3if/203932.CddM.gif 3/23/98 I'DUGI‘II'W_LAKE_6_WSW, HI (203932) nuances: 1361-1990 Precipitatim (in. ) Minnow cum Center Figure 3.5 Monthly precipitation profile for Houghton Lake, Michigan. Source: Midwestern Climate Center. Champaign, Illinois, a cooperative program of the NOAA NCDC and the lllinois State Water Survey Division of the Illinois DNR. http Ilmm «we nim‘ ‘ -_ ,,'3if/203932. PrecM. gif 3/23/98 42 I'DWJAKLSJSW, ”I (203932) Moses: 1331-1390 Ii the: torn Clinch Center Source: Midwestern Climate Center, Champaign, Illinois, a cooperative program of the NOAA NCDC and the lllinois State Water Survey Division of the Illinois DNR. http'llmm sws uirm J "‘ ,.'3if/203932.Sn0wM.gif 3/25/98 Michigan is in a forested region broadly characterized by Barbour and Billings (1988) as containing mixed deciduous species, yet there is considerable compositional variety present. They show forest associations in the southern portion of the Lower Peninsula to be Beech-Maple, with the remainder of the State falling in their Conifer-Deciduous association class. The climate found throughout Michigan favors forest growth, yet there is a north-south and a more subtle east-west transition in typical species, as noted by Harman (1984): Moving south to north through the Lower Peninsula, one notices an increase in the number of evergreen species, particularly in the vicinity of Clare. These species (white spruce, balsam fir, jack pine) are boreal in distribution but extend south into northern Michigan on upland sites where they are usually only minor associates of other native hardwoods (though jack pine occurs extensively on sandy uplands). Apparently sensitive to warm, dry summers and neutral or basic soils, and unable to compete with more temperate-climate associates, they are limited primarily to the 43 northern Lower Peninsula - north of a line from Bay City to Muskegon (the tension zone) - and to the Upper Peninsula, where cooler summers and lower evaporation are typical along with the acidic, sandy soils on which these trees are most competitive. Warmer summer weather with higher evaporation rates combined with lea favorable soils seem to restrict their occurrence in southern Michigan (Harman, 1984. p.87). The Roscommon County study area in the north-central Lower Peninsula is north of the ‘tension zone’ referred to by Harman, and is also north of Clare. The north-south species transition in the Lower Peninsula is noted by Harman to be the apparent result of climatic influence, but the narrowness of the transition zone may be related to the soils distribution. In the northern region of the Lower Peninsula, coarse, sandy soils are more prevalent, whereas to the south of this tension zone he notes heavier glacial till soils are more prevalent. Harman goes on to state that A less conspicuous forest gradient exists from the central to the western portions of the Lower Peninsula. Woodland composition in the central Lower Peninsula is a mosaic of beech-maple mixed with some oak-hickory communities that results from the patchwork of different soils. Westward through the state, however, the more drought-sensitive beech-sugar maple forests become more common, and north of about Muskegon, they are mixed increasingly with yellow birch and hemlock. ...This westward increase of beech, sugar maple, and other drought-sensitive associates may be related to climatic modifications by Lake Michigan (Harman 1984, p.87). 1993 data for the northern Lower Peninsula show that 52% of the timberland was privately owned by individuals and farmers (Figure 3.7). Of the balance, 40% was publicly held, and 8% was owned by corporations (Leatherberry, 1993). Other Miscellaneous public corporations 1 it 8% Individuals &farmers 52% National Forest 1 2S Figure 3.7 1993 timberland ownership in the Northern Lower Peninsula Unit. Source: MDNR Forest Management Division. http://www.dnr.state.mi.us/www/lmd/fia/nlp/fianlp.html 3/23/98 In the Roscommon County area for which classification experimentation was conducted, presettlement forests were primarily Pine-Oak (Sommers etal., 1984). In the northern Lower Pensinsula, "The deciduous members of the oak-hickory species become fewer, and upland pin oak may be the dominant species. Dry sites that have been heavily disturbed are dominated by red maple, bigtooth and quaking aspen, and some pines (both red and jack), but recurrent fires eradicated the pines on some sites. Extensive jack-pine plains characterize sandy sites that have been burned less often or those sites that are a bit less xeric (excessively dry)” (Herman, 1984, p.88). In 1993, the fifth Michigan forest inventory was completed. This wodr was identified by the USDA as the 1993 Michigan Forest Inventory and Analysis (FIA), and details were included for the Northern Lower Peninsula Unit (Leatherberry, 1993) which encompasses the majority of all observed defoliation. The Northern Lower Peninsula 45 Unit includes those Michigan counties on the tier from Oceana through Bay and all other counties to the north of this tier in the Lower Peninsula. The inventory was carried out under the responsibility of the North Central Forest Experiment Station of the USDA Forest Service. Figure 3.8 indicates that in the Northem Lower Peninsula Unit where gypsy moth defoliation was concentrated, Maple- Birch, Aspen, and Oak-Hickory were the most abundant forest types, with a combined total areal extent in that year of approximately 4.7 million acres, and the addition of Elm- Ash-Soft Maple and other hardwoods increases this forested total to approximately 5.5 million acres. This 5.5 million acres represents 73.50% of all forested land in the Unit. Other forest types specifically mentioned in a USDA summary include Jack Pine, Red Pine, and other softwoods, as shown in Figure 3.8. The inventory indicated that 65.36% of the 11.3451 million acres of land in the USDA Northem Lower Peninsula Unit were forested when the 1993 Michigan FlA was conducted (Leatherberry, 1993). 46 01h" 80300068 Red pine E} 1993 Jack pine I “30 g Other hardwoods ! EIm-ssh-soit maple o it OCR-NONI! ”'5‘“‘3'1'3‘3:I'2-2-:-:-:-:- eeeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeeee Aspen eeeeeeeeeeeeeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeeeeeeeeeeeeee eeeeeeeeeeeeeeeeeeeeeeeeeeee Maple-birch o (is {a {.5 5.0 is flea (rt-on acres) Figure3.8 1980and1993arealextsntofmalorforssttypsslnthsuorthernLower Peninsula Unit. Source: MDNR Forest Management Division. http://www.dnr.state.mi.us/www/fmd/fia/nlplfianlp.html 3/23/98 3.2 AVHRR imagery The most recent two months of AVHRR data for the study area at the time of collection by CBS were available through the NOAA Great Lakes Environmental Research Labo- ratory (GLERL), located in Ann Arbor, Michigan. Archived data were available through the NOAA National Oceanographic Data Center (NODC). The activities of the current NOAA CoastWatch Active Access System (NCAAS) include archiving and distributing CoastWatch AVHRR data. Detailed information concerning these facilities and data is available via the World Wide Web as indicated in Appendix A. 47 As part of previous work at the CBS, and as described by Liebens (1994), ten 1993 AVHFiFl scenes were downloaded from NOAA in Ann Arbor (then-recent dates) and Washington (then-older dates). Selection was based on completeness of geographic coverage and degree of cloud cover. The images were selected with the goal of providing cloud-free composites for three times: pre-defoliation, concurrent with peak defoliation, and post defoliation. The pre-defollation composite image approximately covered the period mid-May to mid-June, and consisted of the following five julian dates: 136, 146, 157, 162, and 163. The peak defoliation composite was produced from scenes on julian dates 193, 196, and 197, which corresponded to mid-July. The post defoliation time composite for late August consisted of scenes 243 and 244. image acquisition time was 1300 hours for julian date 157, 2100 hours for julian date 193, and 2000 hours for all other dates, producing sun-angle differences for julian date 157 and 193 compared with the other eight scenes. All images were geometrically registered to scene 157 using a linear method. Registration error was stated to be on 'the order of 0.1 - 0.2 pixels' (Liebens, 1994). Error in the range from 0.5 to 1.0 pixels or less is normally considered acceptable (T ownshend at al., 1992). The NDVI was calculated for these images, and rescaled to values of 0 through 255. The cloud cover was removed through maximum value compositing, in which the highest pixel values of the scenes in each narrow temporal window were used to generate the final composite. This elimi- nated the clouds, which had very low NDVI values. Although not specifically stated in the CRS literature regarding this topic, it appears as though both NOAA-10 and NOAA-11 satellites were used, based on the overpass times just mentioned. NOAA-10 was launched September 17, 1986 and remains in service. 48 NOAA-11 was launched September 24, 1988 and went out of service on September 13, 1994 (USGS EROS Data Center, 1998). Both were classed as third—generation operational meteorological satellites. Uncorrected, die combined usage of NOAA-10 and NOAA-11 data would introduce variations due to differing sensor response characteristics. Also, the difference in overpass times results in different sun angles, which in turn contribute to differences in scene brightness. The different satellite orbital alignments introduce varying ground resolution (coarser farther off-nadir) and varying lengths of atmospheric travel paths. Figure 3.9, Figure 3.10 and Figure 3.11 provide a visual impression of the Lower Peninsula pre-, peak and post defoliation NDVI values, respectively. Though the three images appear somewhat similar, generally moderate values are observed in the earlier pre-defoliation stage, while higher NDVI ratios existed later in the growing season during the period of peak defoliation, with slightly declining values at post defoliation. This relationship is evident at a greater level of detail in the Roscommon County subsets presented as Figure 3.12, Figure 3.13 and Figure 3.14 for pre-, peak and post defoliation phases, respectively. Summary statistics for the defoliated and nondefoliated pixels at the two regional levels are presented later in section 4.1.2. 49 Figure 3.9 Lower Peninsula pre-defollatlon NDVI. Dark green is high, yellow is moderate, and dark red is low. Mid-May to mid-June, 1993. u: 1...... “M '“L-L... ..4 ._.A¢Lu Figure 3.10 Lower Peninsula peak defoliation NDVI. Dark green is high, yellow is moderate, and dark red is low. Mid-July, 1993. 50 Figure 3.11 Lower Peninsula post defoliation NDVI. Dark green is high, yellow is moderate, and dark red is low. Late August, 1993. Figure 3.12 Roscommon County pro-defoliation NDVI. Dark green is high. yellow is mid-range, dark red is low. Mid-May to mid-June. 1993. 51 Figure 3.13 Roscommon County peak defoliation NDVI. Dark green is high, yellow is mid-range, dark red is low. Mid-July, 1993. Figure 3.14 Roscommon County post defoliation. NDVI. Dark green is high, yellow is mid-range, dark red is low. Late August, 1993. 52 3.3 Land Cover Since the forest defoliation which is to be detected must occur in forested areas, the classification procedure would benefit from the knowledge of whether the area represented by a pixel is or is not forest land cover. More specifically, it would be beneficial to exclude from consideration those pixels in nonforested portions of the study The CBS at Michigan State University compiled a land cover map of the state based on the visual interpretation of custom-enhanced Landsat imagery. Both M88 and TM data from the period 1979 to 1986 were interpreted. These scenes were optically enlarged to match stable-base copies of USGS 19 x 29 quadrangies (1 :250,000-scale). A 40 hectare minimum mapping size was employed. The hand-drawn land cover polygons were digitized and then rasterlzed at a cell size of 333.33 meters. Subsequently, these data were aggregated using a 3 x 3 model filter to the final one kilometer x one kilometer cell size using custom software which resolved ties by deferring to the lower numbered of the two land cover categories (i.e. 1-Urban would take precedence over 2-Agriculture, et cetera). In previous work by staff of the CRS, this map was registered to the AVHRR scenes and forested portions of the region were isolated and used to create a binary image, shown in Figure 3.15. In preprocessing operations within a geographic information system, this binary image was used to exclude all nonforested pixels from the data later presented for classification. 53 Figure 3.15 Lower Peninsula forest mask. Forested areas are indicated in black. 3.4 Defoliated Areas Reference information, consisting of observed areas of gypsy moth defoliation in 34 Michigan counties, was acquired by MDNR in 1993 using an airbome video system. Defoliated areas were sketched onto 1:100,000 topographic maps by MDNR following interpretation. These defoliated areas were later digitized for each individual county and were available for this study as vector CMAP files. The most seriously affected portion of Michigan is the northern half of the Lower Peninsula, as shown in Figure 3.16, which depicts counties affected during the study period. Due to a very small occurrence on Bois Blanc Island, Mackinac County in the Upper Peninsula is indicated as a defoliated area in Figure 3.16. However, due to the small size of the affected area, this was lost in the vector to raster conversion. There were no occurrences noted for other portions of the Upper Peninsula in the ‘truth’ data provided by MDNR. Thus, all known defoliated areas in the raster data, which served as the basis for processing operations, were in the Lower Peninsula of Michigan. Small areas were defoliated in Oakland and St.Clair counties in southeastern Michigan. The extent along Lake Michigan was as far south as Muskegon County, but no defoliation was noted in Antrim, Charlevoix and Leelenau counties, which are located in the northwestern Lower Peninsula. Defoliation presence was expressed in two ranges: 75 to 100% and 50 to 75%, with digital information identified to indicate either of these two states, or the absence of defoliation. As noted by Congalton (1991), the accuracy of the ‘truth’ data is often not known with certainty. Such was the case in this investigation, although due care and standard professional practices were exercised in the acquisition, interpretation, digitizing and preprocessing of the information. As is customary, it was assumed that the reference data were correct. However, due to the time lapse between acquisition and use, field verification was not possible in connection with this research. Recall that defoliation totals may be subject to misinterpreted mortality from previous years. 