PRECISION FORESTRY PROCEEDINGS OF THE SECOND INTERNATIONAL PRECISION FORESTRY SYMPOSIUM UNIVERSITY OF WASHINGTON COLLEGE OF FOREST RESOURCES FERIC, THE FOREST ENGINEERING RESEARCH INSTITUTE OF CANADA IUFRO, THE INTERNATIONAL UNION OF FOREST RESEARCH ORGANIZATIONS USDA FOREST SERVICE, PACIFIC NORTHWEST RESEARCH STATION SEATTLE, WASHINGTON JUNE 15-17, 2003 PRECISION FORESTRY PROCEEDINGS OF THE SECOND INTERNATIONAL PRECISION FORESTRY SYMPOSIUM Printed in the United States of America All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopy, recording, or any information storage or retrieval system, without permission in writing from the publisher, the College of Forest Resources. Institute of Forest Resources College of Forest Resources Box 352100 University of Washington Seattle, WA 98195-2100 (206) 685-0887 Fax: (206) 685-3091 http://www.cfr.washington.edu/Pubs/publist.htm Proceedings of the Second International Precision Forestry Symposium, sponsored by the University of Washington College of Forest Resources, the Precision Forestry Cooperative, Seattle, Washington, FERIC, the Forest Engineering Research Institute, Vancouver, BC, IUFRO, The International Union of Forest Research Organizations, Vienna, Austria, and the USDA Forest Service, Pacific Northwest Research Station, Resource Management and Productivity Program, Portland, Oregon. Additional copies of this book may be purchased from the University of Washington Institute of Forest Resources, Box 352100, Seattle, Washington 98195-2100. For additional information on the Precision Forestry Cooperative please visit http://www.precisionforestry.org II TABLE OF CONTENTS Acknowledgments Preface Keynote Speakers VI VIII Opening Remarks and Welcome to the First International Precision Forestry Symposium B. Bruce Bare 1 Precision Forestry – The Path to Increased Profitability! Bill Dyck 3 Precision Technologies: Data Availability Past and Future Daniel L. Schmoldt and Alan J. Thomson 9 Plenary Session A: Precision Operations and Equipment Moderator, Alex Sinclair Multidata and Opti-Grade: Two Innovative Solutions to Better Manage Forestry Operations Pierre Turcotte A Test of the Applanix POS LS Inertial Positioning System for the Collection of Terrestrial Coordinates Under a Heavy Forest Canopy Stephen E. Reutebuch, Ward W. Carson, and Kamal M. Ahmed 17 21 Ground Navigation Through the Use of Inertial Measurements, a UXO Survey Mark Blohm and Joel Gillet 29 Precision Forestry Operations and Equipment in Japan Kazuhiro Aruga 31 Precision Forestry Applications: Use of DGPS Data to Evaluate Aerial Forest Operations Jennie L. Cornell, John Sessions and John Mateski 37 Plenary Session B: Remote Sensing and Measurement of Forest Lands and Vegetation - Moderator, Tom Bobbe Estimating Forest Structure Parameters on Fort Lewis Military Reservation Using Airborne Laser Scanner (LIDAR) Data Hans-Erik Andersen, Jeffrey R. Foster, and Stephen E. Reutebuch 45 Developing “COM” Links for Implementing LIDAR Data in Geographic Information System (GIS) to Support Forest Inventory and Analysis Arnab Bhowmick, Peter P. Siska and Ross F. Nelson 55 Large Scale Photography Meets Rigorous Statistical Design for Monitoring Riparian Buffers and LWD Richard A. Grotefendt and Douglas J. Martin 61 Forest Canopy Models Derived from LIDAR and INSAR Data in a Pacific Northwest Conifer Forest Hans-Erik Andersen, Robert J. McGaughey, Ward W. Carson, Stephen E. Reutebuch, Bryan Mercer, and Jeremy Allan 65 III Enhancing Precision in Assessing Forest Acreage Changes with Remotely Sensed Data Guofan Shao, Andrei Kirilenko and Brett Martin 67 Automatic Extraction of Trees from Height Data Using Scale Space and SNAKES Bernd-M. Straub 75 A Tree Tour with Radio Frequency Identification (RFID) and a Personal Digital Assistant (PDA) Sean Hoyt, Doug St. John, Denise Wilson and Linda Bushnell 85 Plenary Session C: Terrestrial Sensing, Measurement and Monitoring Moderator, Steve Reutebuch Value Maximization Software – Extracting the Most from the Forest Resource Hamish Marshall and Graham West 87 Costs and Benefits of Four Procedures for Scanning on Mechanical Processors Glen E. Murphy and Hamish Marshall 89 Evaluation of Small-Diameter Timber for Value-Added Manufacturing – A Stress Wave Approach Xiping Wang, Robert J. Ross, John Punches, R. James Barbour, John W. Forsman and John R. Erickson 91 Early Experience with Aroma Tagging and Electronic Nose Technology for Log and Forest Products Tracking Glen Murphy 97 Plenary Session D: Design Tools and Decision Support Systems Moderator, Glen Murphy Modeling Steep Terrain Harvesting Risks Using GIS Jeffrey D. Adams, Rien J.M. Visserm and Stephen P. Prisley 99 Use of the Analytic Hierarchy Process to Compare Disparate Data and Set Priorities Elizabeth Coulter and Dr. John Sessions 109 Use of Spatially Explicit Inventory Data for Forest Level Decisions Bruce C. Larson and Alexander Evans 115 Elements of Hierarchical Planning in Forestry: A Focus on the Mathematical Model S. D. Pittman 117 Update Strategies for Stand-Based Forest Inventories Stephen E. Fairweather 119 A New Precision Forest Road Design and Visualization Tool: PEGGER Luke Rogers and Peter Schiess 127 Harvest Scheduling with Aggregation Adjacent Constraints: A Threshold Acceptance Approach Hamish Marshall, Kevin Boston and John Sessions 131 Preliminary Investigation of Digital Elevation Model Resolution for Transportation Routing in Forested Landscapes Michael G. Wing, John Sessions and Elizabeth D. Coulter Comparison of Techniques for Measuring Forested Areas Derek Solmie, Loren Kellogg, Michael G. Wing and Jim Kiser IV 137 143 Posters and Abstracts Can Tracer Help Design Forest Roads? Abdullah E. Akay 150 CPLAN: A Computer Program for Cable Logging Layout Design Woodam Chung and John Sessions 150 List of Contributors 151 List of Attendees 155 Second International Precision Forestry Symposium Agenda 160 V ACKNOWLEDGMENTS Many individuals and organizations contributed to the success of this symposium and the collection of papers in this volume. The conference was sponsored by the University of Washington College of Forest Resources, the Precision Forestry Cooperative, the USDA Forest Service, Pacific Northwest Research Station, Resource Management and Productivity Program, Portland, Oregon, FERIC, the Forest Engineering Research Institute of Vancouver, BC, Canada, and IUFRO, the International Union of Forest Research Organizations of Vienna, Austria. The program was planned by a committee consisting of: Chair Professor David Briggs, College of Forest Resources, University of Washington Scientific Sub-Committee: Hans-Erik Andrsen, College of Forest Resources, University of Washington David Briggs, University of Washington Ward Carson, USDA Forest Service, PNW Research Station Megan O’Shea, University of Washington Steve Reutebuch, USDA Forest Service, PNW Research Station Professor Gerard Schreuder, Acting Director, Precision Forestry Cooperative This committee developed the program and recruited authors for the topics presented. The lead authors, in turn, worked with coauthors and consulted with others to make this a truly international effort. The time and effort of all these contributors resulted in excellent presentations and posters. Papers were reviewed before acceptance for publication, and the input of the many reviewers is much appreciated. The session moderators Alex Sinclair, Vice President, FERIC Western Division, Vancouver, BC, Tom Bobbe, Remote Sensing Applications Center, USDA Forest Service, Salt Lake City, UT, Steve Reutebuch, Team Leader, Silviculture and Forest Models Team, USDA Forest Service, Pacific Northwest Research Station, Seattle, WA, Glen Murphy, Professor, Forest Engineering, Oregon State University, Corvallis, OR, and B. Bruce Bare, Rachel B. Woods Professor of Forest Management, and Dean, College of Forest Resources, University of Washington provided program linkage and kept the conference on schedule. VI We would also like to thank the Precision Forestry Board of Directors for their support in the early planning of the symposium. Chair, Rex McCullough, Weyerhaeuser Company Wade Boyd, Longview Fibre Craig D. Campbell, Boise Cascade Corporation David Crooker, Plum Creek Timber Suzanne Flagor, Seattle Public Utilities Watersheds Division Sherry Fox, Washington Forest Protection Association John Gorman, Simpson Investment Company Peter Heide, Washington Forest Protection Association Edwin Lewis, Bureau of Indian Affairs Yakima Agency John Mankowski, Washington State Department of Fish and Wildlife John Olsen, Potlach Corporation Charley Peterson, USDA Forest Service Mike Renslow, Spencer Gross, Inc. Bryce Stokes, USDA Forest Service David Warren, The Pacific Forest Trust Laurie Wayburn, The Pacific Forest Trust Maurice Wiliamson, Forestry Consultant Megan O’Shea of the University of Washington College of Forest Resources was responsible for conference arrangements and management. Andrew Cooke edited and produced the proceeding CD. Their efforts and those of the registration workers, projectionists, and other volunteers were critical for the smooth operation of the conference and are greatly appreciated. Production of this volume was coordinated through the Institute of Forest Resources; a special thanks to John Haukaas for his assistance in editing. Megan O’Shea handled final editing and publication preparation. The capable efforts of this skilled group are gratefully acknowledged. The proceedings are reproductions of papers submitted to the organizing committee after peer review. We wish to acknowledge the efforts of all the scientists involved in the peer reviews of these proceedings papers. No attempt has been made to verify results. Specific questions regarding papers should be directed to the authors. David Briggs, Chair Second International Precision Forestry Symposium VII PREFACE The need for precision forestry is no longer a choice in managing forest and producing forest products. Driven by both the ever increasing scrutiny over the protection of forest resources, and the economic need to use forest products to the fullest, professional foresters and product managers are demanding quality detailed information about forests they manage and products they make. I am confident that the presentations and discussion we have in the next few days will lead to the implementation of technologies that will move forestry to a higher level of information resolution. Please take note of the fine corporate exhibitors featured on the following page. I am grateful for their participation in this symposium. I want to give special thanks to the College of Forest Resources faculty, staff and students who worked on the Symposium Planning Committee, as volunteers or scientific reviewers. Gerard Schreuder Acting Director, Precision Forestry Cooperative VIII Opening Remarks and Welcome to the First International Precision Forestry Symposium B. BRUCE BARE, DEAN, COLLEGE OF FOREST RESOURCES, UNIVERSITY OF WASHINGTON WELCOME The College of Forest Resources, University of Washington is pleased to host this second international symposium dedicated to Precision Forestry. We hope your participation and ideas will help focus attention on innovative technologies and approaches to guide the future of forestry and the forest industries in Washington State and elsewhere. A few words about our College. HISTORY OF ADVANCED TECHNOLOGY INITIATIVE (ATI) • • • The UW’s Precision Forestry Cooperative is one research cluster funded by the State’s Advanced Technology Initiative (ATI). The ATI is a partnership between the Legislature, private industry, and the research universities of the State of Washington. Washington State Legislature funded six Advanced Technology Initiatives during the 1999/2001 biennium. HISTORY OF ATI • • Each ATI “cluster” is expected to generate new industries or transform existing industries of importance to Washington State. And, each “cluster” is a bridge between research, education, and new economic activity. New leaders are being educated to help transform the industries vital to the State’s economic future. PRECISION FORESTRY • • • • Employs high technology sensing and analytical tools to support site-specific, economic, environmental, and sustainable decision-making for the forest sector. Provides highly repeatable measurements, actions, and processes to grow and harvest trees, as well as to protect and enhance riparian areas, wildlife habitat, esthetics, and other environmental resources. Provides valuable information and linkages between resource managers, the environmental community, manufacturers, and public policy. Links the practice of sustainable forestry and conversion facilities to produce the best economic returns in an ecologically and socially acceptable manner. INNOVATIVE TECHNOLOGIES • • • • • • GPS, GIS for precise ground measurements Remote sensing (LIDAR, INSAR) Wireless systems Real-time process control scanners Visualization Decision support systems (integrated data systems) 1 PRECISION FORESTRY COOPERATIVE FOCUS • • • • Decision Support Systems Remote Sensing and Geospatial Analysis Silvicultural and Ecological Engineering Precision Operations and Terrestrial Sensing • To develop tools and processes that increase the precision of forest data to support better decisions about forests — their services and products, through a collaborative effort with private landowners, public agencies, manufacturers, and harvesters. PRECISION FORESTRY COOPERATIVE GOAL PRECISION FORESTRY SYMPOSIUM • • • Brings scientists, managers, and developers together to work collaboratively. Will provide insights into the current “state of the art” and provide a springboard for new ideas and innovations. We hope you enjoy the symposium, the campus, and the city during your stay with us. 2 Precision Forestry – The Path to Increased Profitability! BILL DYCK Abstract: The market wants good wood and the forest industry wants to see greater profitability. Precision Forestry has a role to play in both developing tools to find the best wood in existing forests and trees, and also in providing the knowledge to grow better wood in the first place. New technologies are being developed that can help us evaluate forests at a macro-level, enhance our ability to estimate stand volumes, and even measure the properties of individual trees and logs. These tools should lead to greater profitability as higher value wood can be allocated to higher value markets. Increased profitability can also be achieved by understanding the interactions of genetics, site and silvicultural management to grow more valuable forests. INTRODUCTION perhaps an optimistic sceptic, when it comes to the development and application of technology for the forest industry. I learned early on that ideas are cheap, there is relatively little that is really new patenting is often a costly waste of time and money, and that implementation is everything. Therefore, I want to start off by making one point with regard to Precision Forestry research and technology and its application to industrial forestry: The term “precision forestry” means different things to different people. To a geneticist it probably means precisely matching the genetics of a tree species to the site to maximise growth. To an industrial forester it might mean precisely managing a forest to match what the market needs. But, to a conservationist it probably means being able to precisely manage a forest to optimise environmental benefits. What the website for this symposium said was that: “Precision