55 Figure 3.16 Michigan Counties With Observed Defoliation in 1993. Based on MDNR interpretations of airborne video. Black areas indicate counties where defoliation was observed. The Roscommon County classification test area, which also experienced defoliation, is shown in white. Nondefoliated counties are depicted as green. A statistical summary generated by CMAP software for the statewide CMAP vector files indicated that in this format there were a total of 41 1,763.2 acres within defoliation polygons, with a mean area of 349.54 acres, yet with a standard deviation of 862.02 acres. At full conversion, this total acreage would fill 986 of the 1.3 kilometer-square raster cells, given 417.6 acres per cell. The largest of the 1,178 polygons was 9,774.6 acres in size, which would cover 23.41 complete 1.3 kilometer-square AVHRR pixels. Twenty-two ’slivers' less than 0.3 acre in size were deleted in preprocessing, so the smallest vector defoliation polygon was 0.3 acres. in the original CMAP vector files, 56 areas interpreted to be 75 to 100% defoliated were classed as 1, and areas interpreted to be 50 to 75% defoliated were classed as 2. There were a total of 1,942 of the 1.3 kilometer-square raster cells classed as defoliated in Michigan, which indicates 328,198 hectares (810,995 acres) defoliated. This exceeds the amount indicated by the vector ’truth’ data. Additional details ccnceming the difference in vector and raster defoliation areas are provided in section 4.1.1.2. 57 Chapter 4 METHODS 4.0 Methods Data preprocessing involved several steps utilizing a number of software packages to prepare the digital files for input to classifiers. Following classification, common measures of remote sensing accuracy assessment were calculated and compared. The hardware and software utilized for this work are commercially available or are freeware, and were provided by the Department of Geography and the Center for Remote Sensing and GIS at Michigan State University, and by the Institute for Parallel and Distributed High Performance Systems at the University of Stuttgart. This research employed Sun Ultra 1 and Ultra 2 workstations and Pentium and 486-class personal computers to run CMAP, lDRiSl, EXCEL, SYSTAT, ERDAS, PCl EASl/PACE. FOXPRO, ARC/lNFO and SNNS software for the various tasks. The general processing requirements were to merge previously digitized county vector reference information into a statewide coverage, convert this to raster format, coregister the result with three precomposited raster AVHRR NDVI files, sample the data, and generate Lower Peninsula and single-county classifications of the data by the use of a convenfional maximum likelihood classifier and also by employing a less conventional artificial neural network approach. Accuracy was then assessed by generating an error, or confusion, matrix, calculating percent correctly classified (FCC) and Kappa values, and comparing these for the different classification methods and spatial extent. Two measures of signature separability were used to determine the statistical similarity 58 between the defoliated and nondefoliated classifier training information. Finally, Kappa values were compared with results achieved by other researchers. 4.1 Data Processing Processing is logically divided into two phases, the preprocessing work and the actual classification of the properly formatted, assembled, and sampled data. Michigan Lower Peninsula data were initially prepared, and a Roscommon County subset was then selected, or windowed, from these for additional analysis. 4.1.1 Preprocessing The preprocessing operations converted both the raster and vector data into a form suitable for classification, with varying amounts of preliminary work required on the digital files received. These files included the AVHRR imagery, land cover information and boundary files, and a map showing known areas of defoliation occurrence during 1993, the period under investigation. 4.1.1.1 Raster AVHRR imagery and Land Cover Data Imagery was received from CBS as raster files in lDRlSl format containing NDVI values, scaled from 0 through 255, as the pixel information content. Three temporally composited scenes were provided, one each for the pro-defoliation, the peak defoliation, and the post-defoliation stages. As detailed in section 3.2, these composite periods correspond to mid-May to mid-June, mid-July and end-August, respectively. The method 59 of calculating these NDVI values was reviewed in section 2.2.2. The forest mask had been applied to all of these scenes, thus excluding nonforefied areas from consideration. The forested portion was approximately 38% of the entire Lower Peninsula of Michigan, at 1.3 kilometer-square resolution. Using six corresponding points from the AVHRR scenes and from a Michigan state and county boundary file, which properly overlayed the defoliation file, the three AVHRR scenes were resampled to the 0.325 kilometer-square preliminary spatial resolution of the reference data, and later aggregated to 1.3 kilometer-square resolution. The reference information was in the Michigan State Plane (Zone 2) coordinate system. A linear (first order) mapping function was used to minimize the adverse effect of any inadvertent poor choices in points and due to the limited number of readily identifiable control points on the data received. Nearest neighbor resampling was performed, thus placing the NDVI value of the closest, or nearest, cell into the new output cell. This approach does not alter the data values, but only moves them to a new location. It is the preferred method for resampling. Following standard procedures for this type of operation, those correspondence points having the highest residuals were omitted, while retaining a sufficient minimum number distributed around the defoliated area. A distribution of six points encircling the primary region of defoliation remained from a total of 19 points identified initially, as depicted in Figure 4.1. The residuals for these six individual points ranged from a low of 0.20 to a high of 0.94 pixels. Six points are twice the mathematical minimum requirement. The conventional level of acceptable root mean square error (RMSE) for such an imagery resampling procedure is one-half pixel (T ownshend et al., 1992). The final overall RMSE, using only these six points and expressed in input image units, was 0.58, which approximates the desired one—half pixel level. The first-order polynomial coefficients of the back transformation, from new cell to old cell, are provided in Table 4.1, and are used in the following equations: ’=bo+b,x+b,y y’=bo+b,x+b,y where x’ and y’ represent the original input, and x and y represent values in the rectified image, and on, b, and b2 are the coefficients listed in Table 4.1. Table4.1 Computed llneartransformation coefficientsofthebacktransformatlon Truncated. Values were carried internally in the ldrisi software module to 20 significant figures. Coefficient X Y b0 ‘ - 321 .68107582 14308523385 b1 0.00023916 0.00000417 b2 - 0.00000238 0.00023674 61 l_ l | | —- Figure 4.1 Control points used in resampling. Black control points enlarged for visibility. Areas of defoliation are red. 4.1.1.2 Defoliated Areas Vector files representing the actual extent of gypsy moth defoliation within each affected Michigan county, as well as a state and county boundary file, were received in CMAP format. In CMAP, these individual county files were merged to provide a single file providing comprehensive information. This file was edited, cleaned and built, and slivers less than 0.3 acres were deleted. This reference information was then exported from CMAP to lDRlSl vector format. This product was classed to distinguish between areas of 75-100% defoliation (class 1), areas of 50-75% defoliation (class 2), and areas of no observed defoliation. These regions were based on MDNR interpretations of airborne video reconnaissance acquired in 1993, and hand-sketched onto 1:100,000 topographic 62 maps, which served as the basis for digitizing. Figure 4.2 is an example showing a large defoliation occurrence in southeastern Roscommon County. in lDRlSI, the vector reference files were converted to raster format, with the final cell size chosen to coincide with the 1.3 kilometer-square mid-latitude pixel size of the NOAA CoastWatch Michigan-Huron AVHRR scene that was the source of the spectral information (Leshkevich, Schwab and Muhr, 1993). The vector-to-raster conversion first utilized cells of size 0.325 kilometer by 0.325 kilometer (to produce a total of 16 pixels per AVHRR cell), which were then aggregated to 1.3 kilometer-square cell size, with each cell having an averaged value in the range 0 to 1 to represent the degree of defoliation. Prior to aggregation, in order to specify the proportion a cell was defoliated, the cells identified as 75-100% defoliated in the original vector data were assigned the midpoint value of 0.875 as the proportion defoliated in the raster format. Similarly, the pixels identified as 50—75% defoliated in the CMAP data were assigned the midpoint value 0.625 as the proportion defoliated. Thus, no call in the aggregated raster data was identified as 100% defoliated, and the maximum defoliation content was 0.875, whereas the minimum was zero. This approach provided subpixel reference information that was used in sampling and later analysis. Since there are 169.0 hectares (417.6 acres) in a 1.3 by 1.3 kilometers cell, the area of 1/16 of such a cell is 10.56 hectares (26.1 acres). 63 Figure 4.2 Reference sketch map for Roscommon County defoliation cluster. Red is 75-100% and dark green is 50-75% defoliation. The southeast comer of Roscommon County is at lower right, with Gladwin County below and Ogemaw County to the right. Base map is USGS Houghton Lake 1:100,000 30 X 60 minute quadrangle. The onscreen information within the IDRISI software and a related technical reference manual both state that the ‘Polyras’ command used for the vector to raster conversion in this work functions as described in the following quotation: “Conversion is done by determining the intersection points of the polygon boundary and lines representing the centers of image rows. A boundary causes a cell to be included in the polygon if its position is more than halfway through that cell” (Eastman, 1992, p. 138-139). This procedure generated considerably more raster defoliation than was found in the original vector format. As the preceding quotation implies, the cause was apparently due to many cases where only partial polygon content within a raster cell forced the raster pixel to be classed as defoliated. The total area mapped with some defoliation in the original vector CMAP files was 166.6346 hectares (411.7632 acres), in contrast to the 328,198 hectares (810,996 acres) represented by the 1,942 raster cells at 1.3 kilometer-square spatial resolution which were identified as defoliated and used as reference information in classifications during this research. However, many 1.3 kilometer-square pixels had lower defoliation content than the vector polygons, which were estimated only at either 87.5% or 62.5%. 4.1.1.3 Sampling In order to develop characteristic NDVI signatures for the three temporal windows represented by the AVHRR composite images, samples of the defoliated and the nondefoliated cells were generated. The defoliated sample was based on within-pixel total defoliated area (defoliation being the primary target to be identified), and all pixels having a defoliation content greater than or equal to 50% were used for training. The need for training pixels possessing greater purity was discussed in Chapter 3. The nondefoliated sample consisted of a random selection of nondefoliated pixels falling near (up to 1.3 kilometers or one pixel away from) all defoliated pixels with gmater than 0% defoliation (not only those with greater than or equal to 50% defoliation). Nondefoliated pixels that were adjacent to 65 defoliated pixels were thus randomly selected. with an attempt made to provide a total number of nondefoliated cells approximately equal to the number of sampled defoliated pixels, in conformance with common practices (Blamire, 1996). This approach favors the ANN method in that the borders of the decision boundary are highlighted, in contrast to the standard approach in maximum likelihood training in which samples are taken from within the class region (Bishof, Schneider and Pinz, 1992; An and Chung, 1994). Congalton (1991) recommends a stratified random sampling approach when few samples are extracted from each class. The complete Lower Peninsula image depicting defoliated and nondefoliated cells is shown in Figure 4.3. There were 1,942 defoliated and 21,295 nondefoliated pixels. This image was sampled as described and the resulting product is depicted in Figure 4.4. In the sampled Lower Peninsula image that was presented to the classifiers for training, there were 372 defoliated pixels and 294 nondefoliated pixels. The Roscommon County analysis was based on data from the full and sampled Lower Peninsula data to reflect only the areas within the county. No separate resampling was done for Roscommon County. The Roscommon County complete and sampled images are presented in Figure 4.5. in the complete county image, there were 160 defoliated and 595 nondefoliated pixels. in the sampled Roscommon County image, there were 55 defoliated pixels and 23 nondefoliated pixels. Figure 4.5 may be compared with Figure 4.2, a reference sketch map of the southeastern portion of Roscommon County. An and Chung (1994) have reported that better results are not guaranteed by larger training sample sizes, citing Wann, Heidiger and Greenbaun (1990) as further corroboration. Lawrence (1991) stressed that examples should be presented in proportions that are fairly even. NDVI values corresponding to the sampled pixels were then extracted from composites coinciding with the three temporal defoliation stages and used to generate characteristic signatures which were later employed to classify other cells. Two techniques implemented in EASl/PACE were utilized in order to assess the degree of statistical similarity between these training signatures so as to discern whether it was reasonable to anticipate sharply defined spectral class partitioning by the classifiers. These signature separability measures were the Transformed Divergence (TD) and the Bhattacharrya or Jeffries-Matusita Distance (BD). While the BD method is considered to have a stronger theoretical basis, both measures produce similar real numbers between 0 and 2. Values in the range from 0 to 1.0 are considered to be very poorly separable, and values in the range from 1.0 to 1.9 are poorly separable. Good separability may be anticipated when the calculated result is in the range from 1.9 to 2.0. Lower numbers reflect a greater degree of statistical similarity, and may be indicative of either high within-class training site variance or improper spectral band combinations (Richards, 1986). Both the TD and the 80 measures were calculated for the Lower Peninsula and the Roscommon County defoliated and nondefoliated signatures. 67 Fl are 4.3 Complete Lower Peninsula reference data. Black areas are defoliated. Green areas are nondefoliated. White is masked. I‘l T I ll Fi ure 4.4 Sampled Lower Peninsula data. Black areas are defoliated. Green areas are nondefoliated. White is masked. - I I Complete Data Sampled Data Figure 4.5 Roscommon County complete and sampled reference data. Black areas are defoliated. Green areas are nondefoliated. White is masked. 4.1.1.4 Roscommon County Scene Extraction For classification by the maximum likelihood method and by the BP ANN technique, a subset of the full Lower Peninsula dataset was utilized. Because of the relatively coarse spatial resolution of the AVHRR sensor, most pixels were ‘mixed,’ or nonhomogenous. These cells with a mixture of both defoliation and nondefoliation content were anticipated to be difficult to discriminate, and consequently a region was selected through visual inspection of the digital files and the 1:100,000 topographic reference sketch maps where, due to more concentrated and extensive known defoliation, it was anticipated that there would be a greater proportion of ‘purer’ pixels than would be found overall in the Lower Peninsula. 69 This region is approximately enclosed within the political boundaries of Roscommon County, and was extracted by specifying the upper left and the lower right corner cell row and column numbers, based on the full Lower Peninsula scene. A vector map of the county boundaries for the State was placed over a raster NDVI scene and comer raster pixels for the region to be extracted were selected that fell under the intersection of county boundaries at the four vector county comers. Roscommon County scenes were extracted from all files used in classification. 4.1.2 Maximum Likelihood Classification Using data prepared as described in the preceding section, maximum likelihood classification was conducted using EASl/PACE commercial software (PCl Inc., Richmond Hill, Ontario, Canada) on Lower Peninsula information as well as on Roscommon County data. The classification technique was briefly summarized in section 2.3.1. Defoliated and nondefoliated training signatures were created for the statewide and county tests, with equal a pn'ori probabilities employed in those instances when a pixel to be classified was within overlapping portions of class signatures in the feature space. The three composited NDVI images for the pro-defoliation, peak defoliation, and post defoliation stages were the ‘bands’ used to provide the spectral information for the classification, with both a Lower Peninsula set and a Roscommon County set used in separate tests. The classifier was forced to assign all defoliated and nondefoliated pixels based on the Mahalanobis distance. with the exception of only those pixels within the 70 specified regions which were masked as nonsusceptible to defoliation. With the Mahalanobis measure of divergence, a lesser distance between a pixel vector and a class vector indicates greater probability of class membership, as reviewed by Jensen (1996). A file representing the most likely class of each pixel was generated following Summary NDVI statistics for the Lower Peninsula full and sampled defoliated and nondefoliated cells are presented in Table 4.2. Similar data for Roscommon County are included in Table 4.3. Note that the highest mean NDVI value for defoliated cells within each group occurs at peak defoliation in every case except for the Roscommon County defoliated cells, which had a high post-defoliation figure. The Roscommon County peak defoliation mean NDVI values for defoliated cells exceeded those for the Lower Peninsula. The signatures presented in Tables 4.2 and 4.3 reveal a deviation from the anticipated reduction in NDVI at peak defoliation in the defoliated pixels. This postulated reduction in NDVI was based on the controlling processes ouflined in sections 2.2.1 and 2.2.2. it was a matter of concern that defoliated and nondefoliated class discrimination based on such similar signatures would likely prove to be very difficult to accomplish. This situation may reflect the inability of the coarse spatial resolution to adequately detect insect-induced transformations in leaf state due to combined returns from multiple reflectors within a very large sensor ground cell. Rather than indicating presence or absence of defoliation, the resulting detector responses summarized in Tables 4.2 and 4.3 may instead primarily represent a broad measurement of typical seasonal growth and later incipient senescence at the spatial resolution of the satellite. Though a 71 thorough investigation of this anomaly is beyond the scope of the current endeavor, additional details regarding signature separability are provided in sections 2.2.1, 4.1.1.3, 5.1.1 and 5.1.2. Section 5.2 includes a discussion of the impact of mixed pixels in this research. Table 4.2 Lower Peninsula NDVI summary statistics Mean, standard deviation. CONDITION STAGE FULL SAMPLE Defoliated Pre-defoliation 135.8, 17.4 135.5. 17.0 Peak defoliation 147.5, 17.9 146.6, 16.4 Post defoliation 144.5,18.1 142.5, 21.6 Nondefoliabd Pre-defoliation 132.2, 20.6 133.0, 19.0 Peak defoliation 145.4, 21 .4 146.0, 22.1 Post defoliation 142.1 . 20.4 142.2. 22.2 Table 4.3 Roscommon County NDVI summary statistics Mean, standard deviation. CONDITION Defoliated Nondefoliated STAGE Pro—defoliation Peak defoliation Post defoliation Pre-defoliation Peak defoliation Post defoliation 72 FULL 128.9, 16.5 154.9. 13.4 153.4, 9.3 125.3. 1 3.7 149.0. 1 8.4 146.2. 1 7.1 SAMPLE 123.9, 16.7 1 52.3. 14.0 154.1 , 8.9 125.6, 10.8 152.4, 12.0 149.9. 8.8 4.1.3 Artificial Neural Network Classification The classification of imagery by artificial neural networks was reviewed in section 2.3.2. Two software packages were used which offered the ability to create and employ artificial neural networks in the supervised classification of spectral information. EASl/PACE was used for the ANN classification reported in this work. EASI/PACE includes an implementation of a simple feedforward BP- supervised classifier with a momentum term. This is the only neural network classification algorithm currently supported in the package. One advantage this software possesses is that mapping the results is easily and automatically accomplished by the module, thus facilitating visual analysis. The Stuttgart Neural Network Simulator (SNNS) was also employed in this research for early comparisons. SNNS is offered free of charge by anonymous File Transfer Protocol (FTP; Stuttgart Neural Network Simulator version 4.1, 1995). This software compiles on a Sun workstation and on other computer systems, is fairly well documented, includes a wide array of algorithms, provides onscreen network analysis and visualization tools, and is supported to a limited extent through an email discussion list. SNNS was known to be capable of producing and processing the small to medium-sized network anticipat- ed for this study (Lutzy and Dengel, 1993). While SNNS offers more speed, control of processing, graphical network analysis capabilities and flexibility in algorithm selection than does EASl/PACE, it lacks a simple method for quickly mapping output information. Therefore, the results reported here 73 were generated by the EASIIPACE algorithm, which is a variant of standard feedforward backpropagation that includes a ‘momentum’ term. The two packages were found to provide nearly identical output in previous research, given the use of an appropriately configured BP with momentum algorithm in SNNS (Brown, Lusch and Duda, 1998). The sampled data described in section 4.1.1.3 was used for training the simple BP with momentum ANN for the Lower Peninsula and county data sets. The trained networks were then assigned the task of classifying all pixels in the respective Lower Peninsula and Roscommon County scenes, with the exception of those masked as being nonsusceptible to defoliation. Tables 4.2 and 4.3 provide summary statistics for NDVI in the Lower Peninsula and Roscommon County, respectively, as briefly discussed in section 4.1.2. The BF network architecture was established in part by the number of variables used as the basis for classification, in this case three (pm-defoliation, peak defoliation, and post defoliation NDVI). These three variables each were assigned to one input element in the network. The architecture was also fixed by the number of classes to be defined, in this case two (defoliation and nondefoliation). These two classes were each assigned to one output element in the network. Thus, the number of processing elements in the input and output layers were established by the data. The number of layers between these two outer layers and the number of processing elements within each internal layer are typically determined heuristically. These internal layers are referred to as ‘hidden.’ The use of too many internal processing elements results in the memorization of the training vectors, while too few hidden nodes render the network unable to learn through weight updates (Hammerstrom, 1993b) 74 The Neural Network Frequently Asked Questions (FAQ) list maintained by Warren S. Sarle of SAS institute, inc. offers the following comments regarding hidden layer architecture optimization: Some books and articles offer “rules of thumb" for choosing a topology - Ninputs plus Noutputs divided by two, maybe with a square root in there somewhere - but such miss are total garbage. There is no way to determine a good network topology just from the number of inputs and outputs. it depends critically on the number of training cases, the amount of noise, and the complexity of the function or classification you are trying to Ieam. There are problems with one input and one output that require thousands of hidden units, and problems with a thousand inputs and a thousand outputs that require only one hidden unit, or none at all. Other rules relate to the number of cases available: use at most so many hidden units that the number of weights in the network times 10 is smaller than the number of cases. Such rules are only concerned with overfitting and are unreliable as well. All one can say is that if the number of training cases is much larger (but no one knows exactly how much larger) than the number of weights, you are unlikely to get overfitting, but you may suffer from underfitting (Sarle, 1998c). Preliminary evaluation of alternative parameter specifications was conducted using SNNS software. As suggested by the preceding quotation, parameter optimization could quite easily have become the subject of a separate, detailed research project. The areas investigated prior to completing the classifications reported in this work included variations in backpropagation algorithms, number of input units, number of hidden layers, number of nodes within each hidden layer, number of output units, ieaming rate, momentum and number of iterations (epochs). Output was evaluated based on logged training and testing errors, through visual interpretation of training and testing error curves, and through the use of the “Analyze” tool within SNNS, which provided the number of patterns classified correctly, incorrectly and those not classified. The tested 75 parameter values were within ranges recommended by the software developers and other researchers. The algorithms evaluated were Quickprop and backpropagation with momentum. The inputs were expanded from the basic pre-defoliation, peak defoliation and post defoliation NDVI values to also include ratios and differences of these figures. Due to the coarse spatial resolution of the sensor compared with the areal extent of the targets, a moving window of averaged NDVI values, or ‘texture’, was not added to the inputs. Architectures with both one and two hidden layers were evaluated, containing between 2 and 50 nodes per single hidden layer and 5 to 30 nodes for each of the two hidden layers. Architectures with both one and two output nodes were W. The ieaming rate was assigned within the range of 0.1 to 1.0, and momentum was restricted to within 0 to 1.0, as suggested in SNNS documentation. ANN training was conducted for up to 1,000 iterations, based on inspections of the graphed descent of the error cunres. in addition to these modifications in ANN parameters, alternative data sampling techniques were also evaluated during this phase. A markedly superior combination of these key elements of the backpropagation ANN approach was not found through these preliminary trials, when assessed as described in the preceding paragraph. Consequently, EASIIPACE configuration recommendations, all within the range of tested parameter values, were adopted for use in the final classifications. Confirmation that ANN parameters have been specified such that the best possible solution was realized is not attainable, and researchers must simply assess whether the resulting accuracy is sufficient for the given application (Garson, 1991; Jain, 1994). Caution must be exercised in assigning processing elements to the hidden layer, since 76 an overabundance could add complexity that serves no useful purpose, thus limiting performance (Eliot, 1993). The simple fully-interconnected network configuration used in producing the reported classifications was of the form 3 input elements, 3 internal elements in one hidden layer, and 2 output elements. The momentum utilized was 0.9 and the ieaming rate was set to 0.1, with a run length of 1000 processing iterations. Network training times for the small file sizes were on the order of 0.25 hour or less on a Sun Ultra 2 workstation. Minimum error thresholds were ignored in order to assign every pixel to the most activated class. An output image was generated for each classification indicating the most likely class for each pixel. While the number of input variables presented for final classification could have been increased through NDVI band differencing or ratioing or through the use of non-ratioed AVHRR channel 1 and channel 2 data, this work restricted ANN (and maximum likelihood) input to the simple case of pre-defoiiation, peak defoliation and post defoliation NDVI ratios. in part, this was to assess whether such input was sufficient to satisfactorily resolve the decision boundaries. in addition, such an approach attempted to avoid the negative effects of added dimensionality while exploiting the advantages cited in secfion 2.2.2 concerning the use of ratios. The implications of the ‘curse of dimen'sionality' (described by Bellman, 1961) in ANN applications have been summarized by Jenna Sinkkonnen to refer to the ail-consuming application of network resources in the representation of “irrelevant portions of the space” due to the ‘exponential growth of hypervolume as a function of dimensionality“ (Saris, 1998b). 