Forestry uses high technology sensing and analytical tools to support site-specific, economic, environmental, and sustainable decision-making for the forestry sector. It provides for highly repeatable measurements, actions, and processes to initiate, cultivate, and harvest trees, as well as enhance riparian zones, wildlife habitat, and other environmental resources. It provides valuable information linkages between resource managers and processors. The Symposium will bring together scientists to present state-of-the-art information on topics such as precision sensing techniques, operations-sensing techniques and their use for decision-making.” What this meant to me was that the audience was going to be interested in a wide range of topics all designed to improve the precision by which we manage forests, whether it is for commercial, environmental, or social benefits. A keynote paper is supposed to be thought provoking and generally delivered at a fairly high level to set the scene for the rest of the meeting. I’m going to attempt to do that, but will focus on one side of forestry – “industrial forestry” – i.e., that part of forestry that seeks to make money from trees, and I’ll look at how I think Precision Forestry can improve profitability. I have been in the Science & Technology game for over 25 years and consider myself to be a sceptical optimist, or Get the market to provide the lead. Technology driven research is almost certainly doomed to fail. The main objective of this presentation is to give you my views on where precision forestry technology can play a role in the industrial sector and specifically how it can help the forest industry become more profitable. HOW PRECISE DOES FORESTRY NEED TO BE? When I started out in forestry, and even relatively recently, there used to be an expression commonly used by foresters: “close enough for forestry”. What that really meant was that in forestry you didn’t have to be very precise, after all, forestry was just cutting down trees and getting them to a sawmill where logs were made into lumber and shipped off the to market; generally a pretty crude business. The business has changed, primarily as logs have become more valuable and cost cutting puts the squeeze on operations. But, how precise do we really need to be? A tree is a tree is a tree! At least if it is the same species, the same size, and the same shape it should be, correct? However, that is not the case. All logs are different even if they are 3 priate technology would have saved his business. Back to the question “how precise does forestry need to be?” The answer is “it depends” and it mainly depends on the market being targeted. One of the main reasons that forestry is complicated and there is a need for greater precision is the enormous complexity of trees. After decades of research we still fail to understand some of the fundamental principles of tree growth and wood. We can certainly grow big trees and quickly, but we don’t fully understand the linkages between growing trees and creating high value wood. While this complexity creates problems, it also creates opportunities, at least for those who take the time and effort to really understand the nature of trees. I suggest there are two paths if followed that will make forestry more precise and lead to greater profitability: (1) know what you’ve got, (2) grow what the market wants. clonal and even if they come from the same tree. In Figure 1 several hundred logs from two radiata pine plantation forests in New Zealand were selected for similar grade and tested for sound speed, a measure of intrinsic wood stiffness and other properties. The results were very revealing as there was wide variability in wood properties, from similar looking logs. The prices paid for the logs were all the same, but the value of the structural lumber from the fastest and stiffest logs was much greater than the industrial grade from the slower logs. Of course it is the forest owner that is missing out on this premium! But it is also the mill owner that is wasting resources processing inferior logs in an attempt to make premium products. Frequency 120 PATH 1 – KNOW WHAT YOU’VE GOT 80 Regardless of whether the market is for forests, standing trees, logs, lumber, or fibre, you need to know what you’ve got and where it is. This path should perhaps therefore read, “Know what you’ve got – in terms the market values” The forest market wants to know where the forests are, how big they are, and what’s in them. It also wants to know the “risk” – how healthy are the forests, what’s their nutritional status, and are there potential liabilities associated with high value conservation areas, endangered species habitat, or cultural sites that need to be protected. We are reasonably good at valuing forests on a very broad basis, but we’re not all that good at rapidly determining risk values, such as nutritional and health status. The tree or stumpage market would like to know much more precisely then what we are currently able to determine both the volume that is in the forest and the value. Ideally, it would also like to know just how variable the quality is in the stand, both within and between trees. The log market wants to know volumes by external log grade (log dimensions, sweep, knot size and spacing) but it is also starting to ask for more than this – hence the crude measure of strength in some log markets of “rings per inch”. Ideally we would like to be able to match specific logs to specific markets, both for lumber, veneer and even chips, but we are only just starting to make progress in this area. 40 4 3.8 3.6 3.4 3.2 3 2.8 2.6 2.4 0 Speed, Km/s Figure 1: The sound speeds (km/s) of a large sample of similar logs from two geographically distinct radiata pine forests in New Zealand demonstrating the large variability in the intrinsic wood properties of the logs. (Industrial Research Ltd data). There is another expression that I’ve heard more recently and that is perhaps more relevant. “Forestry isn’t rocket science, it’s more complicated than that!” I’m not sure who coined the phrase but I believe it is appropriate. The results in Figure 1, and in fact the underlying technology underpinning what is now a commercial tool, is the work of an ex-space physicist Dr Mike Andrews, currently working at Industrial Research Ltd in New Zealand. Clearly, a more precise grading system for the radiata pine logs shown in Figure 1 would have seen greater log segregation based on intrinsic wood properties and greater price differential in the market. However, a word of caution, for Precision Forestry technologies to be useful we need to be careful that they don’t over complicate the business of forestry, or the winners will be concrete and steel. But, on the other hand, new log and wood segregation technologies can play a big role in protecting wood’s place in the market by providing better quality control and product assurance for wood products. There was an example during the Sydney Olympic construction days when a very large laminated beam failed in use because the manufacturer had used low strength components, although they looked just as good as previous material he had used. In this case the application of appro- Technologies to tell us what we have: Seeing the Trees from the Sky Satellite technologies have been very disappointing, at least for industrial forestry applications, and to my knowledge there have been very few examples of satellite technology improving forest management and helping to increase revenue flow. Aerial photography from planes and helicopters, on the other hand, has been the workhorse of remote 4 ance on “seat-of-the-pants forestry”. In the absence of reliable technology, local knowledge and experience becomes extremely important for estimating the inherent wood properties and hence the value of stands and trees. That reliance is changing as new tools become available to assess stands for wood properties. Silvascan-2 developed by Rob Evans and his team at CSIRO in Australia, has proved to be an extremely invaluable tool for improving our evaluating wood properties. This technology can measure fibre properties from an increment core up to 1000 times faster than traditional lab-based methods. The tool is especially useful for measuring microfibril angle, which in the past has been expensive and somewhat unreliable, as well as for determining other cellular properties that translate into useful market values. Many forestry companies are now using Silvascan-2 to improve their inventory assessments of wood properties and values by analyzing increment cores from selected trees. Director (also known as Hitman), a technology developed by IRL in New Zealand and owned and marketed by CHH FibreGen, is being used to determine the structural properties and by inference the value of logs (Figure 3a). This technology is based on time-of-flight sonics and has been demonstrated to reliably predict the average stiffness of lumber produced in logs. Because the MOE of the log is simply equal to density times the speed of sound squared, the technology is basically measuring fibre properties that influence macro properties such as stiffness, strength, and stability. The challenge is to interpret what the log is “saying” and translate this information into meaningful values (Figure 3b). Director is currently being used to identify resource stiffness by stand and by forest, and to a lesser extent to segregate individual logs for high value structural processing, mainly LVL. The future is the development of technology to cost-effectively assess the properties of standing trees and thereby greatly improving value estimates of stands and forests, of particular interest for the stumpage or forests market. Research is currently focused on hand held tools to measure the density and stiffness of trees. sensing. More recently in this area we’ve seen tree counting algorithms developed that enable automatic determination of the number of trees per hectare from digital imagery, and also much better forest boundary identification than what has been possible in the past. There have been new developments in remote sensing technology that show potential for industrial forestry. I’m particularly excited about the promise of hyperspectral imagery for assessing disease infestation and nutrient deficiencies in production forests. Researchers in CSIRO (Nicholas Coops in press), Australia have demonstrated the application of airborne hyperspectral imagery for the assessment of Dothistroma a needle blight disease of radiata pine (Figure 2) and plans are underway to launch this as a commercial service, thus enabling more rapid and more accurate detection of the disease. As well as showing promise for monitoring forest health, it appears that the technology appears to also have application for determining tree species, the nutritional status of forests, and monitoring the spread of weeds. Healthy Un-healthy Figure 2. Hyperspectral image of Dothistroma infection in radiata pine. (N. Coops et al, CSIRO in press). Seeing the Trees from the Ground While being able to measure everything remotely is the dream, we still need to measure trees and forests from the ground. Traditional ground-based forest inventories give a reasonable estimate of tree volumes by species and to some extent external log grades, but as a rule we tend to be rather poor at estimating the true value of stands. New technologies are coming onto the market that will change all this. Instead of simply estimating log grades, laser tree-profiling technology collects digital images of trees that can be fed into optimising software to predict values as well as volumes per hectare. Currently this is too expensive to be used as more than just an audit tool but it does point to the future. However, what is really exciting is our increasing ability to see into trees and quantify some of the more valuable intrinsic properties such as density, stiffness and specific fibre properties. Wood is a very complex biomaterial that is poorly understood by forest managers and scientists alike, hence the reli- Figure 3a: Application of Director sonics technology to pine logs (IRL photo). 5 somewhat understandable in that until recently we haven’t been able to rapidly measure wood properties, but all that’s changing and we no longer have the lack of tools as an excuse. We are now entering what I consider to be the fourth stage of industrial forestry – High Performance Wood (Figure 4). Stage 1 consisted simply of felling old growth natural forest and processing the logs into lumber. Stage 2 was the start of plantation forestry in which vast areas of trees were planted, often to replace dwindling supplies from natural forests. In Stage 3 we started to get more sophisticated and practiced more intensive silviculture resulting in improved genetics, faster growth, and generally fatter trees by an early age, but the focus was simply on what the trees looked like and had nothing to do with wood quality. Stage 4 is what I optimistically refer to as the stage of “high performance wood”, and this is where Precision Forestry comes in. Precision Forestry for growing better wood that is. I do not accept the argument that we cannot predict what the market wants 25 years out, or even 100 years out. If we look back at what the market for wood products has wanted for the last 1000 years it has been for strong and stable material, and for some applications, attractive wood. Getting a bit of durability is a bonus, but if we focus on strong and stable wood then we have to be on the right track. We could add to this list with a few other obvious parameters, such as defect-free (internal checks) and blemish-free wood (resin pockets etc). For some fibre applications we will want strong, coarse fibres, whereas for others we want short fibres and often fibres that will collapse to give a soft finish. But, let’s keep this simple and focus on solid wood products. Our inability to reliably produce strong, stable, and attractive wood at a reasonable cost is at least partially responsible for the introduction of substitute products, including wood composites. So, where does Precision Forestry come into growing a good crop of trees, or better stated, growing good wood? It comes in everywhere, starting with genetics and ending with Of course, having managed to precisely locate your forests and determine what is in the trees, you then need to ensure you extract maximum value, which gets into log processing technology. That is a whole new subject; so instead of going forward down that path, let us go back to the beginning – growing what the market wants. 