77 4.2 Accuracy Assessment There is a lack of general guidelines for selecting a method for assessing classification accuracy (Congalton, 1991; Foody, McCullough and Yates, 1995a), though common approaches include percent correctly classified and Kappa. The percent correctly classified overestimates classification accuracy because an adjustment for agreement due to chance is not applied. Kappa, though addressing chance agreement, has been shown by Foody (1992) to constantly underestimate the overall accuracy of the classi- fication being evaluated. Further, Congalton (1991, p. 45) has noted that “only simple random sampling completely satisfies' an assumption on which Kappa analysis is based. Another method, Tau, has been proposed as an improvement over Kappa (Ma and Redmond, 1995). in the manner of Foody, McCullough and Yates (1995a), the percent correctly classified is presented in this work. Further, a crosstabulation of the classed output versus reference information was performed to enable the creation of an error matrix and the calculation of the more conservative Kappa statistic, as described by Jensen (1996). Producer and user accuracies were thus also available for comparison. The error matrix compares the pixel assignment for each class in the reference and the classification output. The percent correctly classified is the total number of cells classed correctly divided by the total number of pixels classified, with the result multiplied by 100. PCC thus may range from zero to 100%. Producer accuracy reflects the percent correct for each class based on the reference image. User accuracy considers the percent correct for each class based on the classification output. The Kappa statistic uses the total number of pixels under consideration, the sum of correctly classed pixels, 78 and the products of row and column totals to produce the amount by which the classification exceeds random assignment, which would have a value of 0%. The Kappa for purely nonrandom assignment would be 100%. Tests were conducted to assess whether the Kappa values for each classification were significantly different than zero. The results should be interpreted with caution, however, because by taking every defoliated cell for accuracy assessment, the variance was underestimated because the cells were not independent of each other. A significance test was performed to assess the level of difference between ML and ANN error matrices for the Lower Peninsula. Such a test for the two classifications of Roscommon County data was not possible due to the presence of only values equaling zero in one row of the ANN matrix. Muchoney and Haack (1994) used SPOT HRV and panchromatic (PAN) data acquired on June 15, 1987 and July 4, 1988 in a comparison of the ability of several classifiers to detect change caused by gypsy moths in the forest leaf state of a small locale in Virginia. Though their methods differ from those employed in this study and the results are thus not directly comparable, Muchoney and Haack’s published summary statistics were contrasted with the results of this investigation. Readers should be aware that Hudson and Ramm (1987) summarized the occurrence of several erroneously published equations for the calculation of the Kappa statistic, and provided the correct version to the remote sensing community. They further noted that the correct formulation was presented by Fleiss, Cohen and Everitt (1969), and that the measure was originally conceived by Cohen (1960). Jensen (1996) presented a correct formulation, algebraically equivalent to that of Fleiss, Cohen and Everitt (1969) and 79 identical to the correct formulation noted by Hudson and Ramm (1987) to have been published by Bishop, Feinberg and Holland (1975). 80 Chapter 5 RESULTS, DISCUSSION AND CONCLUSIONS 5.0 Results, Discussion and Conclusions This work represented an attempt to utilize readily available AVHRR satellite imagery to detect Michigan forest defoliation caused by gypsy moths by employing multitemporal signatures and ancillary GIS data in supervised classification exercises. The maximum likelihood technique, a conventional approach, and BP ANN, a somewhat nonconventional method for this application, were the supervised classifiers chosen for the task. The analysis integrated spectral, temporal, and spatial information in the classification and provided a test of the potential to automate and improve upon the standard MDNR defoliation assessment, which is based upon human interpretation of airborne video reconnaissance, by using the commme applied NDVI indicator of vegetation existence and condition. In this Chapter, the results obtained in all trials are reported, research questions are answered, the implications are discussed and conclusions are drawn. 5.1 Results As described in section 4.2, the accuracy of each classification result was quantitatively assessed through the construction and interpretation of error matrices and the calculation of the percent correctly classified and the Kappa statistic. The results obtained for each region and classifier are summarized in the two following sub-sections. 81 5.1.1 Lower Peninsula Classification All Michigan Lower Peninsula data were sampled and presented for maximum likelihood and artificial neural network classification, using the methods presented in Chapter 4. Complete data for the region were presented for classification based on signatures developed from the samples. A summary of both trials is presented here. Table 5.1 Error Matrix for Maxlrnum Likelihood classification of Lower Peninsula Data Reference image Defoliated Nondefoliated now TOTALS Defoliated 1449 15015 16464 8.8% Nondefoliated 493 6280 6773 92.7% COLUMN TOTALS 1942 21295 23237 74.6% 29.5% 33.3% PCC 0.9% Kappa In Table 5.1, results are presented for the maximum likelihood classification of Lower Peninsula data. It can be seen that 74.6% of the cells that were actually defoliated were correctly classed, yet only 8.8% of the pixels classed as defoliation were really defoliated. The majority of cells were incorrectly assigned to the defoliated class, indicating a high rate of commission error, and resulting in this difference in accuracy levels from the producers and users perspectives. While only 29.5% of the actual nondefoliated cells were correctly classed, 92.7% of those identified as nondefoliated were actually in that class. Overall, 33.3% of the nonmasked pixels in Michigan were correctly classified, however the low Kappa of 0.9% suggests that chance played a large role in the classification accuracy. The Kappa value was significantly different than zero at a 95% confidence level. 82 Table5.2 ErrorMatrleorNeuraiNetworkClasslflcatlonofLowerPeninsulaData Reference image Defoliated Nondefoliated now TOTALS Defoliated 1799 18511 20310 8.9% Nondefoliated 143 2784 2927 95.1% COLUMN TOTALS 1942 21295 23237 92.5% 13.1% 19.7% PCC 1 .1% Kappa The data in Table 5.2 reveal that, in the Lower Peninsula, the ANN classifier was able to correctly identify 92.6% of those pixels known to be defoliated, but only 8.9% of the cells classed as defoliated were actually members of that class. Again, a high rate of commission error is observed. Only 13.1% of the truly nondefoliated pixels were correctly assigned, yet in the classed image 95.1% of cells identified as nondefoliated were correct. The overall 19.7% correctly classified was low, and the Kappa value of 1.1% again indicated much agreement due to chance. The Kappa value was significantly different than zero at a 95% confidence level. At a 95% confidence level, the Lower Peninsula ML and ANN classifications were not significantly different. Both the maximum likelihood and ANN classifiers correctly classed a substantial percentage of the known defoliation, while being far less successful in identifying the nondefoliated pixels. This occurred because both techniques erroneously assigned over 60% of cells to the defoliated class. While the maximum likelihood method had an overall percent correctly classified equal to 1.7 times that of the ANN approach, both were shown 83 by the Kappa statistic to have exceeded agreement due to chance by only a very small margin. The signature separability measures TD and BD, described in section 4.1.1.3, both indicate a very poor separation between the defoliated and nondefoliated classifier training signatures. The TD value was 0.074 and the BD measure was 0.071. The signatures are statistically very similar, and thus do not contain suitable information to enable discrimination between the classes. 5.1.2 Roscommon County Classification As discussed in section 4.1.1.4, it was anticipated that classification accuracies would be adversely affected by the preponderance of mixed pixels in the statewide coverage, due to the coarse spatial resolution of the spectral sensor. For purposes of comparison, the Roscommon County area was selected due to the widespread occurrence of known defoliation within the county boundaries. All Roscommon County data were sampled and presented for maximum likelihood and artificial neural network classification in a manner similar to that performed for the Lower Peninsula. Complete data for the region were presented for analysis based on signatures developed from the training samples. The results from each trial are summarized here. Tabie5.3 ErrorMstrleorMaxlmum LikelihoodClasslficadonofRoscommonCountyData Reference image Defoliated Nondefoliated ROW TOTALS Defoliated 88 164 252 34.9% Nondefoliated 72 431 503 85.7% COLUMN TOTALS 160 595 755 55.0% 72.4% 68.7% PCC 22.7% Kappa As detailed in Table 5.3, the maximum likelihood classification of the Roscommon County data resulted in the successful identification of actually defoliated cells in 55.0% of the cases, and 34.9% of the cells classed as defoliated were really members of that class. A greater level of accuracy was achieved in the correct identification of truly nondefoliated pixels, at 72.4%, and 85.7% of the pixels classed as nondefoliated were actually in that category. The overall value of 68.7% correctly classified cells in the county was partially the result of chance, yet the 22.7% Kappa statistic confirms that considerable agreement was produced by reasons other than chance. The Kappa value was significantly different than zero at a 95% confidence level. The 22.7% Kappa statistic of the Roscommon County maximum likelihood classification is consistent with results reported by Muchoney and Haack (1994), which ranged from 20.6% to 45.4% and averaged 31.6%. They evaluated five classifiers for the detection of gypsy moth defoliation in a 148 kilometer2 (57 miles”) study area in Virginia. For comparison, the area of the Roscommon County scene was 1,571.7 kilometel’ (606.8 miles”). They provided summary data for methods termed merged PCA, PCA eigenimage (density level slicing), image differencing, spectral-temporal, and post-classification. The maximum likelihood decision rule was applied in the merged PCA, spectral-temporal, and post-classification 85 approaches. Their work was based on June 15, 1987 peak-defoliation SPOT HRV-XS data and July 4, 1988 peak defoliation SPOT HRV-XS and PAN information, and employed 1:24.000 US. Forest Service aerial sketch maps as reference information, digitized and converted to raster format. Defoliation categories for 1988 were heavy defoliation from 61% to 100% defoliated, moderate defoliation from 31 to 60% defoliation, and nondefoliated in the range 0% to 30% defoliation. Categories for 1987 were only identified as defoliated or nondefoliated. Consequently, they combined the 1988 data to codings similar to 1987, including moderate defoliation in the defoliation category. The minimum defoliation polygon mapped in their work was approximately 4 hectares (10 acres). Significantly, the results achieved in the single-year multitemporal maximum likelihood classification of Roscommon County 1,300 meter-square AVHRR data were comparable to those produced using the much finer resolution two-year multitemporal SPOT HRV-XS 20- meter-square multispectral data, which was resampled to 10 meters based on panchromatic data. The Kappa statistic for the maximum likelihood classification of Roscommon County data was very similar to that produced by Muchoney and Haack using spectral-temporal and merged PCA classifiers, which had values of 20.6% and 24.9%, respectively. Their highest Kappa value was 45.4% for image differencing, and their lowest Kappa was the 20.6% figure for spectral-temporal. The average overall accuracy among the five classifiers in the Virginia study was 64.2%, with the spectral-temporaland merged PCA classifiers having 61 .1%.and 63.3%, respectively. A high of 74.2% for image differencing and a low of 60.7% for merged PCA were realized. This compares with the overall percent correctly classified value of 68.7% for the Roscommon County maximum likelihood trial. 86 stle5.4 ErrorMatrleorNeuralNetworkClessificatlonofRoscommonCountyData Reference Image Defoliated Nondefoliated ROW TOTALS Defoliated o o 0 0.0% Nondefoliated 160 595 755 78.8% COLUMN TOTALS 160 595 755 00% 100.0% 78.8% PCC 0.0% Kappa in the ANN classification of Roscommon County data, total confusion exists with regard to the identification of defoliated areas, given the input data and selected classification parameters. In this trial, the error matrix in Table 5.4 shows that every pixel was assigned to the nondefoliated class, resulting in 100.0% correct identification of the known nondefoliated cells. There were 78.8% of classed nondefoliated cells placed in the proper category. Not one defoliated cell was classed correctly with reference to either the ‘truth’ image or the classed image, and the level of accuracy in the classification of the targeted defoliated pixels was thus zero. While overall 78.8% of all cells were classed correctly, the wholesale assignment of every pixel to the dominant nondefoliated class resulted in a Kappa statistic equal to 0.0%, indicating pure chance agreement. Note that there was an unequal weighting of defoliated cells in the training sample, which would be interpreted by the ANN classifier as biased a priori information, as mentioned in section 2.3.2 and section 4.1.1.3. A significance test could not be perlonned to assess the level of difference between ML and ANN classifications for Roscommon County due to values of zero in the ANN error 87 matrix. The 22.7% Kappa resulting from the ML method indicated superior performance compared to the ANN approach which had a Kappa value of 0.0%. For the defoliated class that is of particular interest in this research, the maximum likelihood method returned superior results in both producer and user accuracy. The ANN PCC was 10.1% higher than that of the maximum likelihood method because all pixels were assigned by ANN to the dominant nondefoliated class, though the Kappa statistic was far superior for the maximum likelihood classification. One similarity between the maximum likelihood and ANN approaches was that each realized a higher PCC in the Roscommon County exercise than was achieved in the Lower Peninsula classification. As also found for the Lower Peninsula data, the signature separability measures TD and BD both indicate a very poor separation between the defoliated and nondefoliated classifier training signatures used for Roscommon County data processing. The TD value was 0.252 and the BD measure was 0.232. While these values represent an improvement over those in the Lower Peninsula analysis by a factor of approximately 3.3 to 3.4, the signatures are nonetheless statistically very similar, and thus do not provide sufficient information to enable discrimination between the classes. 5.2 Answers to Research Questions and Discussion 1) Can the maximum likelihood supervised classification technique be successfully applied to the problem of detecting defoliated areas? At the county level, the maximum likelihood method used in this research was demonstrated to generate a 55% producer classification accuracy for the primary defoliated target, with 88 approximately 35% of pixels classed as defoliated actually being defoliated. The results compare favorably with those achieved in a study by other researchers (Muchoney and Haack, 1994). However, the relatively low accuracy would most likely prove insufficient for a monitoring program upon which land management decisions would be based. On a Lower Peninsula basis, the low 8.8% user accuracy of cells classed as defoliated would no doubt be unacceptable to land managers. Additional details related to this question are provided in answer to the fifth research question. it is apparent that overlap exists in the class probability density functions. 2) Can the ANN classification technique be successfully applied to the problem of detecting defoliated areas? Based on this trial, and given the previously described efforts to optimize architecture and leaming parameters, it must be concluded that the defoliated targets cannot be satisfactorily discriminated from the nondefoliated pixels by the ANN approach. While high producer accuracy was achieved at the Lower Peninsula level, the low user accuracy and Kappa statistic indicate substantial misciassification and dependence on chance agreement. A more intensive investigation of the BP ANN method in this particular application may very well enhance between-class boundary definition sufficiently to enable improved classification, though the very poor training signature separability remains a critical concem. 3) How do the accuracy levels of the maximum likelihood and the ANN techniques compare In this application, given the final parameters utilized? As one might surmise from the answers to the two preceding questions, this work shows that, given the methods adopted and the spectral information presented, the maximum likelihood classification technique was better able to identify those pixels confirmed by the 89 interpretation of airborne video reconnaissance to contain trees defoliated through the action of gypsy moths. In other recent work, Skidmore et al. (1997) also concluded that the ANN method failed to yield significant advantages over more common classification techniques in mapping forests using Landsat TM and GIS data. There were similarities between the methods in the general degree of percent correctly classified when the results were stratified by the areal extent of the classed region. Both techniques had lower levels of accuracy, expressed as PCC, in the Lower Peninsula exercises. This accuracy increased by a factor of two to four in the Roscommon County tests. At the Lower Peninsula level, both techniques generated defoliated producer accuracy in excess of 70%, and defoliated user classification accuracy was nearly identical for the two methods. Kappa values, though low, were very similar in the Lower Peninsula classifications. The inability of the ANN method to identify defoliation in the Roscommon County test, and the comparatively superior overall performance of the maximum likelihood technique in the same trial, led to the greatest difference in results between the two classification methods. Using the Kappa statistic as the judging criterion, the 22.7% achieved by the maximum likelihood classification of Roscommon County data was the best outcome realized in these trials. 4) Are the classification accuracy levels for defoliated and nondefoliated pixels similar? The two classification methods returned varying results depending on the method by which accuracy was assessed. in the Lower Peninsula tests, the ANN approach provided superior producer accuracy for the target defoliated cells, while the maximum likelihood technique had an advantage in terms of producer accuracy in the discrimination of nondefoliated areas. Both methods generated similar results for each category when user accuracy was evaluated. As a percentage of the reference image, the ANN technique applied to the full Lower Peninsula data set displayed a greater ability to accurately identify defoliated pixels, by a margin of 18% (92.6% - 74.6%). As a percentage of the classed image for the same situation, the results were nearly identical, with only 0.1% greater accuracy for the ANN approach (8.9% - 8.8%). As a percentage of the reference image for Lower Peninsula data, the maximum likelihood technique more accurately classed nondefoliated pixels by a margin of 16.4% (29.5% - 13.1%). As a percentage of the classed image for this situation, accuracy was similar, with a difference of only 2.4% (95.1% - 92.7%), in favor of the ANN method. As a result of the wholesale assignment of all pixels to the Roscommon County nondefoliated class by the ANN technique, the accuracy rate for the nondefoliated pixels was far superior to that for the defoliated pixels. which were classed with an accuracy of zero. The margin of difference in ANN classification accuracy between defoliated and nondefoliated cells as a percentage of the reference image was 100.0%. As a percentage of 91 the classed image, this difference was 78.8%. In both cases, the margin was in favor of the nondefoliated class. The least difference between defoliated and nondefoliated producer and user accuracy levels among the four trials was realized for both accuracy viewpoints by the maximum likelihood Roscommon County classification. In this case, there was a difference of 17.4% between the producer accuracies for defoliated and nondefoliated classes. There was a 50.8% difference between the user accuracies for defoliated and nondefoliated classes. The greatest polarization in terms of producer accuracies occurred in the ANN classification of Roscommon County data, with a difference of 100%. The greatest difference when considering user accuracies occurred in the ANN classification of the Lower Peninsula data, though the 86.2% value was similar to those for the three trials other titan the Roscommon County maximum likelihood test. 5) At a subpixel level of detail, what insight may be gained from an analysis of pixel homogeneity? Due to the ‘fuzziness’ of the resulting spectral information, pixels containing a combination of the desired class and another class, or several other classes, are difficult for any hard classifier to properly discriminate and assign when partitioning the feature space. When considering the effect of mixed pixels in these classifications, several additional questions may be raised in connection with the work described, which will be answered in this section in relation to the primary issue of pixel homogeneity, or lack thereof. 92 This discussion takes a different approach from that of Foody (1996), who investigated the relationship between mixed pixels and the output values of an ANN classifier after spatially degrading finer resolution imagery to approximate AVHRR data. He found that, though the output activation values did not have a strong relationship to the pixel content, it was possible to rescale these figures to markedly improve the relationship. What is the abundance of mixed pixels In the Lower Peninsula and Roscommon County data sets? it is certain that most defoliated pixels in the entire data set are mixed. Table 5.5 shows the Lower Peninsula and countywide abundance of defoliated pixels for each of four ranges of true defoliation percentages. Only 19.1% of all Lower Peninsula defoliated pixels (372 of 1,942) had true defoliation levels greater than or equal to 50%, which was used as the selection criterion for the defoliated training sample. In Roscommon County, 34.4% of the defoliated pixels had a percent defoliated greater than or equal to 50%. Table 5.5 Lower Peninsula and Roscommon County defoliation by actual percent defoliated quartiles Actual Percent Lower Lower Roscommon Roscommon Defoliated Peninsula Peninsula Co. Number of Co. Percent of Number of Percent of All Defoliated All Defoliated Defoliated Defoliated Pixels Pixels Pixels Pixels 75 to 100 78 4.0 11 6.9 50 to 75 294 15.1 44 27.5 25 to 50 473 24.4 30 18.7 0 to 25 1097 56.5 75 46.9 COIUIIII 701818 1942 100.0 160 100.0 93 How do the frequency dlstrlbutlons of the per-plxel proportion defoliated compare for all defollated pixels In the Lower Peninsula and In Roscommon County, and do these differ from the distribution for the correctly classified defoliated pixels In the best outcome realized among the four classlficatlons performed? Figure 5.1 and 5.2 provide a clear visual impression that the majority of pixels had low levels of defoliation in both the Lower Peninsula and the Roscommon County data sets. In contrast to the Lower Peninsula histogram in Figure 5.1, the frequency distribution in Figure 5.2 for Roscommon County appears somewhat bimodal, and differs markedly from that for the Lower Peninsula, though the highest frequencies are still in the lowest proportions defoliated. The spatial distribution of the information presented in Figure 5.2 is shown in Figure 5.3, and it can be seen that the primary region of high percent defoliation occurs near the center of the large cluster of defoliated pixels. 0.03 QED 0.18 0.27 0.3 0.5 0.54 0.63 0.72 0.81 OR) Figure 5.1 Frequency distribution of proportion defoliated for all Lower Peninsula defoliated pixels. 94 0CD one 0.18 0.27 0.3 0.46 0.64 0.63 0.72 0.81 can Figure 5.2 Frequency distribution of proportion defoliated for all Roscommon County defoliated pixels. Figure 5.3 Spatial distribution of proportion defoliated for all Roscommon County defoliated pixels. Red is high percent defoliation, yellow is mid-range, and green is low. 95 Figure 5.4 shows the frequency distribution for the actual proportion of defoliation for those Roscommon County pixels correctly classed as defoliated by maximum likelihood. The bimodal distribution apparent in Figure 5.2 for all defoliated pixels in the Roscommon County data remains evident, and is slightly more pronounced. 