1 Site G11 Stem #4 25.8m Amplitude 0.75 0.5 0.25 0 0 100 200 300 400 500 600 700 Frequency, Hz Figure 3b: Sonics trace from a radiata pine stem. PATH 2 – GROW WHAT THE MARKET WANTS The market wants good wood! We now know how to grow big trees quickly, but we have yet to determine how to reliably produce good wood. Critics of this statement claim that the definition of “good wood” depends on the end use, and while this is true to a certain extent, we can definitely state what constitutes “bad wood”. If we don’t know what good wood is, or at least understand what we don’t want in wood, then we really haven’t got much hope in growing what the market wants. For decades forest growers have focused on very unsophisticated markets – the log market and the tree market (or the market for forests). Consequently we’ve either strived to grow volume per hectare or volume per stem. Other than branch size and straightness, there has been relatively little focus on wood quality. Even worse, in some countries, especially where the forests are government owned, there has only been a focus on getting a new crop started and above the weeds with little attention to where the final harvest might end up. What we should really be asking of course is, what does the market want and how do we go about growing what the market wants. We need to put a lot more thinking into growing wood than we have in the past. Stage 1 – Old growth forests Stage 2 – Plantations Stage 3 – Growing big and straight I’ve yet to meet a forester who can tell me the formula for growing good wood. Stage 4 – High performance wood Geneticists who believe that genetics is the answer to everything have taken us for a ride down the wrong forest path. And for the most part we’ve basically ignored the influence of site and management on wood properties. This is Figure 4: The four stages of industrial forestry. Timing will vary by country. 6 harvesting. In fact, it goes back to the molecular level and understanding how wood cells respond to site and management stimulus. The reason that some NZ radiata pine is “trash” and treated as such in some markets, is not because there is anything wrong with the species, it’s the way we’ve grown some of our forests. The so-called “S-diagram” in Figure 5 provides a framework to indicate why we need to be more precise when growing trees. The key is to have a reasonable understanding of what the market wants, and then to have a much better handle on how genetics, site, and management impacts on what is produced. est), but also in how we select our sites and manage our trees. In fact we don’t really even attempt to manage trees, but we tend to manage stands and forests. The best pruned radiata pine stands in New Zealand are worth twice the value of the average pruned stands, and the reason is a combination of genetics, site, and management practices. It is the influence of both the site and management of the individual trees that results in the differences in wood quality that we get within a forest. We are only just starting to really understand how much genetics pre-determines wood quality, and that trees growing next to each other on basically the same site and with the same management, will produce very different wood. A Move to Genotype Forestry? One way to overcome the effect of genetics on variable wood properties is to use genotype (clonal) forestry. This is done for short rotation pulp and paper hardwood crops and is starting to be employed for longer-rotation conifers. However, while this will certainly reduce variability, there is no guarantee in my mind that it will lead to higher value forests as I’m not convinced that we have even begun to understand the relationship between genetics and wood quality. The promises of molecular biology and tools such as “marker-aided selection” are there, but are they real or are they just hype? Choosing the wrong genotype (i.e., clone) can have disastrous results unless we are 100% certain that we have gotten everything correct, not just the one trait that we might be selecting for. We see this in our genetics programmes where we’ve focused on volume and form and have had virtually no understanding how selecting for these traits would affect other features that are actually more important for the ultimate wood market. Figure 5:The quality of the lumber and fibre products derived from a tree is dependent upon the ultrastructure and molecular properties of the wood cells, which are in determined by a combination of genetics, site, and management. (S-diagram from University of Canterbury). A Move to Site-specific Forestry? I suggest that we can make more progress producing what the market wants, i.e., good wood, by moving to more sitespecific forestry. I do not believe that we are ready to match genotypes to site, but we can certainly match families to site and avoid some of the more serious impacts of disease, water logging, certain wood quality defects etc. and also make gains in productivity. We can also begin to be more precise in managing sites to produce better trees and better wood by first of all understanding the effects of soils and climate on wood properties. We can also be much more precise in how we manage weed competition by, for example, careful chemical selection and precision application, and in how we manage nutrition, which should be on a site basis, rather than a stand basis. There is also a need for much better understanding as to the impacts of silvicultural interventions, such as pruning Producing good wood products that the market wants is very similar to producing good wine. It’s getting the combination of genetics, site selection, and management regime just right, and then of course processing the grapes in the best possible way. Certainly genetics is important to wine, hence we are able to make choice of a cab sav or a sauvignon blanc depending on our mood at the time. But, as any wine drinker knows, it’s possible to buy a very good cab sav and also a very poor one. What makes the difference? That really comes down to site – particularly soil and climate – and then to management – how the vines were managed to optimize the quality of the grapes. Skill in processing good grapes is also very critical of course, but as wine makers have told me, “anyone can make a good wine in a good vintage year”. In forestry we tend to be very imprecise not only in selecting genetics (we have tended to choose what grows fast7 Forestry research and technology developments have not been all that good at really understanding what the market wants, but it has not necessarily been all our fault. Often we ask for input, but we ask the wrong people or we ask the wrong question, and therefore we get the wrong answer. Or we get the correct answer but we do not know enough to provide the solution. There is little doubt in my mind that what the market for industrial forestry really wants is good wood. We have two ways to produce this good wood (1) find it in our existing forests, and (2) grow it in the first place. To become more profitable we need to better understand what wood is, particularly what good wood is, what key properties we need to measure in all stages of the value chain, and we need to understand what this means to the end user. We then need to develop tools that can help us to make these assessments, but we have to be able to implement this technology in such a way that the costs do not outweigh the benefits. Precision Forestry is required in both enhancing our ability to “know what we’ve got” and also in understanding how to “grow what the market wants” as we need research and technology to understand what is in the forest, right down to the tree and log level, and we need to be much more precise matching genetics with site and silvicultural management. A greatly improved ability to know what we have got and to grow what the market wants will lead to greater profitability, provided we can do all this cost effectively. and thinning, shelter wood management etc on final wood quality. It appears that thinning may have a detrimental effect on wood quality as it stimulates cell growth producing steeper microfibril angles in the secondary cell walls leading to reduced stiffness and stability. It also appears to be responsible for increasing the amount of compression wood in a stem, possibly a response to greater wind movement in the stand. The Challenges are There! Forestry needs to focus on genetics, site, and forest management practices that will produce the best cells, which in turn will lead to the best wood. Sounds difficult? You bet it is! The alternative is “hit and miss forestry” which in many cases will lead to reasonable wood, in some cases be a total disaster, and if we are really lucky, will lead to really good wood that the market can’t get enough of. Ironically, in New Zealand, the best wood that I know of has come from untended fire-regenerated stands of radiata pine that were harvested at age 50. Clearly there is a role for Precision Forestry to focus on the underlying mechanisms that influence wood quality, as ultimately, the market that we are targeting is not simply demanding better forests, but it is demanding better wood or it will turn to substitutes. CONCLUSIONS This is a keynote paper so I will I wrap up with a couple of salient lessons for Precision Forestry, and to do this I want to go back to the point that I made at the start of this presentation: ACKNOWLEDGEMENTS Several people have provided input to this paper and I particularly thank Mike Andrews (IRL), Nicholas Coops (CSIRO), Peter Carter (CHH), Rick Walden (Smart Forests), and Brian Rawley. Get the market to provide the lead. Technology driven research is almost certainly doomed to fail. 8 Precision Technologies: Data Availability Past and Future DANIEL L. SCHMOLDT AND ALAN J. THOMSON Abstract:Current precision and information technologies portend a future filled with improved capabilities to manage natural resources with greater skill and understanding. Whereas practitioners have historically been data limited in their management activities, they now have increasing amounts of data and concomitant sophistication in data management, analysis, and decision tools. Expanding precision forestry technologies beyond traditional reliance on optics-based tools offers new opportunities for forest resource interrogation. However, as data become more immediate and information rich, traditional views of data availability may lose some relevance. Technical constraints are becoming less daunting and social and ethical responsibility and sensitivity are gaining prominence. Because data that might be deemed private or protected can be readily moved and combined with other data, new concerns arise about who uses those data and how they use them. Capabilities built into newer analysis and decision support tools add further apprehension about privacy, accuracy, and accessibility. It does not require an extraordinary string of suppositions to imagine when regulation and legal decisions will promulgate certain safeguards for data management and for software that handles data. Such restrictions could likely limit data availability in currently unforeseen ways—counteracting, to some extent, technology-based advances in data availability. Still, irrespective of those possibilities, there are actions that natural resource professionals can take to lessen potential future restrictions on data availability. These include defining an “information space” for each precision technology, understanding language and knowledge flows, and planning for integrated systems and processes that holistically address information needs and uses. INTRODUCTION of habit. Once bacterial counts become part of our everyday information environment, though, we have to alter our consciousness to incorporate a more “dirty-aware” reality that accepts our existence with microbes. Similarly, a fire manager, presented with large amounts of real-time data and the capability to manipulate it, begins to see the landscape in a truly dynamic way. Now, decisions that he or she makes can be continually updated, or tweaked, as conditions change. Dynamic decision making creates increased confidence and control for the manager, and minimizes the likelihood that field judgments will be questioned later. In fact, decision support systems can track the decision making process for subsequent audit. Not only are more data available more often, but the time between measurement and application is shrinking rapidly. Whereas, at one time, field crews collected volumes of information on the ground, recorded data with pencil and paper, and entered data into a computer back in the lab for analysis, it is now possible to collect more spatially dense data much faster without going into the field, in some instances. The former process could take many days (or weeks) for relatively low resolution, while current technologies can potentially reduce the time to just hours. Such just-in-time information promises to bring decision making out from behind the computer display and into the field (e.g., Clark 2001). Here, then, managers and field operators can react One of the prominent thrusts in agriculture, food, and natural resource systems brings increasingly data-rich environments into everyday use. Consumers, for example, might soon be able to scan a package of chicken in the refrigerator and know exactly where the product was grown and processed, and what its current shelf-life is based on bacterial counts (Pathirana et al. 2000). In other cases, land managers might have real-time information about fuel loads across large geographic areas and simulate a large number of hypothetical ignition scenarios based on 24-hour weather forecasts. For sustainably grown timber, chain-of-custody verification might rely on programmable identification devices (Simula et al. 2002) or chemical markers (q.v., companion article in this volume). Significant scientific and technical hurdles still remain and modifiers, such as “soon” and “large number,” are as yet undefined, but theoretically there is nothing to prevent either scenario from becoming reality, as feasibility is well established in both cases. These data have the capacity to tell us more about the world in which we live and work, and also can alter our professional, and emotional, viewpoints of that world and how we interact with it. In the examples above, we don’t currently give much thought to bacterial counts on the food we eat, although we wash food, such as chicken, as a matter 9 more quickly to changing conditions and have a broader, more informed picture of the resource being managed. The advantages of high spatial and temporal data resolution for researchers and practitioners are obvious, so data volume and rapidity have been primary scientific thrusts. However, we are entering a phase where subtle shifts are occurring in what “data availability” means. While, there are still many forest and forest product characteristics that we would like to measure and apply effectively, the technical hurdles to doing so are not insurmountable. Those previous science and technology limits to data availability may soon be supplanted with other availability issues, such as data-use policy and legal restrictions. Then, the issue becomes not one of technically capable information technology (IT), but rather one of human-centric IT (Schmoldt 2001). That is, how well these information tools fit within organizational and social cultures and how well they reflect the users’ ethical standards and expectations. Just because physical hurdles to data generation have been reduced does not mean that limitations on data application, based on ethical concerns (Thomson and Schmoldt 2001), won’t be equally problematic. In the sections that follow, we describe and illustrate the three phases of precision technologies: basic research; engineering and technology development; and application and adoption. While most of the companion papers in this volume deal with the latter two phases, the basic research phase cannot be completely ignored as it provides the scientific basis for a technology’s capabilities and limitations. The second issue addressed in this paper is the growing importance of ethics in data collection and data use. There needs to be awareness by scientists and practitioners regarding ethical standards of conduct and how they may dictate development and use of precision