0.CD one 0.18 0.27 0.3 0.45 0.54 0.53 0.72 0.81 090 Figure 5.4 Frequency distribution of proportion defoliated for all Roscommon County defoliated pixels correctly classed as defoliated by ML classifier. Roscommon County was selected for detailed analysis because it was one of the few counties to experience widespread defoliation during 1993, and thus a higher percentage of pixels were anticipated to be more thoroughly defoliated than in most other counties. This higher level of pixel defoliation saturation is seen when comparing the percent defoliation frequency distributions for the Lower Peninsula and Roscommon County. In this sense, 96 Roscommon County was not representative of all counties in the entire State of Michigan. Counties having arbitrary political boundaries, this was not considered significant since the task was solely to develop characteristic spectral signatures in order to enable the identification of defoliated sites through multitemporal NDVI input vectors. Such a scene encompassing a dense yet extensive defoliation cluster with a higher frequency component at maximum pixel purity levels served the research purpose since the data were screened to include only broadly similar forested areas. Selecting only the intensely defoliated Roscommon County cluster would, however, have certainly yielded too few pixels for classifier training and evaluation purposes (Jensen, 1996). The less mixed, or more homogenous, pixels were desired in order to optimize classifier training, as is customary in supervised classification efforts (Jensen, 1996; Schowengerdt. 1983). The Roscommon County maximum likelihood results will serve as the focus for analysis in the following questions ccnceming the relationship between accuracy and pixel homogeneity. Based on the Kappa statistics presented in section 5.1, this classification was clearly superior to the other three tests, and a detailed review of the accuracy at subpixel levels may shed light on how this was achieved. Recall that the Kappa values indicated negligible agreement beyond chance for the other classifications. Were cells outside the percent defoliated training range also accurately classified? In the training, only pixels having greater than or equal to 50% defoliation were used in the defoliation sample. Furtheranalysis, summarized in Table 5.6, revealed that 60.2% of the pixels correctly classed as defoliated in Roscommon County had a true defoliation level of less than 50%. In addition, 60.0% of all the defoliated pixels with a true defoliation level of 97 greater than or equal to 50% were correctly classed. while 50.5% of all the defoliated pixels with a true defoliation level less than 50% were correctly classed. The maximum likelihood classifier thus did correctly identify many of these more highly mixed pixels outside the range of those presented during signature development, though the Kappa statistic did indicate chance assignment played a role in the overall accuracy realized. Table 5.6 Roecomrnon County maximum likelihood results for training sample range and tor category outside training range ActualPercentDefollstlon PercentovaerallCorrectly PercentCorrectofAll Range Classed Defoliation for Defoliated Pixels in Range I“IMO Greaterthanorequaito50% 35/88239.8% 35/55=60.0% Lessthan50% 5338860296 53/105=50.5% How does classification accuracy vary for differing pixel mixture levels of correctly and Incorrectly classed defoliation, and are the less mixed, or purer, defoliated plxels Indeed more accurately classed? The relationship between the classification accuracy for defoliated pixels and the degree of pixel class purity (which may also be termed the mixture level) is shown for the Roscommon County maximum likelihood classification in Table 5.7 and Figure 5.5. As anticipated, the percentage of pixels correctly classified as defoliated increases as the level of actual defoliation increases. Larger percentages of pixels were incorrectly classed as the level of actual defoliation decreased. Correctly classed defoliated pixels represent a larger percentage of the pixels in the highest percent defoliation range (75 to 100%) than incorrectly classed defoliated pixels (81.8% versus 18.2%, respectively). incorrectly classed defoliated pixels represent a larger 98 percentage of the lowest percent defoliation range (0 to 25%) than correctly classed defoliated pixels (53.3% versus 46.7%, respectively), though by a much smaller margin than the previous situation (63.6% and 6.6%, respectively). Table 5.7 Roscommon County maximum likelihood defoliated accuracy and class homogeneity Actual Defoliation Defoliation Defoliation Defoliation Defoliated Correct Percent Correctly Correctly Incorrectly Incorrectly Pixels In Divided by Defoliation Classed Classed Classed Classed Range Incorrect Ran (Mixture Number of Percent of Number of Percent of Total Ratio Level) Pixels Range Pixels Range (Percent) 75 to 100 9 81.8 2 18.2 11 (6.9) 4.50 50 to 75 28 63.6 16 36.4 44 (27.5) 1.75 25 to 50 16 53.3 14 46.7 30 (18.7) 1.14 0 to 25 35 46.7 40 53.3 75 (46.9) 0.88 Column 88 NA. 72 MA. 160 (100.0) N.A. Totals 90 80 if § 70 it I E 60 4— .E a 50 1' —e— Correctly Gassed E 40 4 —I—- incorrectly Gassed .6 30 .- E 20 ~ 3. 10 i 0 f i .L t o 20 4o 60 80 100 Percent Defoliation Figure 5.5 Subpixel percent of range versus percent defoliation for Roscommon County ML classification. The four midpoints of the percent defoliation ranges were used to plot the percent defoliation. 99 The percent of pixels in a defoliation range (y-axis) for the correctly classed pixels increases with increasing actual percent defoliation (x-axis) (Figure 5.5), while the percent of pixels in a range for the incorrectly classed pixels decreases with increasing actual percent defoliation. The crossover point, or point at which the correctly classed and the incorrectly classed curves intersect, occurs at approximately 25.5% actual defoliation, as scaled from Figure 5.5. A vertical line above and below this point and intersecting the x-axls separates two regions where either the correctly classed curve (on the right side) or the incorrectly classed curve (on the left side) dominates the percent defoliation range (y-axis). The correctly classed defoliated pixels with greater than 25.5% actual defoliation possess a larger percent of the defoliation range than the incorrectly classed defoliated pixels. The incorrectly classed defoliated pixels with less than 25.5% actual defoliation claim a larger percent of the defoliation range than the correctly classed pixels. What are the Implications concernlng the sensitivity of the classifier, and what do the classification results suggest ccnceming the area of detectable defoliation wlthln pixels? it is noteworthy that the preponderance of correct classifications generated by the maximum likelihood classification technique begins to occur at a level of defoliation on the order of only 1/4 pixel in areal extent. Recall that the training sample consisted entirely of pixels with greater than or equal to 1/2 pixel defoliation, and that this portion represented only 34.4% of the total defoliated pixels. The pixels in this same training range of 50% to 100% defoliation represented only 39.8% of all correctly classed defoliated pixels, as noted in Table 5.6. Thus, 60.2% of the correctly classed defoliated pixels were outside the sampled defoliation range presented to the classifier for signature development. The maximum likelihood 100 classifier thus exhibited some ability to identify nonhomogenous pixels, yet there was a corresponding penalty in that 65.1% (164/252) of pixels classed as defoliated were actually nondefoliated, as found through interpretation of the error matrix in Table 5.3. The foregoing indicates the existence of overlapping decision boundaries. When considering the application of this general AVHRR-based approach to gypsy moth defoliation detection, and in related fieldwork, it could be useful to know what areal extent of defoliation occurrence is required to begin achieving more correct classifications than incorrect classifications for defoliated pixels. Percent defoliation is equivalent to the areal extent of defoliation within a pixel divided by the total area of the pixel. For the 1.3 kilometer-square mid-latitude spatial resolution of the AVHRR pixels employed in this exercise (1 .3’ = 1.69 kilometer” area), the 25.5% actual pixel defoliation crossover value (value of x at the intersection of the correct and incorrect curves in Figure 5.5) represents an area of 43.1 hectares (106.5 acres). Thus, correctly classed within-pixel defoliated areas greater than 43.1 hectares (Le. greater than 25.5% actual defoliation) possess a larger percentage of the defoliation range than the incorrectly classed defoliated pixels. This percentage difference steadily increases with greater pixel percent defoliation from the crossover point to the maximum x-axis value, or highest percent defoliation, as seen in Figure 5.5. Another way to consider this accuracy information in relation to pixel mixing is to note that the ratio of correct to incorrect is 1.0 at the intersection of the correct and incorrect curves and greater than 1.0 to the right of the intersection. It is less than 1.0 to the left of the intersection, which occurs at 25.5% defoliation. Figure 5.6 illustrates this relationship 101 between the correctfincorrect ratio and the percent defoliation. At the maximum percent defoliation, the correctfincorrect ratio is 4.50 (81.8%l18.2%), up from 0.88 (46.7%/53.3%) for the lowest percent defoliation range. O .L . 12.5 37.5 62.5 87.5 Percent Defoliation Figure 5.6 Correct/Incorrect ratio versus percent defoliation for Roscommon County ML classification. The curve is nonlinear. The steepness of the curve between the 62.5% and 87.5% defoliation values indicates the relatively rapid improvement in classification accuracy within this zone as percent defoliation increases. The 62.5% and 87.5% values represent range midpoints within the 50-75% and 75-100% ranges, which also reflect the full range of the defoliated training sample used in the classification. At lower levels of defoliation content, 102 the improvement in classification accuracy was less pronounced for equal changes in percent defoliation. This general relationship also existed in the maximum likelihood classification of the Lower Peninsula data, but was not observed in the artificial neural network classifications at either regional level, as depicted in Figure 5.7. For reference, the percent defoliation levels in this figure and in Figure 5.6 correspond to the units of area provided in Table 5.8. Recall, however, that due to the method of converting the vector ranges of MDNR-interpreted defoliation percentages observed on video imagery, the maximum possible percentage of defoliation is 87.5. The foregoing has provided insight into the limitations imposed by the coarse resolution AVHRR receiver, given the areal distribution of target occurrences. If one considers that pixel class purity is the single greatest determinant of classification accuracy, the use of a finer spatial resolution sensor would be expected to improve classification accuracy, assuming cloud-free scenes could be acquired in the desired temporal windows at suitable wavelengths. For example, the Landsat Thematic Mapper (TM) imagery at similar wavelengths to those used by the AVHRR sensor has a spatial resolution of 30 meter- square (0.03 kilometer-square), in contrast to the 1,300 meter-square (1.3 kilometer- square) AVHRR resolution. The resulting 0.0009 kilometei‘ area of a TM pixel is very small when compared with the 1.69 kilometer2 AVHRR pixel, and 1,877.78 TM pixels are used to cover the same surface area as just one AVHRR pixel. However, note that Muchoney and Haack (1994) did not achieve substantially better results using relatively fine resolution SPOT data. 103 .9 i a: 10 i + M m 8 + M M. FDS m 6 + nos ML 4 i 2 "V 0 . 12.5 37.5 62.5 87.5 Percent [bfolletlon Figure 5.7 Correct/Incorrect ratio versus percent defoliation for Lower Peninsula and Roscommon County ML and ANN classifications. In the legend, Ml denotes Lower Peninsula data, and ROS signifies Roscommon County data. NN is the ANN classification, and ML is the maximum likelihood classification. Table 5.8 Areal extent and correct/Incorrect classification ratio for key values of percent defoliation AVHRR Percent Roscommon County Defoliation Area Defoliation Area Defoliation (midpoint Maximum Likelihood km' acres of range, except for Correct/incorrect 25.5 and 100.0) Ratio 100.0 NA. 1.69 417.6 87.5 4.50 1 .48 365.4 62.5 1.75 1.06 261.0 37.5 1.14 0.63 156.6 25.5 1.00 0.43 106.5 12.5 0.88 0.21 52.2 1.0 kilometer” = 247.10538 acres. 1 hectare = 2.4710538 acres. The use of 30-meter-square pixels instead of 1,300-metersquare pixels would, of course, result in a large increase in file size and processing time, as well as increased costs. For coverage of the Lower Peninsula of Michigan, on the order of 3 billion pixels would be needed prior to aggregating to produce a map of percent defoliated with approximately 188 million 30-meter TM pixels. For comparison, there were just under 100,000 pixels in the aggregated Lower Peninsula AVHRR image. The acquisition of a suitable number of cloud- free scenes at a sufficiently high temporal resolution could be problematical in this region, however. 5.3 Conclusions and Recommendations for Future Research Overall, among the four different classifications performed, the accuracy of the spectral- temporal classification of gypsy moth defoliation ranged from 19.7% to 78.8%. Producer accuracy for the defoliated class reached a high of 92.6% in the Lower Peninsula ANN test, and the highest user accuracy for the same class was 34.9%, generated in the Roscommon County maximum likelihood experiment. However, classification accuracy in only one of the four trials exceeded chance agreement by more than a negligible margin, as determined by Kappa tests. This result, 22.7%, was achieved by the Roscommon County maximum likelihood classification, and was consistent with values reported by Muchoney and Haack (1994) for several different classification techniques using spectral information with a much finer spatial resolution. Considerable misciassification resulted from the use of nonseparable training signatures which were in turn caused by severely mixed coarse resolution spectral pixels, rendering the output unsuitable for use in contributing to the management of the gypsy moth forest 105 defoliation detection problem. The applicability of the work was thus limited at least in part by a mismatch between the smaller typical target sizes and the large ground resolution of the sensor. This occurred even though a study period was selected that provided extensive clusters of defoliation in Michigan, compared to most other years for which data were available. The sensor selection involved a choice between conflicting needs for high temporal as well as high spatial resolution, given adequate spectral and radiometric resolution and minimal costs of data acquisition. With limited training data and poorly separable training signatures, the generalizing nature of the maximum likelihood technique yielded greater accuracy in the Roscommon County trial than did the attempt by the artificial neural network to seek detail with insufficient supporting information. The percent correctly classified values for the two classifiers were similar at each regional level, though both produced higher percentages in the Roscommon County analysis. The Roscommon County maximum likelihood accuracy results indicate that, while problems involving spatial resolution existed, the NDVI values representing the spectral component did provide information that enabled the identification of the targets. The use of individual channel data, in addition to NDVI, may facilitate improved definition of the class boundaries. it is unlikely that an alternate sampling method would produce improved defoliation or nondefoliation recognition accuracy. At the Lower Peninsula level, 19.2% of all defoliated pixels were used in training, and these possessed the highest percentages of defoliation content, or the greatest purity. In the Roscommon County investigation, 14.4% of all defoliated pixels were employed in training, and these were also at the highest levels of percent defoliation. 106 In addition to the nonhomogenous nature of the pixel class content, another limiting factor encountered with this data was the paucity of defoliated training cells. For the ANN technique, this resulted in a limited amount of training pixels per weight in the architecture, even though a minimal number of internal, or hidden, processing elements were specified. For the defoliated target, this ratio was equal to 24.8 for the Lower Peninsula and only 1.5 for Roscommon County. A very slightly higher Kappa statistic resulted for the Lower Peninsula ANN classification. The sampling method favored the ANN approach by selecting nondefoliated training cells at the spatial class boundaries instead of in the midst of similarly classed pixels as is typical in maximum likelihood analysis. While suboptimal ANN parameter specification may have contributed to reduced overall classification accuracy, the similarly low level of accuracy achieved by the maximum likelihood method suggests that the realized outcomes were primarily influenced by other causes. Analysis of subpixel information provided confirmation that, as anticipated, correct defoliation classification was directly, though nonlinearly, proportional to the percent of each cell that was defoliated. Correct results rose dramatically when the defoliation content exceeded 62.5%, as evidenced by the county level study that was configured to provide a greater abundance of more highly defoliated pixels than were available on a Lower Peninsula basis. Correct classification of cells with low percent defoliation did occur. Given the finding from this thesis research that greater accuracy might be achieved by minimizing ground mixing through the use of a finer resolution sensor, why, then, did such an approach fail in other research to produce accuracies much greater than those realized l07 with the coarse resolution AVHRR sensor? Under peak defoliation leaf-off conditions, lower NDVI values were expected due to the removal of the high-NDVI leaf component of the spectral reflectance signal. This would facilitate the identification of affected regions. Yet, with the absence of a dense canopy, one might surmise that any reduction in NDVI through leaf loss may have been countered by the greater contribution in the near-infrared of the undergrowth to the reflectance received at the sensor, as discussed by Spanner et al. (1990). This would thus inhibit multispectral classification of defoliation, even with data acquired at a fine spatial resolution. However, visual inspection of Michigan CIR imagery indicated minimal NIR contribution from the understory, and this explanation seems doubtful (Lusch, 1998). This question remains unanswered, yet may involve differences in methodology, and also the tradeoff between boundary pixel influence and spectral variance as a function of spatial resolution as considered by Markham and Townshend (1981) and Woodcock and Strahler (1987). The influence of soil moisture variability on NDVl may play some role in partially explaining the lack of substantial improvement in classification accuracy when a finer spatial resolution sensor was employed by other researchers. Todd and Hoffer (1998) investigated the impact of changes in vegetation cover, soil texture and soil moisture content on NDVI, and confirmed that, as anticipated, increases in the percentage of vegetative cover yielded increases in NDVI. They observed that, for equal percentages of vegetation, NDVI levels increased for soils with a greater moisture content. “NDVI values associated with low reflecting soils were elevated compared with those associated with high reflecting, brighter soils' and ‘differences in soil type were less important than differences in soil moisture, in their impact on NDVI“ (Todd and Hotter, 1998, p. 918). They also found that the effects of 108 soil on the NDVI ratio were most pronounced when cover density was in the range from 20% to 80%. The coarse spatial resolution of the AVH RR imagery, the lack of radiometric and atmospheric calibration, variations in time and platform of acquisition, maximum value compositing, and timing of observations compared with forest defoliation stages were key sensor concerns in this endeavor. The suitability of the network architecture, adequacy of the training set, and the hazard of local error minima were the primary issues in the ANN classification. Compliance with Gaussian requirements was a factor in the maximum likelihood tests. The reference data against which both classifier outputs were judged was assumed to meet accuracy requirements, yet verification was not possible. There was a substantial lapse of time between AVHRR data collection and the acquisition of information that formed the basis for the forest mask applied in preprocessing. The required vector to raster conversion resulted in an increase in the extent of classed defoliation, and at the county level, there was a sizable difference in the number of training pixels for each class. Misregistration error between the reference information and satellite imagery did exist, though this was within generally accepted limits. The targets in this detection exercise were of variable, and often small, extent, especially when compared with the ground pixel size of the multispectral acquisition sensor. Ehrlich, Estes and Singh (1994) have stated that the lack of radiometric calibration in the AVHRR sensor is the primary shortcoming of this platform. Vermote and Kaufman (1995) offered a method for the absolute calibration of AVHRR channels 1 and 2. Further, the negative influence of atmospheric variability and differing sun angles were left uncorrected in the fine-resolution study by Muchoney and Haack (1994) and in this work, leaving the 109 two studies comparable in this regard. A spectral adjustment in the multitemporal images, as performed by Mikkola (1996) using the multiple regression method of Olsson (1993), could render the frequency distributions of digital numbers in all scenes more directly comparable, and this is recommended for any future research. Given the issues mentioned in the preceding paragraphs, this research attempted to fully exploit the generally reported robustness of the artificial neural network approach, and, to a certain extent, tested the limiting assumptions and suitability of the maximum likelihood classifier. It is recommended that the Forest Management Division of the Michigan Department of Natural Resources and other interested parties seek alternate approaches to the automated identification of scattered occurrences of forest defoliation caused by gypsy moths, as were typical during the 1993 season. These may include either the use of additional techniques to enhance multitemporal AVHRR data quality or the use of a sensor with finer spatial resolution than employed in this study. The former approach may remain limited by the contrast between the areal extent of the target and the sensor ground cell, and by the resulting aggregated spectral information. The latter option may raise additional concerns involving temporal resolution, and did not yield greatly superior accuracies when attempted by Muchoney and Haack (1994). 110 APPENDIX 111 Appendixa Selected World Wide Web Resources Related to . Michigan Forest Defoliation Detection Using Multitemporal NDVI Signatures Derived From AVHRR Satellite Imagery This listing is also available Online at http:llwww.ssc.msu.edul~geo/stu/dudalwwwgyp.html Accessto all sitesverified on April 2, 1998 Mutations Michigan Department of Natural Resources (MDNR) http://www.dnr.state.mi.usl MDNR Forest Management Division (FMD) http://www.dnr.state.mi.us/www/fmd/fmdhome.html Michigan State University (MSU) Department of Geography http://www.ssc.msu.edu/~geo/ MSU Center for Remote Sensing and Geographic lnfonnation Science httpzllwww.crs.msu.edul Great Lakes Environmental Research Laboratory (GLERL) http'J/www.glerl.noaa.gov/cw/cw.html National Oceanic and Atmospheric Administration (NOAA) httpzllwww.noaa.gov/ NOAA National Oceanographic Data Center (NODC) http://www.nodc.noaa.gov/ NOAA NODC Midwestern Climate Center httpzllmcc.sws.uiuc.edul 112 mm 1993 Michigan Forest Inventory 8. Analysis http:llwww.dnr.state.mi.uslwww/fmd/fia/surveyun.html Climate Summaries for the Midwest - Station Choices in Michigan http:l/mcc.sws.uiuc.edu/Summary/Michigan.html The Atlas of Michigan Project http:llwww.msu.edu/~atlasl Wishes Gypsy Moth in North America hwy/gypsy.fsl.wvnet.edu:80/gmottv Michigan's Gypsy Moth Education Program http:llwww.ent.msu.edulgypsyed/ MDNR FMD Forest Health Page http:llwww.dnr.state.mi.us/wwwffmd/pest/forhealhtml Evaluation of Forest Susceptibility to the Gypsy Moth Across the Coterminous United States http:llgypsy.fsl.wvnet.edu/gmoth/suscept/suscept.html WW Advanced Very High Resolution Radiometer http:l/edcwww.cr.usgs.gov/glislhyper/guide/avhrr The Remote Sensing Tutorial - An Online Handbook http:l/code935.gstc.nasa.govl'l'utorial/Start.html Satellite Environmental Monitoring of the Great Lakes: A review of NOAA's Great Lakes CoastWatch program. http:llwww.glerl.noaa.govlctrb/ctrb_0081 1 .htrnl NOAA CoastWatch Active Access System (NCAAS) http://www.nodc.noaa.gov/NCAASIncaas-home.htrnl 113 Classification PCl Maximum Ukellhood Classification http:llwww.pci.on.ca/cgi-binlpcihlp/MLCIALGORITHM What is an Artificial Neural Network? http:/Iwww.emsl.pnl.gov:2080/proj/neuronlneurallneural.ann.htrnl Neural Network Frequently Asked Questions (FAQ) ftp:/lftp.sas.com/pub/neural/FAO.html Sommers EASIIPACE and lmageWorks from PCl Geomatics Group http:llwww.pci.on.ca/ Stuttgart Neural Network Simulator (SNNS)from the University of Stuttgart - Version 4.1 Manual http:l/www-ra.informatik.uni-tuebingen.de/SNNSIUserManual/UserManualhtml IDRISI from Clark Labs http:llwww.idrisi.clatku.edul SNNS FTP Software Download Site in lpub/SNNS/ as file SNNSv4.1.tar.z fth/ftpjnfonnatikuni-stuttgartde/ 114 REFERENCES CITED 115 REFERENCES CITED Ahem, F. J., J. Sirois, W. D. McColl, R. P. Gauthier, T.T. Alfoldi, W. H. Patterson, and T. A. Erdle, 1991. Progress toward improving aerial defoliation survey methods by using electronic imagers. Photogrammetric Engineering & Remote Sensing. 57(2): 1 87-1 93. Albert, Dennis A., 1995. Regional landscape ecosystems of Michigan, Minnesota, and Wisconsin: A working map and classification (Fourth Revision: July 1994). North Central Forest Experiment Station, Forest Service, United States Department of Agriculture, General Technical Report NC-f 78, 250 p. Allen, Thomas R., Thaddeus J. Bara, and Stephen J. Walsh, 1994. Observed biases in the conterrninous U.S. AVHRR satellite data set: A North Carolina case study. Geocarto lntemational, 9(2): 52-62. An, Ping, and Chang-Jo F. Chung, 1994. Neural network approach for geological mapping: technical background and case study. Canadian Journal of Remote Sensing, 20(3): 293-301. Avery, Thomas E., and Graydon L Berlin, 1992. Fundamentals of Remote Sensing and Airphoto Interpremtion. Macmillan, New York, 472 p. Bailey, David, and Donna Thompson, 1990. How to develop neural-network applications. Al Expert, 5: 38-47. Barbour, Michael G., and William Dwight Billings (eds), 1988. North American terrestrial vegetation. Cambridge University Press, New York. Bellman, R., 1961. Adaptive Control Processes: A Guided Tour. Princeton University Press. Benediktsson, Jon A., Philip H. Swain, and Okan K. Ersoy, 1990. Neural network approaches versus statistical methods in classification of multisource remote sensing data. IEEE Transactions on Geoscience and Remote Sensing, 28(4): 540- 552. Bischof, H., W. Schneider, and A. J. Pinz, 1992. Multispectral classification of Landsat- images using neural networks. lEEE Transactions on Geoscience and Remote Sensing, 30(3): 482-490. Bishop, Y. M. M., S. E. Feinberg, and P. W. Holland, 1975. Discrete Mulfivanate Analysis - Theoryand Practice. MIT Press, Cambridge, Mass. 575 p. Blamire, P.A., 1996. The influence of relative sample size in training artificial neural networks. Intemational Joumal of Remote Sensing. 17(1): 223-230. 