technologies. which may not necessarily reflect biological realities. These and other issues have hindered data availability and application in the past and continue to present challenges for some recent technologies. Still, as the papers in this volume and cited works elsewhere demonstrate, forest science and management have increased access to data, collection frequency, and possess more powerful tools to manipulate the data. PRECISION TECHNOLOGIES AND CURRENT DATA AVAILABILITY Technologies, i.e., tools, processes, and materials, ensue from scientific discovery. Biophysical and chemical phenomena must first be understood before they can be translated into useful devices and products. For example, the optical properties of the atmosphere and plants, and the physics of collecting light at great distances, must be known before remote sensing makes sense. Similarly, the mathematics of optimizing constrained production functions must be developed before solution algorithms can be written. These scientific developments provide fundamental knowledge for subsequent, possibly unforeseen, technologies. In some research settings (e.g., a university or federal laboratory), end uses for science endeavors may not always be immediately apparent; neither are they necessarily offered as justification for the research. In other cases, a longterm goal (process, device, product) drives the science, with a proof-of-concept targeted as the immediate research objective. The latter is more common in the private sector. In relatively few cases, however, do agriculture and forestry applications drive research efforts related to precision technologies. Once various precision technologies have been developed, though, they often find ready application to re- Definition Before proceeding further, it is important to provide a definition for the broad area of “precision technologies.” For most intents and purposes inherent in this paper, the following should suffice: Instrumentation, mechanization, and information technologies that measure, record, process, analyze, manage, or actuate multi-source data of high spatial and/ or temporal resolution to enable information-based management practices or to support scientific discovery This definition applies equally well to technologies that might be employed in agriculture, food, and environmental systems. While the definition doesn’t explicitly state so, biophysical, chemical, and engineering sciences provide the bases for these technologies, and information technologies (IT) often provide the application mode—although, in some cases practices are realized through the use of electro-mechanical devices driven by microprocessors to actuate a response. Basic Research Introducing precision technologies into forest environments is difficult for many reasons. First among those are scale issues. Our measurements must be possible at spatial scales in the millimeter range (nitrogen fixation in the soil) and also at the kilometer range (stand health, stand timber volume). Events occurring over short time periods (e.g., stomatal aperture) can be equally important to much longerperiod phenomena (e.g., tree diameter growth). Second, there is tremendous variability over time and space when repeatedly measuring the same phenomenon. While this creates problems for taking consistent measurements, our ability to take frequent measurements helps us understand that variability and better deal with it. Third, most of our measurement modalities to date have relied on optics, which limits our observations to line-of-sight interrogation. Fourth, when taking measurements at finer spatial and temporal resolutions, we then often aggregate those data—in our models and decision support systems—to, somewhat arbitrary, coarser resolutions that suit anthropocentric needs, 10 of communicating MEMS that can measure ecological variables across an entire watershed, for example. Convergence of the Internet, communications, and information technologies with techniques for miniaturization has placed sensor network technology at the threshold of a period of major growth. Emerging technologies can decrease the size, weight, and cost of sensors and sensor arrays by orders of magnitude, and increase their spatial and temporal resolution and accuracy. Large numbers of sensors may be integrated into local- or wide-area systems to improve performance and lifetime, and decrease life-cycle costs. Communications networks provide rapid access to information and computing, eliminating the barriers of distance and time for tracking endangered species, detecting insects and pathogens, monitoring engineered structures and air and water quality. The coming years will likely see a growing reliance on and need for more powerful sensor systems, with increased performance and ecological functionality. search and management of environmental and ecological systems. Engineering and Technology Development The second phase of technology R&D involves applied engineering, wherein scientific discoveries are turned into new prototypes. These early stage technologies undergo testing and validation (either in the laboratory or in the field) to establish their capabilities and limitations. It is at this point where theoretical expectations and operational realities often come into conflict, and solutions and compromises must be tried, tested, and resolved. Companion IT also needs to be developed to make the new technologies operationally effective. The following paragraphs highlight several emerging technologies—biosensing, micro-electromechanical systems (MEMS), and sensor networks—that offer new and innovative possibilities for precision forestry, and mitigate some of the aforementioned difficulties working in forest environments. In agriculture, food, and the environment, there is an ever-increasing need to detect and measure minute quantities of chemicals or microbes (e.g., biosecurity) occurring in both indoor and outdoor environments, and to do so almost instantaneously (just-in-time information). Areas of particular interest include: food production and processing, agricultural products, pest management, surface and ground water, soils, and air. Universities, federal laboratories, and other federal agencies have been developing biosensing technologies to measure trace levels of biological and chemical materials in real-time. Biosensing includes systems that incorporate a variety of means, including electrical, mechanical, and photonic devices; biological materials (e.g., tissue, enzymes, nucleic acids, etc.); and chemical analysis to produce detectable signals for the monitoring or identification of biological phenomena. In a broader sense, the biosensing includes any approach to detecting biological elements (or their chemical signatures) and the associated software or computer identification technologies (e.g., imaging) that identify biological characteristics. Because of the scale of these biological entities and the masses involved, new advances in nanoscience and nanotechnology are proving useful. MEMS integrate mechanical elements, sensors, actuators, and electronics on a common silicon platform. MEMS make possible the realization of complete systems-on-a-chip. Sensors gather information from the environment by measuring mechanical, thermal, biological, chemical, optical, or magnetic phenomena. The electronics then process the information derived from the sensors, and through some decision making capability direct the actuators to respond by moving, positioning, regulating, pumping, or filtering, thereby controlling the environment for some desired outcome or purpose. For many environmental applications, the actuation step will take the form of a wireless transmission of data collected. These devices become particularly useful and powerful, however, when combined into networks Application, Adoption, and Economics Enabling technologies are converging with fields of application, e.g., agriculture and forestry, to provide the measurement, storage, analysis, and decision-making needs of producers and processors. In many cases, though, innovations are frequently adopted in clusters; e.g., genetically improved rice + fertilizer + insecticide. Here, there would be little economic payback for applying costly agronomic treatments to low-yield rice, whereas the same treatments applied to an improved rice strain would be more readily adopted. The marriage of remote sensing and geographic information systems in forestry represents another cluster example. In agriculture, for example, techniques are currently being developed to: (1) make precise measurements and continuously monitor field and plant conditions through sensors and instruments, (2) organize large volumes of data with spatially referenced databases, and (3) analyze and interpret that information using decision support systems that make economically favorable choices. The greatest “technology push” has been in precision agriculture (PA)—where information technologies provide, process, and analyze multisource data of high spatial and temporal resolution for crop production operations. Very similar technologies are being developed and promoted in the forestry arena for timber production and ecological assessments. Despite this “push,” the “pull” by the end-user community has been hesitant and weak, although most producers admit that they will have to adopt PA technology eventually. Currently, most see initial cost, uncertain economic returns, and technology complexity as limiting factors. These empirical observations are consistent with Rogers’s theory of innovation diffusion (Rogers 1995). Furthermore, in light of recent and anticipated regulatory requirements for nutrient release and water/air quality, many producers feel that the environmental benefits of precision agriculture might 11 be the eventual driving force for technology adoption. Nevertheless, small- and medium-sized producers (both in agriculture and forestry) have a distinct disadvantage versus large producers. In high-volume food and fiber production, economies of scale and narrow profit margins provide an economic advantage to large producers. Furthermore, large producers tend to have more education and are less technology averse than smaller producers. These characteristics of food and fiber production suggest that most technological advances, including precision agriculture/forestry, are not scale neutral. Furthermore, the factors limiting PA adoption, noted above, are also less problematic for larger producers, giving them an additional competitive advantage. One way for smaller producers to combat these competition trends is to create, or reach into, unique markets where their small size is an advantage. Value-added products expand the profit margin for producers that are positioned to provide enhanced value to consumers—which is more often the case for small producers that deal with small quantities of raw products and have more direct access to consumers. In addition, smaller producers can become more competitive in a technology world by mitigating the barriers to adoption. By spreading the initial cost of technology over many producers and by sharing information about how to use the technology, smaller producers can obtain the adoption capabilities held by large producers. One way to accomplish these tasks—that has been applied successfully by nonindustrial private forest (NIPF) landowners—is by forming landowner cooperatives (Stevens et al. 1999). These cooperatives are grass-roots activities (as distinct from existing agricultural cooperative enterprises) wherein members share equipment, information, and market power to achieve some common goals for managing their operations. A nominal fee is usually charged members and the cooperative becomes a business entity. In the eastern U.S., approximately 60% of timberland resides in NIPF ownerships. Yet, only a small portion of that acreage is actively managed. In the past several decades, the number of forestland owners has been increasing, with more non-farm and absentee owners. This new cohort of owners also has diverse interests. As with agriculture, precision forestry technologies are more readily adopted for use on large ownerships (industrial and public), but if economic and educational hurdles can be overcome, smaller ownerships will also participate, either individually or in groups. manipulation, data availability, and “alternatives” formulation and selection. The ethical concerns addressed below do not include intentionally malicious behavior, such as computer crime, software theft, hacking, viruses, surveillance, and deliberate invasions of privacy, but rather explores the subtler, yet important, impacts that data collection and use can have on people and their social, cultural, corporate, and other institutions. Privacy Improper access to personal information is the issue that “privacy” usually brings to mind. Any unauthorized access to information about an individual or their property can be an invasion of privacy, just as unauthorized access to one’s property has traditionally been considered invasive. However, even authorized access may lead to privacy concerns, when access to separate data sources is used to combine information (Mason, 1986). For example, one institution may record landowners’ names and land ownerships, while another may be authorized to store land records and timber values for tax purposes. Individually, the databases are properly authorized, but if the records are combined by a third party, it may be possible for unauthorized parties to gain financial data about individual landowners. As environmental databases increase in size, complexity, and connectivity, projects that involve adding data fields or combining data or knowledge sources must consider the ethical implications of those activities. In recent years, a new privacy issue has arisen in the area of geographic information systems (GISs), related to location protection. For example, many cultural sites on public lands are protected either by law, policy, or regulation. Yet, entering site locations in a GIS may disclose locations for unethical use. One way around this problem is to define a polygon that contains a site or group of sites, without disclosing exact point locations. A similar situation exists in relation to biodiversity and rare species protection. Innovative approaches are required to facilitate resource monitoring and protection while simultaneously ensuring there is no loss of privacy resulting from location disclosure. Current remote sensing technologies allow anyone to “look into” someone else’s property—assessing, without the owners knowledge or consent, timber or crop value that can be used for insurance, bank loan, or taxation purposes. Even when such data have been collected legitimately, there is no guarantee that adequate safeguards have been instituted to protect unauthorized access and use. Future technologies will create even greater opportunities for remote intrusion. As more and more data become available on the Internet, unintended use has become a major problem. Web surfers can borrow data from different source or combine data inappropriately from many sources either misusing it or misattributing it. Similar concerns related to re-packaging of information may arise where public funding of government research places researcher and research information in the public domain. Enterprising organizations, then, turn ETHICS AND FUTURE DATA AVAILABILITY Ethics is the study of value concepts such as “good,” “bad,” “right,” “wrong,” and “ought,” applied to actions in relation to group norms and rules. Therefore, it deals with many issues fundamental to practical decision-making. Precision and information technologies lie at the heart of modern decision making, including data/information storage and 12 that public-domain information into company revenue. While not illegal, it may be unethical if there is minimal value added to the publicly available, and public-funded, information. expense of global sustainability (Hart, 1999). Scale is a key determinant of indicator usefulness: some indicators that are useful at the household or community level are difficult to measure at the regional level, and some regional indicators may have little meaning at the community or household level. Because indicators compress so much ecological, economic, or social information into a single variable or set of variables, it is especially crucial that they are chosen, measured, and interpreted carefully. Accuracy may also be influenced by the sequence in which operations are applied. In theory, error limits of predictions should be supplied; however, while error limits of individual equations may be known, it is rare that models actually compute the consequences of combining multiple equations. Mowrer (2000) examines error propagation in simulation models and presents several approaches (Monte Carlo simulation and Taylor series expansion) to project errors. This has become an active research topic recently (q.v., Mowrer 2000), as several models are typically used in combination to predict future conditions. Key language and terminology used to frame a question can significantly influence the applicability of data or information. This is true for any information system in which the user is forced to converse using concepts unfamiliar to them. This cultural mismatch is of special significance in studies of Native peoples, where the interview subject may have concepts and values very different from those of the questioner. For example, the term “forest” is a key concept for resource management, but certain Native peoples have no concept for forest in their culture or any word in their native tongue. Instead, they have a more holistic view of the land that includes trees, plants, animals, and people (Thomson, 2000). Once such basic cultural differences are identified, the important challenge becomes one of understanding the ramifications of those differences, how they affect data needs and data use. Statistics, images, graphs, and maps are all methods of summarizing, presenting, or filtering information. Ethical decisions behind the selection and transformation of material can significantly affect the accuracy with which recipients may perceive a situation. When a situation is highly charged or contentious, objectivity in portraying information becomes critically important. Certain decision support software may increase considerably the power of users to make or influence decisions that were formerly beyond the limits of their knowledge and experience. For example, upper management may gain direct access to lower level data and information summaries. This helps bypass intervening distortions, resulting in more accurate perceptions. Greater accuracy is dependent, however, on support software that has itself been developed with appropriate ethical considerations and higher level managers must be willing and able to use the software to achieve distortion-free information sharing. This type of situation has been a bane of statisticians for years. Very powerful software packages have allowed users to perform all manner of inappropriate statistical tests on data without full Accuracy A software developer’s ability to know and predict all states (especially error states) is low for complex systems. At first sight, it would appear that a software developer would be ethically bound to correct all system errors. However, dealing with errors can raise ethical dilemmas: 15-20% of attempts to remove program errors introduce one or more new errors. For programs with more than 100,000 lines of code, the chance of introducing a severe error when correcting an original error is so large that it may be better to retain and work around the original error rather than try to correct it (Forester and Morrison, 1994). The frequency of disclaimers, software updates and patches, as well as the lack of substance to software warranties, result from software developers’ recognition of this problem. The ultimate effect is larger and more complex software, whose size is less related to functional capability than it is related to software age and the battery of “fixes” that it has received over time. Similar ethical conflicts arise with decision support tools, where modelers and developers realize that a model’s results can only be broad approximations in many cases. Another problem related to accuracy is determining which specific information to use. For example, it is often difficult to select appropriate socio-economic or biological indicators or to choose among predictive models. An indicator is something that points to an outcome or condition, and shows how well a system is working in relation to that outcome or condition. For example, in a forest simulation model, tree diameter at breast height (dbh) is a key indicator of treatment effects. However, there may be a range of potential equations available to predict dbh. One equation may simply predict dbh from tree height, while another equation may predict it from both height and crown width. The equation selected will have different consequences with regard to accuracy, precision, data costs, and suitability for extrapolation. This choice relates, in turn, to precision and bias in the estimators used. Requirements of the intended user and usage should guide the choice. When a social or economic indicator is being used, ethical considerations are even more significant. If the indicator misrepresents a value set, then it cannot be considered accurate. Indicators have long been used in predictive systems (Holling, 1978): such indicators must be relevant, understandable, reliable, and timely. In natural resource disciplines, with their current emphasis on sustainability, indicators must have additional characteristics. Sustainability indicators must include community carrying capacity; they must highlight the links between economic, social and environmental well-being; they must be usable by the people in the community; they must focus on a long range view; and they must measure local sustainability that is not at the 13 knowledge of what they are doing. While current statistical software manuals contain a great deal of information regarding model specification and assumptions, they cannot replace a well-founded understanding of basic statistics by the experimenter. lie in the adoption phase. Even though engineering and technology development aspects may seem daunting and time-consuming, it is the economic, cultural, and educational issues that often doom or advance technology use. This suggests that more thought needs to be given to that final phase. Some issues that need to be addressed in the development phase (or pre-diffusion) are: intended users, intended uses, workflow changes, education and training, economics, associated IT changes or requirements, favorable or unfavorable regulations, early adopters, commercialization entities, and user communities. All these factors can impact if, and how, a new technology is accepted and used. Privacy has long been considered an inherent right of individuals in a “free” society. Initially, this involved protection of the individual from unwanted or unwarranted invasion of their physical space. More recently, privacy has been extended into an individual’s information space, as well. For precision technologies currently under development in natural resource and agricultural domains, more real threats are likely to arise from unintentional and unforeseen information breaches than from any intentional conspiracy. These occur when information sources are combined or used in unintended ways. As long as information about individuals exists and is accessible by others, individual privacy can potentially be compromised. During technology development, designers need to be cognizant of users, co-developers, publics, cultures, special interest groups, commercial enterprises, governments, and other groups that might be affected directly or indirectly by their products. Designers must also consider the information their technology uses or generates, and the decision-making landscape that it affects or creates. Use of appropriate language is at the heart of many accuracy issues. Even if an information system does not estimate the accuracy of results explicitly, it is important to make end-users aware of the variability in potential outcomes, and the assumptions and trade-offs that have contributed to it. Similarly, non-textual rendering of system outputs should be designed to address accuracy concerns in the flow of knowledge. It is also essential to address the way in which knowledge flows through organizational hierarchies, and to ensure its appropriate use at different organizational levels. As with accuracy issues, language lies at the heart of many accessibility issues. Information delivery must be geared to concepts appropriate to the intended audience, and information overload avoided, as knowledge can be inaccessible if the recipient is swamped with information. Limitations of technical accessibility by some groups may require developing an integrated range of systems and processes to ensure access by all stakeholders in a decision environment. While there will always be some ethical culpability on the individual’s part, much responsibility still rests with organizations to institute standards of ethical conduct that create an atmosphere of social morality for their employees and members. Self regulation is always more readily accepted and effective than regulation from governmental institutions, Accessibility Appropriate access to data and software has both technical and intellectual components. To make use of software, a person must have access to the required hardware and software technology, must be able to provide any required input, and must be able to comprehend the information presented. For example, for a Web-based system, users must have reliable connections to the Internet and sufficient bandwidth. Each end-users must also have a browser compatible with the material sent to it (including such things as the appropriate Java classes for use with applets) and any helper applications or browser plug-ins for viewing and hearing content. If an intended audience lives in a developing country, or in a remote area, such technological issues may be critical. For this reason, when software or a data base is developed, its implementation should be part of an integrated process that includes the full range of affected individuals. This may include specifying duties for a suite of “actors” such as technology transfer officers or field personnel. Accessibility is also limited if results are presented inappropriately. For example, data may be aggregated at a fixed scale that may have limited value for many users. In other cases, language and concepts beyond the end-user’s understanding or vernacular might render a decision support system useless for a large audience segment. While it is neither practical nor possible to accommodate all who might “stumble onto” data or software, primary target audiences need to be defined and understood. In a digitally networked age, the ability to connect systems, databases and information-rich environments becomes more possible but also more problematic. The goal of seamless, transparent, and “user-friendly” information access makes interoperability a required attribute of databases, systems, and vocabularies. This desired attribute requires both technical and human dimensions to enhance interoperability within regional, national, and global forest information systems. Interoperability ensures that systems, procedures, and cultures of an organization are managed in such a way as to maximize opportunities for exchange and re-use of information, whether internally or externally. Because end-users of data are not necessarily local or regional and because large-scale forest assessments are becoming more important (e.g., carbon management), standards and protocols for forest data are looming on the horizon. SOME STEPS TO TAKE Once basic, scientific principles have been demonstrated, the biggest hurdles to realizing an operational technology 14 which may not always fully understand the issues involved. By thinking in advance about ethical issues that may eventually impinge on data availability, organization might alleviate potential future restrictions or reduce their impacts. sensitive biosensor for Salmonella. Biosensors and Bioelectroncis 15: 135-141. Rogers, E. M. 1995. Diffusion of Innovations, Fourth Edition. The Free Press, New York. Schmoldt, D. L. 2001. Precision agriculture and information technology. Computers and Electronics in Agriculture 30(1/3): 5-7. LITERATURE CITED Clark, N. 2001. Applications of an automated stem measurer for precision forestry. Pages 93-98 in D. Briggs (ed.) Proceedings of the First International Precision Forestry Cooperative Symposium, College of Forest Resource, Univ. of Washington, Seattle WA. Simula, M., J. Lounasvuori, J. Löytömäki, M. Rytkönen. 2002. Implications of forest certification for information management systems of forestry organizations. Forest information technology 2002 international congress and exhibition. 6 pp. www.indufor.fi/documents%26reports/pdf-files/ article07.pdf Forester, T., and P. Morrison. 1994. Computer Ethics. MIT Press, Cambridge, Mass. Hart, M. 1999. Guide to sustainable community indicators. Hart Environmental Data, North Andover, MA. Stevens, T. H., D. Dennis, D. Kittredge and M. Richenbach. 1999. Attitudes and preferences toward cooperative agreements for management of private forestlands in the Northeastern United States. Journal of Environmental Management 55:81-90. Holling, C.S. 1978. Adaptive environmental assessment and monitoring. John Wiley & Sons, Chichester. Mason, R.O. 1986. Four ethical issues of the information age. MIS Quarterly 10(1): 5- 12. Thomson, A.J. 2000. Elicitation and representation of Traditional Ecological Knowledge, for use in forest management. Computers and Electronics in Agriculture 27(1-3): 155-165. Mowrer, T. 2000. Uncertainty in natural resource decision support systems: Sources, interpretation, and importance. Computers and Electronics in Agriculture 27(1-3): 139-154. Thomson, A. J., and D. L. Schmoldt. 2001. Ethics in computer software design and development. Computers and Electronics in Agriculture 30(1/3): 85102. Pathirana, S. T., J. Barbaree, B. A. Chin, M. G. Hartell, W. C. Neely, and V. Vodyanoy. 