116 Brown, Daniel G., David P. Lusch, and Kenneth A. Duda, 1998. Supervised classification of types of glaciated landscapes using digital elevation data. Geomorphology, 21: 233-250. Center For Remote Sensing, 1991. CMAP Computer Mapping, Display, and Analysis Version 2.1 User's Manual. Michigan State University, East Lansing. Chen, Tao, and Mikio Takagi, 1993. The classification of NOAA images using multilayer neural network. Procedings of the 6th Australian Joint Conference on Artificial intelligence. Melbourne, November 1 6-1 9: 409-414. Cihiar, J., L. St-Laurent and J. A. Dyer, 1991. Relation between the normalized difference vegetation index and ecological variables. Remote Sensing of the Environment, 35: 279-298. Clark Labs, 1995. The ldrisi Project onscreen manual, ldrisi for Windows Version 1.01.002. Clark University, Worcester, MA. Cohen, J., 1960. A coefficient of agreement for nominal scales. Educational and Psychological Measurement, 20(1): 37-46. Congalton, Russell 6., 1991. A review of assessing the accuracy of classifications of remotely sensed data. Remote Sensing of the Environment, 37: 3546. Decatur, Scott Evan, 1989. Application of neural networks to terrain classification. Procedings of lntemational Joint Conference on Neural Networks, Washington, 0.0., 1: 283-288. Dottavio, C. Lisette, and Darrel L. Williams, 1983. Satellite technology: An improved means for monitoring forest insect defoliation. Journal of Forestry. January. Duda, Richard O., and Peter E. Hart, 1973. Pattern classification and scene analysis. John Wiley 8: Sons, New York, New York, 482 p. Eastman, J. Ronald, 1992. IDRISI Technical Reference, Version 4.0 rev. 1. Clark University, Worcester, Massachusetts, 213 p. Ehrlich, D., J. E. Estes, and A. Singh, 1994. Applications of NOAA-AVHRR 1 km data for environmental monitoring. lntemational Joumal of Remote Sensing, 15(1): 145- 161. Eliot, Lance B., 1993. Prefilter your neurons. AI Expert, 8:9-11. Fleiss, J. L., J. Cohen, and B. S. Everitt, 1969. Large sample standard errors of Kappa and weighted Kappa. Psychological Bullefin. 72(5): 323-327. Foody, Giles M., 1992. On the compensation for chance agreement in image classification accuracy assessment. Photogrammetric Engineering & Remote Sens- ing, 58(10): 1459-1460. 117 Foody, Giles M., 1995. Land cover classification by an artificial neural network with ancillary information. international Journal of Geographical lnfonnation Systems. 9(5): 527-542. Foody, Giles M., 1996. Relating the land-cover corhposition of mixed pixels to artificial neural network classification output. Photogrammetric Engineering 8 Remote Sens- ing, 62(5): 491 -499. Foody, Giles M., Mary B. McCullough, and William B. Yates, 1995a. Classification of remotely sensed data by an artificial neural network: Issues related to training data characteristics. Photogrammetric Engineering 8 Remote Sensing, 61 (4): 391 -401. Foody, G. M., M. B. McCullough, and W. B. Yates, 1995b. The effect of training set size and composition on artificial neural network classification. Intemafi'onal Journal of Remote Sensing, 16(9): 1707-1723. Foschi, Patricia 6., and Deborah K. Smith, 1997. Detecting subpixel woody vegetation in digital imagery using two artificial intelligence approaches. Photogrammetric Engineering & Remote Sensing, 63(5): 493-500. Garson, G. David, 1991. Interpreting neural-network connection weights. AI Expert, (6): 47-51. Goward, Samuel M, Brian Markham, Dennis G. Dye, Wayne Dulaney, and Jingli Yang, 1991. Normalized difference vegetation index measurements from the Advanced Very High Resolution Radiometer. Remote Sensing of the Environment, 35: 257- 277. Green, Kass, and Brian Cosentino, 1996. Using satellite imagery to detect and monitor forest change. Geo info Systems, 6(2): 22-30. Hammerstrom, Dan, 1993a. Neural networks at work. IEEE Spectrum, 30(6): 26-32. Hammerstrom, Dan, 1993b. Working with neural networks. lEEE Spectrum, 30(7): 46- 53. Hara, Yoshihisa, Robert G. Atkins, Simon H. Yueh, Robert T. Shin, and J. A. Kong, 1994. Application of neural networks to radar image classification. IEEE Transactions on Geoscience and Remote Sensing, 32(1): 100-109. Harman, Jay, 1984. Environmental significance of a Great Lakes location. Chapter 5 in Sommers, Lawrence M., Joe T. Darden, Jay R. Harman, and Laurie K. Sommers, 1984. Michigan, a Geography. Westview Press, Boulder, Colorado, 254 p. 118 Hepner, George E, Thomas Logan, Niles Ritter, and Nevin Bryant, 1990. Artificial neural network classification using a minimal training set: Comparison to conventional supervised classification. Photogrammetric Engineering and Remote Sensing, 56(4): 469-473. Huang, W.Y., and R. P. Lippman, 1988. Comparisons between neural net and conventional classifiers. Proceedings of lntemational Joint Conference on Neural Networks, San Diego, 4: 485-493. Hudson, William, and Carl W. Ramm, 1987. Correct formulation of the Kappa coefficient of agreement. Photogrammetric Engineering and Remote Sensing, 53(4): 421 -422. Hush, Don R., and Bill G. Home, 1993. Progress in supervised neural networks: What's new since Lippmann? IEEE Signal Processing Magazine, 10(1): 8-39. Ingells, Norris, 1995. Gypsy moth numbers are down. Lansing State Journal. January 4. Ingells, Norris, 1998. Attack gypsy moth with vigor. Lansing State Journal, March 18, 1-2 D. Jain, Anil, 1994. Personal conversation. Jensen, John R., 1996. introductory Digital image Processing: A Remote Sensing Perspective, second edition. Prentice Hall, Upper Saddle River, New Jersey, 316 p. Joria, Peter E., Sean C. Aheam, and Michael Connor, 1991. A comparison of the SPOT and Landsat Thematic Mapper satellite systems for detecting gypsy moth defoliation in Michigan. Photogrammetric Engineering a Remote Sensing. 57(12): 1605-1612. Klimasauskas, Casimir C., 1992. Neural networks: an engineering perspective. lEEE Communications Magazine, 30(9): 50-53. Lawrence, Jeannette, 1991 . Data preparation for a neural network. AI Expert, 6(11): 34-41. Leatherberry, Earl C., 1993. Forest statistics for Michigan’s Northern Lower Peninsula Unit, 1993. Resource Bulletin NC-157. US. Department of Agriculture, Forest Service, North Central Forest Experiment Station. St. Paul, MN, 48 p. Also available via MDNR Forest Management Division [Online] http:llwww.dnr.state.mi.usiwww/fmdifia/nIp/fianlp.html 3/23/98. Leshkevich, George A., David J. Schwab, and Glenn C. Muhr, 1993. Satellite environmental monitoring of the Great Lakes: A review of NOAA's Great Lakes CoastWatch program. Photogrammetric Engineering & Remote Sensing, 59(3): 371 -379. Also [Online] http:llwww.glerl.noaa.govlctrb/ctrb_00811.html Liebens, Johan, 1994. Procedures used in gypsy moth project. Unpublished memorandum. January. 119 Liebhold, Andrew, Kurt Gottschalk, and Doug Mason, 1998. Evaluation of Forest Susceptibility to the Gypsy Moth across the Coterminous United States. USDA Forest Service, Northeastern Forest Experiment Station. [Online] httpvlgypsyfsl.wvnet.edu/gmoth/suscept/suscept.html 3122/98. Lillesand, Thomas M., and Ralph W. Kiefer, 1994. Remote sensing and image interpretation, third edition. John Wiley & Sons, Inc., New York. Lippman, Richard, 1987. An introduction to computing with neural nets. lEEE Acoustics, Speech, and Signal Processing Magazine, 4: 4-22. Lusch, David P., 1993. Spectral reflectance characteristics of vegetation, soil, and water. Handout I in Lusch, DP. and Hudson, W.D., 1993, introduction to digital remote sensing, Fort Worth, Texas. USDA Soil Conservation Service, National Employee Development Office, 267 p. Lusch, David P., 1994. Unpublished notes. Lusch, David P., 1998. Personal communication. Lusch, DP. and F.J. Sapio, 1994. Time-Series Analysis Applied to Mapping Gypsy Moth Defoliation Using Multi-Date AVHRR Imagery. USDA, Forest Service, State Cooperators Meeting. St. Paul, MN, March 2, 1994. Lutzy, Ottmar, and Andreas Dengel, 1993. A comparison of neural net simulators. lEEE Expert, 8(4): 43-51. McConnack, Michael D., and Ron Day, 1993. How artificial intelligence impacts E & P productivity. World Oil, 214(10): 81-87. McKeown, David M., Jr., 1987. The role of artificial intelligence in the integration of remotely sensed data with geographic information systems. lEEE Transactions on Geoscience and Remote Sensing, 25(3): 330-348. Ma, Zhenkui, and Roland L. Redmond, 1995. Tau coefficients for accuracy assessment of classification of remote sensing data. Photogrammetn‘c Engineering 8 Remote Sensing. 61(4): 435-439. Markham, B. L., and J. R. G. Townshend, 1981. Land cover classification accuracy as a function of sensor spatial resolution. Proceedings of the 15" lntemational Symposium on Remote Sensing of Environment. Ann Arbor, Michigan, p. 1075- 1 090. Michigan State University Extension, 1998. Built-in controls of gypsy moth - the 3 P's. Michigan’s Gypsy Moth Education Program. [Online] http:llwww.ent.msu.edu/gypsyed/docsienemies.html 3/31l98. 120 Mikkola, K., 1996. A remote sensing analysis of vegetation damage around metal smelters in the Kola Peninsula, Russia. lntemational Journal of Remote Sensing. 17(18): 3675-3690. Montgomery, B.A. (editor), 1988. Gypsy Moth in Michigan 1987. Extension Bulletin E- 2142, Cooperative Extension Service, Michigan State University, East Lansing. Muchoney, Douglas M., and Bany N. Haack, 1994. Change Detection for Monitoring Forest Defoliation. Photogrammetric Engineering 8 Remote Sensing, 60(10): 1243- 1251. Murray, Dan, 1994. Tuning neural networks with genetic algorithms. Ai Expert, (9): 26- 31. Olsson, H., 1993. Regression functions for multitemporal relative calibration of thematic mapper data over boreal forest. Remote Sensing of the Environment. 46: 89-102. Openshaw, Stan, 1994. Neuroclassification of spatial data. Chapter 3, in Neural Nets: Applications in Geography, Bruce C. Hewitson and Robert G. Crane, eds. Kluwer Academic Publishers, Boston. Paola, Justin D., and Robert A. Schowengerdt, 1993. A review and analysis of neural networks for classification of remotely sensed multispectral imagery. Research institute for Advanced Computer Science, NASA Ames Research Center. Technical Report 93.05, (NASA-CR-f 94291 ), June. Paola, Justin D., and Robert A. Schowengerdt, 1995a. A detailed comparison of backpropagation neural network and maximum-likelihood classifiers for urban land use classification. lEEE Transactions on Geoscience and Remote Sensing, 33(4): 981-996. Paola, Justin D., and Robert A. Schowengerdt, 1995b. A review and analysis of backpropagation neural networks for classification of remotely sensed multispectral imagery. lntemational Journal of Remote Sensing. 16(16): 3033-3058. Paola, Justin D., and Robert A. Schowengerdt, 1997. The effect of neural-network structure on a multispectral land-useiland-cover classification. Photogrammetric Engineering 8 Remote Sensing, 63(5): 535-544. Richards, J. A., 1986. Remote Sensing Digital image Analysis. Springer-Venag, Berlin. Roesch Jr., Francis A., Paul C. Van Deusen, and Zhiliang Zhu, 1995. A comparison of various estimators for updating forest area coverage using AVHRR and forest inventory data. Photogrammetric Engineering 8 Remote Sensing, 61 (3): 307-311. Rumelhart, D.E., and J. L. McClelland, 1986. Parallel distributed processing: Explorations in the microstructure of cognition. MIT Press. 121 Sarle, Warren 8., ed., 1998a. Neural Network FAQ, Part 5 of 7: Free Software and Part 6 of 7: Commercial Software [Online] ftp://ftp.sas.comlpub/neuraVFAQ.html 3/22/98. Sarle, Warren 8., ed., 1998b. Neural Network FAQ, Part 2 of 7: Leaming [Online] ftp:l/'itp.sas.comlpub/neuraliFAQ.html 3/21/98. Sarle, Warren 8., ed., 1998c. Neural Network FAQ, Part 3 of 7: Generalization [Online] ftp'J/ftp.sas.comlpub/neurallFAQ.html 3/21/98. Sapio, Frank, 1998a. Entomophaga maimaiga! Better late than neverl Michigan Department of Natural Resources, Forest Management Division. [Online] http:l/www.dnr.state.mi.uslwww/fmdipestlentom.html 3/17/98 Sapio, Frank, 1998b. Unpublished intemal documents. Michigan Department of Natural Resources, Forest Management Division. Lansing, Michigan. Schowengerdt, R. A., 1983. Techniques for image Processing and Classification in Remote Sensing. Academic Press, New York, 249 p. Skidmore, A. K., B. J. Tumer, W. Brinkhof, and E. Knowles, 1997. Performance of a neural network: Mapping forests using GIS and remotely sensed data. Photogrammetric Engineering 8 Remote Sensing, 63(5): 501 -514. Sommers, Lawrence M., ed., 1977. Atlas of Michigan. Michigan State University Press, East Lansing, Michigan, 242 p. Sommers, Lawrence M., Joe T. Darden, Jay R. Herman, and Laurie K. Sommers, 1984. Michigan, a Geography. Westview Press, Boulder, Colorado, 254 p. Spanner, Michael A., Lars L. Pierce, David L. Peterson, and Steven W. Running, 1990. Remote sensing of temperate coniferous forest leaf area index: The influence of canopy closure, understory vegetation and background reflectance. lntemational Journal of Remote Sensing, 11(1): 95-111. Stuttgart Neural Network Simulator version 4.1, 1995. Provided via binary FTP transfer from host: ftp.infonnatik.uni-stuttgart.de in lpub/SNNS as file SNNSv4.1.tar.Z. Talerico, R.L., J. E. Walker, and T. A. Skratt, 1978. Quantifying gypsy moth defoliation. Photogrammetric Engineering 8 Remote Sensing, 44(11): 1385-1392. Todd, Stella W., and Roger M. Hotter, 1998. Responses of spectral indices to variations in vegetation cover and soil background. Photogrammetric Engineering 8 Remote Sensing. 64(9): 915-921. Townshend, John R. G., Christopher 0. Justice, Charlotte Gurney, and James McManus, 1992. The impact of misregistration on change detection. lEEE Transactions on Gewcience and Remote Sensing, 30(5): 1054-1059. 122 USDA Forest Senrice, 1998a. E. Leopold Trouvelot - Perpetrator of Our Problem. [Online] http:l/gypsy.fsl.wvnet.edu:80/gmoth/trouvelot/trouvelot.html 3/21/98. USDA Forest Service, 1998b. Gypsy Moth Digest - Suppression Costs. [Online] http:llwww.st.wvnet.edu/fl1p/gm_digest/costs.html 3/20/98. USGS EROS Data Center, 1998. Advanced Very High Resolution Radiometer. [Online] http:l/edcwww.cr.usgs.gov/glis/hyperlguide/avhrr 3/21/98. Vennote, E., and Y. J. Kaufman, 1995. Absolute calibration of AVHRR visible and near- infrared channels using ocean and cloud views. lntemational Journal of Remote Sensing, 16(13): 2317-2340. Wan, Eric A., 1990. Neural network classification: A Bayesian interpretation. IEEE Transactions on Neural Networks, 1(4): 303-305. Wann, M., T. Hediger, and N. N. Greenbaun, 1990. The influence of training sets on generalization in feed-forward neural networks. Proceedings of lEEE Fourth lntemational Conference on Neural Networks. 3: 137-142. Woodcock, Curtis E., and Alan H. Strahler, 1987. The factor of scale in remote sensing. Remote Sensing of Environment, 21 : 311-332. 123 IlullllljljlIljlllljllljljjll