2000. Rapid and 15 16 Multidata and Opti-Grade: Two Innovative Solutions to Better Manage Forestry Operations PIERRE TURCOTTE Abstract: You can’t manage what you don’t measure. Two novel systems recently developed by FERIC address this dilemma: MultiDAT allows forest contractors to maximize their machine uptime and Opti-Grade provides an integrated package for optimal forest road management. MultiDAT is a multi-purpose datalogger for forestry managers. MultiDAT can record machine functions, machine movement, machine location and collect operator feedback. The associated software can analyze the data and produce reports on which optimal decisions can be based. The MultiDAT is designed specifically for heavy equipment operating in areas where communication systems are not existent or very expensive. Opti-Grade is a road management system to help focus grading or re-profiling activities where they will have the greatest impact on the road condition for the money invested. Opti-Grade is used to collect important information on the condition of the road network on a regular basis, using equipment installed on a log truck. This data is used to schedule maintenance activities. This new concept is considerably more efficient than the traditional concept of grading whole road segments. apply for forest machines, because almost all reporting on truck systems are based on distance and not on time. We had no choice but to develop a complete system. CONTEXT OF THESE RESEARCH PROJECTS The two projects described here were done at the Forest Engineering Research Institute of Canada (FERIC) during the last 3 years. FERIC is a private research organization that has served the Canadian forest industry for more than 25 years, and started recruiting new members in the United States this year. Projects are determined for the most part by the industrial members and orientations are reviewed on a yearly basis, which accounts for very practical, results-oriented research. FERIC programs cover all aspects of forest operations, from silviculture to harvesting and transportation, but because of the general theme of this symposium, two projects closely related to precision forestry will be presented. Why measure machine utilization? Improving machine utilization has a major impact on the profits of the contractors who conduct a large portion of forest operations. The documentation of the daily operating hours and the nature of all delays is the first step in determining the possible avenues of improvement. To illustrate the importance of machine utilization in today’s context, let’s analyze the effect of improving the utilization of a typical harvester operating in Eastern Canada. This example is based on the following assumptions (all $ in Canadian funds): Cost of the harvester: Useful life: Resale value: Insurance: Repairs and maintenance: Interest rate: Operation schedule: Cost of operator: Cost of fuel: Productivity: Origin of the MultiDAT development Foresters have been using paper chart recorders for more than 30 years to track the utilization of their equipment. They are still in use today in many operations. This crude system has many limitations. The precision of the recording is usually only about 5 to 10 minutes, the charts get easily damaged and the operators can falsify the recordings. When members asked FERIC to develop or find a simple electronic device, we first tried to adapt a recorder designed for trucks. We realized that the truck paradigm does not Revenue: 17 $480,000 5 years $144,000 $24,000 per year $96,000 per year 8% 4,000 hours per year $30 per hour $18.40 per hour 40 cubic metres per productive machine hour $3.00 per cubic metre EXAMPLES OF USE These assumptions do not necessarily represent an average, but are only typical values used to represent the relative importance of the costs. Figure 1 shows the variation of net profit that the owner of this harvester would make if the utilization rate would vary from 70% to 85%. It is easy to see that the owner would make 7 times more profit if the machine is used at 85% instead of 70%. Weyerhaeuser Company Limited in Dryden, Ontario have used 20 MultiDAT Junior to support a productivity improvement program in partnership with two major contractors. For more than a year, they followed a fleet of harvesting equipment, tracking downtime and finding solutions to reduce its impact and occurrence. In some cases, the utilization of skidders improved from 50% to more than 80%. They used very simple configurations and connection methods, relying mostly on the motion sensor to determine machine activity. In Quebec, Gestion Remabec inc., a large contractor, is using the MultiDAT on harvesters for two purposes. First, they monitor the utilization using the sensors, but they also use the GPS option and record the travel path of the machine. When harvesting is completed in a block, they provide their clients with a map showing the area harvested. This map is used to make sure that the operator was not in violation when working close to the block boundaries. The Saskatchewan Department of Transportation is using the MultiDAT exclusively on graders. The MultiDAT PC software has a speed analysis function that can be used to determine approximately the sections of road that were graded. In general, grader operators drive at a higher speed when they are traveling than when they are grading. In the Atlantic Provinces, J.D.Irving Ltd. is gradually implementing the MultiDAT on all their contractors’ equipment. The company provided the MultiDATs free of charge to the contractors in exchange for their commitment to use them and provide utilization reports. In Windsor, Quebec, Domtar Inc. used a GPS- equipped MultiDAT in 2002 to track site- preparation equipment. Although at the time, the MultiDATs were using Garmin25 GPS receivers with no WAAS capabilities, the results were very good; most of the time within 1% of the area measured by walking the block with a GPS receiver after the work was completed. In 2003, they are using GPS-equipped MultiDATs on all of their site-preparation equipment. These MultiDATs now use the CSI SX-1 WAAS receiver, with submetre accuracy. In 2002, the MultiDAT was also used by FERIC researchers to evaluate the productivity of new models of Tigercat equipment in a Tembec Industries Inc. harvesting operation in Ontario. The operations were followed in detail for more than 3 months, and the MultiDAT recordings were downloaded each week by the supervisors and transferred to FERIC for analysis. MultiDAT has been used successfully on harvesters, feller bunchers, skidders, forwarders, bulldozers, excavators, graders, sand trucks, mobile chippers, and loaders. Figure 1: Profit VS Utilization DESCRIPTION OF THE MULTIDAT The MultiDAT has the following characteristics: - 4 channels programmable as timers, counters, or frequency meters - Internal motion sensor sensitive to low frequency vibrations - Optional WAAS GPS receiver - Typical autonomy: 2 to 6 weeks The MultiDAT comes in two versions. The regular unit has an operator interface that allows easy input of operator number, work codes and delay codes. The MultiDAT Junior has all the characteristics of the regular unit without the operator interface. The recordings are downloaded and transferred to a PC using a PDA, either a Palm OS or Windows CE device. The average size of a MultiDAT download file is 50 Kb for one week of recording, which means that a data shuttle can easily download many MultiDATs before being synchronized with a PC. The MultiDAT is built for the harsh conditions in which heavy equipment operates. It is enclosed in a heavy aluminum casing, all components can withstand temperatures varying from -40 to 85 deg. C, and the supply voltage can be from 10 to 28 volts. Almost everything on the MultiDAT is configurable. The number of sensor channels used, the threshold of the motion sensor, the number of work and delay codes used and their meaning. Even the report format is fully configurable. STRONG AND WEAK POINTS The strongest point of the MultiDAT is its versatility. The recorder itself and the PC software can be configured 18 CURRENT DEVELOPMENT PROJECTS in a very simple way to simply record the working hours of a bulldozer for example. They can also be configured in a more complex way to identify the harvesting time, idle time and traveling time of a harvester in a given block. A second strong point is the ease of use for the operator. As shown in Figure 2, the operator interface is very simple, there is no scroll down menus and all the activity and stop codes are visible at a glance. Current projects include the development of geo-fencing for the MultiDAT. Some analysis software for truck dataloggers give the user the possibility of defining polygon boundaries called geo-fences. The GPS position recordings can then be analyzed in relation to these geo-fences, and additional information can be derived such as the time spent in a given region, or the number of times that a truck passed a given location. With the MultiDAT, the geo-fences are pre-determined, and the recorder can be configured to record only the time of entry into and exit out of each polygon. This method requires much less memory than recording all the GPS positions and does not require intensive computer processing after the data is downloaded. The first application that we envision for this development is the management of wood flow in the mill yard. By tracking the passage of loaders between zones, we will attempt to determine the volume of wood that is moved between various sections of the yard and better balance the tasks assigned to each loader. Finally, we are working on the development of a blade contact sensor for graders. This sensor will provide a more accurate map of the road sections that were graded. Since February 2003, the MultiDAT is fabricated and distributed under license by Geneq (www.geneq.com). Figure 2: Operator interface The first weak point of the MultiDAT is, paradoxically, its versatility, which sometimes makes the initial configuration laborious. Many of the first MultiDAT users needed support to configure the recorders and the reports, simply because they did not know what was the most effective configuration. With later versions of the software, we provided more configuration templates and we are using the experience of the first users to determine the configurations that give the best results. We are thus transforming the MultiDAT from a simple activity recorder into a methodology to improve machine utilization. The second weak point is the data shuttle. Although since the inception of the MultiDAT, users could choose between a $150 consumer PDA and a $1500 rugged field computer, so far only consumer PDAs have been used as shuttles. Batteries have given users the most problems. The average user is not aware that a Palm computer is never turned off, even when the display is switched off, and that it wears out its batteries over a period of 3 to 6 weeks even when sitting in a drawer. That is, until users lose their first data files and have to re-install the shuttle program in the PDA. In 2003, we are introducing shuttles using Windows CE, and that should improve the shuttle performance drastically. The data and shuttle programs can now be backed up in flash memory, and the PC communication is much simpler, using the USB cable that is provided with the PDA. Still, keeping ahead of product development for the shuttle is not an easy task. We saw product lives of 9 months with the Palm PDAs and we hope that the Windows CE devices will stay on the market longer. THE OPTI-GRADE SYSTEM During the last two years, FERIC also worked on the development of another precision forestry tool, the OptiGrade system. Opti-Grade follows two simple principles: 1- Measuring road roughness lets you identify sections that need grading—and those that don’t. 2- Using graders efficiently means grading only the sections that need it, and travelling at full speed over sections that don’t. HOW DOES IT WORK? Before the Opti-Grade system is used, the road network (all km markers, bridges, intersections, etc.) must be surveyed using GPS. Then, the recording equipment is installed in one of the trucks that regularly travel the road being monitored. While a sensor continuously measures the roughness of the road, a GPS receiver determines the time and position of each measurement. A datalogger stores the road roughness values, plus the position and recording time of each value. 19 When the truck enters the mill yard, the recordings are transferred by spread spectrum radio to an office computer. The Opti-Grade software then analyzes the roughness and GPS data to determine which road sections require grading and calculates the travel speed of the truck on each section. The user can display a map showing the road roughness for different sections as seen in Figure 3. Because the kilometre markers of the road have been surveyed when the Opti-Grade system was set-up, the map can show the correspondence between those markers and the sections to be graded. typical grading schedule. On this example, the operator only needs to grade 22 kilometres out of the total 63 kilometres of the road, thus saving 65% of the normal grading time. TYPICAL RESULTS The Opti-Grade system is in use at more than 20 locations in Canada. The users generally attempt to maintain the same road quality while reducing grading cost. Experience with the system has shown that the grading costs can be reduced by up to 30% on average. For a company maintaining 400 km of roads, this can represent an annual saving of more than $70,000, or a payback of a few months. The companies using Opti-Grade are also very interested with the speed and time recordings provided by the system. They use this information to establish more precise cycle times and thus fairer trucking rates. Finally, analysis of Opti-Grade data helps to identify road segments that need constant maintenance. Investments in road improvements can thus be justified more easily. The software then prepares an optimized grading schedule based on three criteria: 123- a roughness level above a trigger threshold the minimum length of road to treat the minimum distance between sections to be treated The next morning, the grader operator uses that schedule to determine the sections to grade. Figure 4 shows a Figure 3: Map of road roughness Figure 4: Grading schedule 20 A Test of the Applanix POS LS Inertial Positioning System for the Collection of Terrestrial Coordinates Under a Heavy Forest Canopy STEPHEN E. REUTEBUCH, WARD W. CARSON, AND KAMAL M. AHMED Abstract: The Applanix POS LS backpack-mounted inertial land positioning/navigation system was used to collect terrestrial coordinates along a previously surveyed closed traverse. A total station surveying instrument was used to establish 26 ground-level stakes along a 1 mile traverse under the dense canopy of a 70 year-old conifer forest in the Capitol State Forest near Olympia, Washington. The Applanix POS LS was initialized at a fixed monument and carried through the forest along the traverse 12 times. Coordinate readings were collected continuously both at the survey posts and between posts. Both the system’s location accuracy and its potential for developing terrain profiles were evaluated. The system’s average real-time position accuracy was 2.3 ft (1.6 ft Stdev., 7.0 ft max.) and average post-processed accuracy was 1.4 ft (0.9 ft Stdev., 4.0 ft max.), measured at each survey stake. An earlier study provided a 5 by 5-foot, gridded digital terrain model (DEM) derived from high-density LIDAR data. Profiles generated from the LIDAR DEM were compared with profiles measured by the POS LS system. Average post-processed elevation difference along the profiles was 0.7 ft (1.0 ft Stdev., 4.5 ft max.). INTRODUCTION In general, the INS and GPS components of an integrated POS system complement each other’s strengths or, rather, compensate for weaknesses (Farrel and Barth, 1999). With an initial coordinate fix and the acceleration vectors sensed by the IMU, an INS can integrate the velocity vectors and compute the coordinate path as the unit moves; however, positional accuracy is eroded as instrument drift accumulates over time. In contrast, GPS errors are not accumulated over time but, rather, GPS accuracy is maintained by regular, frequent, independent readings. In short, an INS can measure direction and distance in the short run, but benefits greatly by regular coordinate updates from the GPS to correct drift. The GPS is good over the long run, and benefits greatly from the INS data acquired between GPS recordings. As a surveying device operating in the open under a good constellation of GPS satellites, the POS LS will deliver coordinates accurate in the range of RTK GPS capabilities—approximately 4 inches or better. However, when the GPS signal is blocked, under a forest canopy for example, it reverts to sole dependence upon its INS, and, the error due to drift will begin to diminish the position accuracy over time. The effect of INS drift can be mitigated in two ways: 1) by position updates—the obvious technique of re-initializing the system position with either GPS readings or by periodically Applanix* [a Canadian company that has developed a number of position and orientation systems (POS) based upon inertial navigation systems (INS)] has recently produced a system designed for land surveyors (Gillet et al., 2001). The POS LS system combines an INS [with its embedded inertial measurement unit (IMU)], a roving global position system (GPS) unit, and a computer datalogger into a backpack system weighing about 40 pounds. When utilizing the internal roving GPS receiver, the POS LS unit is intended to be used with a user-supplied realtime kinematic (RTK) GPS basestation that would provide the necessary carrier-phase ambiguity resolution, thereby supplying frequent, accurate coordinate updates to the INS system. Applanix has established with other products, such as the airborne POS AV system, that uninterrupted, postprocessed data from such a GPS/INS system can deliver coordi-nates accurate in the range of inches. However, Applanix anticipates that the land sur-veyor will on occasion lose the GPS signal—for example, under a forest canopy. The question then becomes: what accuracy can one expect from the POS LS system under these less than optimal GPS conditions? * Use of trade or firm names in this publication is for reader information and does not imply endorsement by the U.S. Department of Agriculture of any product or service. 21 re-visiting known points; or, 2) by zero velocity updates (ZUPTs)—a method used to obtain a velocity re-initialization that has been incorporated into the POS LS instrument. If the INS unit is momentarily held still (at zero velocity), Elosegui et al., 1995; Firth and Brownlie, 1998; Lachapelle and Henriksen, 1995). The POS LS unit offers an alternative method for collecting geographic positions under such adverse conditions. The accuracy of the POS LS system is a function of the frequency of coordinate updates—from either a GPS signal or the input of known coordinates—and the frequency of ZUPTs. In this initial study, we only examined POS LS coordinate accuracy at a fixed ZUPT interval (nominally 30 seconds) under forest canopy. We examined both real-time accuracy of the POS LS unit and accuracy obtained by postprocessing POS LS positions after each trial run (traverse) was closed on a known point. It is also important to note that in our test, because of extremely dense canopy conditions, GPS positions were not collected with the POS LS unit. Instead, the unit was initialized over previously surveyed reference points for each trial run. These reference points were located in a clearcut adjacent to the forested area. In practice, the GPS unit in the POS LS could have been used in the clearcut to accurately establish the initial location of the instrument before entering and after emerging from dense forest. METHODS On May 21-22, 2002, Applanix technical personnel brought a POS LS instrument to the Capitol State Forest near Olympia, Washington for trials under the canopy in our forest test site. The forest is managed by the Washington State Department of Natural Resources. Our test site has a mix of forest canopy cover, ranging from 70-year-old conifer cover to recently clearcut areas. It has been the site of several other geomatic (Reutebuch et al., 2003) and forestry (Curtis, et al., in press) research trials. LIDAR data sets were collected in 1998, 1999, and 2000, and a high resolution, 5x5-ft gridded digital terrain model (DTM) was produced from the 1999 data. Additionally, a closed traverse, total station survey was performed under a full-canopy segment of the forest and the staked-points were available for use in assessing the accuracy of the POS LS system. Our test of the POS LS unit was build primarily around re-visiting these surveyed points. A comparison was also conducted between POS LS position elevations and elevations interpolated from the LIDAR-based DTM. The closed traverse survey loop consisted of 26 points, marked with 2x2-inch wooden pegs driven into the forest floor down to ground level. Two reference points, marked as 1A and 2A, were established from local HARN points with a carrier-phase, survey-grade GPS instrument. Other points, spaced around a roughly circular traverse of approximately 1 mile in length, were established with a Topcon ITS-1 total station survey instrument. Closure calculations showed the horizontal accuracy was 1:2840 and vertical closure was 1.1 inches. After adjustment, the horizontal and vertical accuracy of the ground points were within 6 inches and 1 inch, respectively. Figure 1: Applanix POS LS system is held steady in one position during a ZUPT. the POS LS software can re-initialize the velocity vector to zero and thereby correct for accumulated velocity drift. In Figure 1, note how the operator uses a blue staff to help hold the system steady during a ZUPT. The desired time interval between ZUPTs is user defined; however, longer intervals increase position errors. When under dense canopy (when the position is being updated using only INS data) an audible announcement and text display on the POS LS datalogger informs the user when it is time for either a ZUPT or position fix (acquiring a new GPS location in a clearing or moving to a known point). A ZUPT is also automatically initiated when the INS unit senses that it is stationary. Additionally, the operator can manually initiate a ZUPT at any time. At the end of a survey, the location of the POS LS unit is accurately established by either acquiring a high-accuracy GPS position in an opening, or by returning to a known reference point. This allows the traverse data to be post-processed to derive more accurate, adjusted positions. OBJECTIVES It is well established that GPS is not reliable for surveys under or near a forest canopy due to obstruction of GPS satellite signals or signal multi-path problems (Darche, 1998; 22 Sets of POS LS coordinate data were collected continuously at a once per second rate over the course of the closed traverse. Most of the readings were collected while in transit between survey stakes; however, specific blocks of recordings were noted. These blocks were: 1) Alignment Fix: The operator set the backpack at reference point 1A to establish the initial position and allow the system to determine true north. 2) Point Visitation: After alignment, with the instrument on his back, the operator located himself over a survey point (i.e., the 2x2-inch peg) and held himself steady enough to record several seconds of consistent coordinate readings. (Note: A vertical bias of 3.0 ft was subtracted during data reduction to account for the height of the unit’s recording point above the peg in these standing positions). 3) ZUPTs: When alerted by the unit, the operator stopped with the instrument still on his back and held steady—at zero velocity—for several seconds. 4) Position Fix: The operator took the instrument off his back and set it on a survey stake for several seconds and commanded the system to update position. (A survey stake approximately midway through our closed traverse, was used as this intermediate point). that the operator would see on the datalogger coordinate readout in the field as the POS LS is being carried in the forest. Each real-time data file begins with an initial Position Fix. The data from that initial point forward were computed by dead reckoning based upon the IMU readings and INS projections augmented by the operational ZUPTs. The ‘post-processed’ data result from the same recordings, but they depend upon a final Position Fix at the end of each run. An algorithm implemented in the Applanix POSPac software is designed to adjust to zero the error at this fixed terminus and to minimize the error over each run. Both these data sets were examined in this study to quantify the real-time point-by-point accuracy that one can expect in the field, and the accuracy obtainable from further POSPac refinements accomplished after data collection in the office. RESULTS Both the ‘real-time’ and ‘post-processed’ data from this test are presented similarly. Tables 1 and 2 summarize the basic results, including average of coordinate errors at all survey stakes, standard deviation, and maximum error associated with each run. Plots display error accumulation over time for ‘real-time’ and ‘post-processed’ data and differences between them (fig. 2 and 3). Run descriptions A ‘run’ is defined by an initial Position Fix and, with the exception of Run 5, is termi-nated by a final Position Fix. (Run 5 was terminated by an unexpected battery failure and, therefore, did not have a terminal fix and could not be postprocessed). Each run took a certain time—recorded and shown in seconds—and covered a certain point-to-point distance—computed as the accumu-lated, straight-line distance between the points visited. The count of the actual num-ber of points visited is shown in the tables as well. The average run time, number of points, and length was 48 minutes, 2,440 points, and 2,472 ft, respectively. Coordinate and orientation data were collected continuously while following the closed traverse through several loops. We divided this continuous stream of data into runs. Each run was initiated at a Position Fix and terminated later at another Position Fix with Position Visitations and ZUPT updates registered in between. Twelve runs were made during our test. Data Management and Reduction During the two days of POS LS field testing, our operator (Joel Gillet from Applanix) tramped through our rough, forested terrain for a total of nearly 6 miles while stopping at 175 known positions to either re-initialize the instrument or record points as coordinate data. The task took over ten hours and the instrument, recording constantly at the rate of one coordinate set per second, collected nearly 40,000 points. From these data, Applanix delivered to us two types of coordinate files: 1) the ‘real-time’ files that held lists of field recorded ‘time, X, Y, Z’ data, and 2) the ‘post-processed’ files of the same data after adjustment. All coordinate data had been transformed into the State Plane System, Washington South Zone, NAD83, Mean Sea Level Elevation, NAVD88 datum, International Feet. The ‘real-time’ and ‘post-processed’ data are purposely distinguished in this report. The ‘real-time’ data are those Coordinate errors at survey stakes Tables 1 and 2 present the average coordinate errors (defined at each point as the coordinates collected over the survey stake (generally the average of ten readings) minus the survey stake coordinates) for each run for the ‘real-time’ and the ‘post-processed’ data, respectively. The overall error means and standard deviations, weighted in proportion to the number of points visited within each run, are computed and displayed at the bottom of each table. For the ‘real-time’ runs, the mean horizontal error at the stakes was only 2.3 ft (1.6 ft stdev, 7.0 ft max). The mean real-time elevation error was 1.4 ft (1.0 ft stdev, 5.4 ft max). For the ‘post-processed’ runs, the mean horizontal error at the stakes was only 1.4 ft (0.9 ft stdev, 4.0 ft max). The average post-processed elevation error was a remarkable 0.4 ft (0.3 ft stdev, 1.4 ft max). 23 Error Plots at survey stakes versus time differs by a tenth of a foot from the average of coordinates over a range of plus and minus 5 seconds, is taken as a point where the operator is in motion and between points. Table 3 summarizes the post-processed data in this ‘betweenpoints’ class. The table shows the total number of points recorded in a run, and the total number of points recorded while moving. There is remarkably little difference between the LIDAR DTM elevations and the POS LS elevations (DTM elevation minus the POS LS elevation). The mean, standard deviation, and root-mean-square difference over all the moving points were 0.7, 1.0, and 1.3 ft, respectively, with differences ranging from –4.1 to 4.5 ft. Figures 2 and 3 display the coordinate error as it accumulated over time during each run. The plots distinguish each run and are organized to contrast error drift in the ‘realtime’ data (fig.2) and the ‘post-processed’ data (fig.3). Comparisons with a LIDAR DTM As noted earlier, coordinates were being collected constantly at a one second interval while the operator traveled between points. We have distinguished these ‘betweenpoints’ blocks of data by noting the operator movements. By our definition, any point with a coordinate (X, Y) that Table 1. Real-time POS LS system error computed from ground survey stakes. Stakes Total Horizontal Error Vertical Error Run Run Length Visited Time (ft) (ft) No. (ft) (no.) (sec) Avg. Stdev. Max. Avg. Stdev. Max. 1 3732 20 6931 1.9 1.2 3.9 1.4 0.8 2.4 2 2241 15 2496 1.7 1.2 5.2 1.4 0.9 2.8 3 3799 20 5201 2.4 2.1 5.8 2.6 1.6 5.4 4 2141 12 2553 1.9 1.1 3.7 2.0 1.8 4.6 5 2597 12 3103 3.5 2.2 6.9 1.1 0.4 1.6 6 957 7 1290 4.0 2.5 7.0 0.8 0.4 1.4 7 2237 12 2018 2.7 1.5 4.6 1.2 0.7 2.3 8 253 3 806 1.4 0.7 1.9 0.7 0.4 1.1 9 3486 17 3104 2.7 1.8 5.3 1.6 0.8 3.1 10 2239 13 2245 2.0 1.1 3.6 1.2 0.6 1.8 11 3739 19 3192 1.9 1.5 4.8 0.8 0.7 2.0 12 2239 12 1819 1.6 0.6 2.4 1.3 0.8 2.3 Weighted Avg., Stdev., Max. Error--all runs 2.3 1.6 7.0 1.4 1.0 5.4 *Time spent collecting data at each stake and traveling to next stake. Combined Error (ft) Avg. Stdev. Max. 2.4 1.4 4.6 2.2 1.4 5.7 3.7 2.5 7.1 2.8 1.9 5.4 3.7 2.2 7.0 4.1 2.5 7.1 3.0 1.6 5.0 1.5 0.8 2.2 3.1 1.9 5.7 2.4 1.2 3.9 2.1 1.5 5.1 2.2 0.9 3.1 2.8 1.8 7.1 Time per Stake* (sec) Avg. Stdev. Max. 328 397 1912 162 127 529 256 198 893 207 137 522 259 208 797 177 62 256 165 101 374 254 226 515 180 98 402 167 100 351 165 95 415 148 87 312 Table 2. Post-processed POS LS system error computed from ground survey stakes. Stakes Total Horizontal Error Vertical Error Run Run Length Visited Time (ft) (ft) No. (ft) (no.) (sec) Avg. Stdev. Max. Avg. Stdev. Max. 1 3732 20 6931 1.1 0.6 2.6 0.3 0.3 0.8 2 2241 15 2496 2.3 1.1 4.0 0.5 0.3 1.0 3 3799 20 5201 1.5 0.9 3.1 0.5 0.4 1.4 4 2141 12 2553 1.3 0.8 2.4 0.6 0.5 1.3 6 957 7 1290 0.9 0.4 1.4 0.2 0.2 0.5 7 2237 12 2018 1.2 0.9 2.6 0.2 0.2 0.7 8 253 3 806 0.4 0.2 0.6 0.1 0.1 0.3 9 3486 17 3104 1.3 1.2 3.4 0.3 0.3 0.8 10 2239 13 2245 0.9 0.6 1.8 0.3 0.3 0.9 11 3739 19 3192 1.5 1.1 4.0 0.4 0.3 1.2 12 2239 12 1819 1.3 0.8 2.5 0.4 0.3 0.7 Weighted Avg., Stdev., Max. Error--all runs 1.4 0.9 4.0 0.4 0.3 1.4 *Time spent collecting data at each stake and traveling to next stake. 24 Combined Error (ft) Avg. Stdev. Max. 1.2 0.6 2.7 2.5 1.1 4.0 1.7 0.9 3.1 1.5 0.8 2.7 0.9 0.3 1.4 1.3 0.9 2.6 0.4 0.2 0.6 1.4 1.2 3.5 1.0 0.7 2.0 1.6 1.1 4.0 1.4 0.8 2.5 1.5 0.9 4.0 Time per Stake* (sec) Avg. Stdev. Max. 328 397 1912 162 127 529 256 198 893 207 137 522 177 62 256 165 101 374 254 226 515 180 98 402 167 100 351 165 95 415 148 87 312 Figure 2: Real-time combined (horizontal and vertical) position error over time. Figure 3: Post-processed combined (horizontal and vertical) position error over time. 25 DISCUSSION seconds (about 13 minutes) in Run 8, to 6931 seconds (almost 2 hours) in Run 1. The errors are generally dependent upon time, however, as is apparent in both Table 1 and Figure 2, there are exceptions. It does seem safe to expect a total vector error of less than 3 ft with a maximum error less than 8 ft for operations under 30 minutes in length. Table 2 and Figure 3 show the results when the same data are post-processed. Generally, as is apparent from the results, one can expect the error to be cut by half—total vector errors less than 1.5 ft and maximums under 4 ft for a 30 minute operation. Table 3 presents results in a format that should aid in our evaluation of the instrument’s potential for collecting data for a local DTM or linear profiles (stream, roads, trails, etc.). As mentioned above, the elevation difference statistics in Table 3 are based upon POS LS ‘moving points’ (17,635 points in total) compared to elevations interpolated from our LIDAR DTM. The DTM is gridded at 5 by 5 ft. Its accuracy was scrutinized closely and reported by Reutebuch et al. (2003). The statistics in this LIDAR DTM evaluation were based upon the differences between the DTM and the elevations of a larger set of surveyed ground locations. Using a subset of 121 points under the same portion of the forest canopy where this POS LS test was conducted, we computed a mean LIDAR DTM error of 1.02 ft, a standard deviation of 0.95 ft, and minimum, maximum error of – 1.97 and 4.30 ft. Clearly, the weighted means and standard deviations for the POS LS system (Table 3) are very comparable. As Reutebuch et al. (2003) make clear, the elevation differences are small and, most likely, can be attributed primarily to the Our tables and plots were developed to help evaluate the usefulness of the POS LS instrument in a forestry context, particularly in those situations where GPS is unreliable or known to be inaccurate. Three situations are of interest: 1) How well would the instrument serve as a tool for locating specific field coordinates—a plot center, for example, or the boundary points of a unit—in real-time? 2) How well would the instrument serve as a tool for collecting and post-processing coordinates to record, for example, an existing plot center or stream bed under a riparian canopy? 3) How well would the instrument serve as a tool for collecting and post-processing the coordinates necessary to define or evaluate a terrain profile or a digital terrain model in areas of dense canopy? In the first situation, the operator would use the POS LS in a ‘real-time’ mode—out in the forest, using the real-time coordinate read-out to navigate. With the other situations, the operator would collect data and then post-process in the office to prepare an accurate coordinate file. Both Table 1 and Figure 2 demonstrate typical error patterns in runs initiated at a known point and accumulated over time in the field. The total time lapse varies from 806 Table 3. Differences between LIDAR DTM and POS LS elevations while unit was in motion (excludes data collected while the POS LS unit was at rest). Elevation Difference Run Total Moving (LIDAR DTM elevation minus the POS LS elevation, ft) (no.) Points Points Avg. Stdev. RMS* Min. Max. 1 4986 3111 0.6 1.0 1.2 -3.0 4.3 2 2324 1628 0.4 0.9 0.9 -2.7 3.9 3 4222 2649 0.8 1.1 1.3 -4.1 4.2 4 2345 1434 1.5 1.0 1.8 -1.8 4.0 6 1065 710 1.1 0.7 1.3 -1.5 3.1 7 1896 1309 0.8 0.9 1.2 -1.0 4.4 8 253 154 0.9 0.4 1.0 -0.1 1.7 9 2982 2113 0.8 1.0 1.3 -2.3 3.9 10 2079 1187 0.4 0.8 0.9 -2.0 3.6 11 3033 2159 0.8 0.9 1.4 -1.5 4.5 12 1660 1181 0.5 0.9 1.0 -1.4 3.3 Weighted Avg., Stdev., RMS, Min., Max. Difference--all runs 0.7 1.0 1.3 -4.1 4.5 *Root-mean-square difference between LIDAR and POS LS elevations 26 smoothing effect of the DTM, the slight positive bias that was noted in the LIDAR DTM, the operator climbing over large logs, small random errors in the ground survey, and/ or micro-topography of the actual forest floor. Darche, M. 1998. A comparison of four new GPS systems under forestry conditions. Special Report 128. Forest Research Institute of Canada, Pointe-Claire, Quebec, Canada. 16p. CONCLUSIONS AND RECOMMENDATIONS Elosegui, P., J. Davis, R. Jaldehag, J. Johansson, A. Niell, I. Shapiro. 1995. Geodesy using the Global Positioning System: The effects of signal scattering on estimates of site position. J. Geophys. Res.,Vol.100: 9921-9934. The POS LS does seem to have great potential in forestry. In an operating mode that is typical for forestry and, unimpeded by heavy canopy cover, the POS LS post-processed data are considerably better than that of a roving GPS instrument. This is true for its ‘real-time’ mode as well. Therefore, whether a forester is navigating to a point or preparing to record and later post-process coordinate data, the POS LS offers a considerable accuracy improvement over a roving GPS instrument. Currently, the unit is quite expensive and heavy compared to conventional GPS units. However, when accurate positions under heavy canopy were needed in the past, foresters have been forced to use more labor-intensive, expensive and heavy ground survey methods and equipment. And, as happened with GPS units, it is expected that both the cost and weight of the POS LS unit will decrease as the system is miniaturized in the future. It is clear that ZUPTs are very important to the accuracy of the POS LS system, as they are the only means of compensating for IMU drift when reliable GPS signals are unavailable and known points are not nearby. However, ZUPTs will in some circumstances be an operational impediment— requiring the operator to stop too frequently and thus slow down progress of both navigation and data collection. The average time between ZUPTs in the runs of this test was 38.2 seconds, and the average time spent stopped for a ZUPT was 16.6 seconds. One can expect to erode the accuracy if the ZUPTs are less frequent; however, there are many operations in forestry where less accuracy would be acceptable. Therefore, we would recommend that a series of tests be designed to test the positional accuracy versus ZUPT frequency relationship. There are certain to be many situations where this relationship will be of interest. Farrel, J.A. and M. Barth. (1999). The Global Positioning System and inertial navigation. McGraw-Hill, New York, NY. Firth, J. and R. Brownlie. 1998. An efficiency evaluation of the global positioning system under forest canopies. NZ Forestry, May 1998: 19-25. Gillet, J., R. McCuiag, B. Scherzinger, E. Lithopoulos. (2001). Tightly coupled inertial/GPS system for precision forestry surveys under canopy: test results. First International Precision Forestry Symposium, University of Washington, College of Forest Resources, Seattle, WA, June 17-20, 2001: 131-138. Lachapelle, G. and J. Henriksen. 1995. GPS under cover: the effect of foliage on vehicular navigation. GPS World, March 1995: 26-35. Reutebuch, S., R. McGaughey, H. Andersen, and W. Carson. In press. Accuracy of a high resolution LIDAR-based terrain model under a conifer forest canopy. Canadian Journal of Remote Sensing, Vol. 29, No. 5, pp. 527-535. ACKNOWLEDGMENTS Support for this research was provided by the USDA Forest Service, Pacific Northwest Research Station, and the Precision Forestry Cooperative within the University of Washington, College of Forest Resources. The authors wish to acknowledge the Washington State Department of Natural Resources for their generous contributions, by allowing use of the test site and assistance from the Resource Mapping Section, that made this study possible. We also acknowledge Brian Johnson and Joel Gillet of Applanix Corporation for all their help with both field work and data processing. Finally, we wish to thank Andrew Cooke, University of Washington student, for all spreadsheet setup and graphics he provided for this paper. REFERENCES Curtis, R.O., D.M. Marshall, and D.S. DeBell, (eds.). In press. Silvicultural options for young-growth Douglas-fir forests: The Capitol Forest Study—establishment and first results. U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station, Portland, Oregon, General Technical Report PNW-GTR-XXX. 27 28 Ground Navigation Through the Use of Inertial Measurements, a UXO Survey MARK BLOHM AND JOEL GILLET Abstract: While portable inertial navigation systems have been successfully developed to facilitate land surveying under tree canopy, and are currently used in the Oil and Gas exploration business, their inherent cost has so far been an hindrance to their acceptance in other markets. A new generation of inertial survey instruments is being developed to keep the advantages of the previous generation, such as portability, productivity and low environmental impact, but at a lower cost, more to the level of traditional survey instruments and geodetic GPS receivers. The US army corps of engineers having identified this need after a first phase of demonstration of “Innovative Navigation Equipment and Methodologies to Support Accurate Sensor Tracking in Digital Geophysical Mapping (DGM) Surveys” during the year 2001, has financed further studies of lower cost portable inertial navigation systems by Applanix Corporation and Blackhawk Geoservices. This paper presents the joint efforts made by these two companies to define the requirements for a lighter, smaller, less expensive inertial Position and Orientation System, that can be directly integrated to existing field instrumentation (such as a Geophysical instrument), for use under canopy. This instrument could be of value to forestry applications. 29 30 Precision Forestry Operations and Equipment in Japan KAZUHIRO ARUGA Abstract: A higher level of various forest operational activities will be required to meet the higher demands on forest resources. The Japanese Forestry Agency has been developing equipment required for more efficient and precise operations. A tool consisting of GPS, an electrical compass, a laser range finder, digital calipers, and PDA has been developed to measure background information such as topography and forest conditions more easily. In order to access steep terrain of more than 30 degrees, monorails equipped with a crane or a grapple will start to be introduced into forestry. Furthermore, an autonomous monorail for transportation has been developed. According to a harvester and a forwarder, remote controlled or autonomous machines have been studied to increase forestry productivity as well as to reduce forest environmental impact. INTRODUCTION SURVEY TOOL AND SIMULATION The world’s population is over 6 billion in 2001 and is projected to be near 9.3 billion by 2050. This increase in the world population will demand more resources such as freshwater, energy, food, logs, lumber, pulp, and other forest products. Besides providing wood products, the forest has other important functions. Trees and other vegetation on forestlands remove carbon dioxide from the air and release oxygen. Streamside trees provide shade and cool water for fish and other aquatic species during hot summer months; they supply large woody debris to streams that provide and maintain fish and wildlife habitat, and they are a critical component of wetlands and river banks assisting in the protection of water quality and habitat for fish and wildlife. The forest is also a major source of outdoor recreation where people can fish, hunt and engage in other outdoor activities. A higher level of various forest operational activities will be required to meet these higher demands on resources. An increase in large and heavy forestry machines and more forest roads will be needed to increase operational productivities to meet the higher demands on resources. However, it is difficult to use existing forestry machines in mountainous areas. Even if they could be used, their operational cost would become high. Furthermore, they have had a negative impact on forest environments by causing soil disturbance and residual stand damage even on gentle slopes. Proper planning and implementation of forestry operations minimize the negative impact. The Japanese Forestry Agency has been developing equipment required for more efficient and precise operations. This paper describes this equipment that includes a forest survey tool, a forestry operation simulation tool, and remote controlled or autonomous machines. Implementation of a more efficient and precise forestry operation requires more precise and accurate data on topography and forest conditions. Topography can be measured accurately by LIDAR and much research has been conducted to develop a filter to obtain more accurate topography from raw data of LIDAR. Forest conditions include tree number, location, species, height, diameter, and volume. The Japanese Forestry Mechanization Society and the Japanese company, Timbertech, have developed the survey tool, Formas, consisting of a GPS, an electrical compass, a laser range finder, digital calipers, and PDA under a Japanese Forestry Agency project. Since these components used in Formas already exist, this project is aimed at integrating the equipment to measure location, height, and diameter and to calculate volume more easily. Though automated individual tree measurements with LIDAR were studied (Andersen et al. 2002), this project tries to develop a smaller groundbased laser detector which scans topography and trees in three dimensions. More accurate simulation will result with more precise and accurate data on topography and forest conditions. Many simulations on forestry operations have been performed throughout the world. In Japan, Sasaki simulated a mobile yarder with C++ (Sasaki and Kanzaki 1998). Also, Zhou simulated a mobile yarder and processor on steep terrain, and a harvester and forwarder on a gentle slope with GPSS (Zhou and Fujii 1995). In addition, Sakurai simulated a mobile yarder, a processor, and a forwarder (Sakurai 2001). However, these simulations are used only for research. In order to use these simulations in the forestry industry, it is necessary to collect data associated with specific sites and 31 equipment. Finally then, verification of simulations on operational sites should be conducted. FOREST ROAD AND MONORAIL Road density in the forest is 13 m/ha in Japan. Over the course of 40 years, the Japanese Forestry Agency is planning to construct roads in the forest up to 18 m/ha. Unfortunately, 18 m/ha is not high enough to conduct forestry operations with a mobile yarder and small forwarder (used typically in Japan). As forest roads must be constructed based on a forest road standard (the safest means), costs can exceed 100,000 yen/m. A low volume road, called a strip road is constructed without strict standards in order to complement the forest road. Its width is about 2 m. A small forwarder and small truck can be driven on this road. Its cost is about 10,000 yen/m. However, a strip road in the mountainous area is subject to minor landslides. In fact, strip road studies in Japan specifically mentioned that minor landslides were related to topography, vegetation, soil, climate, and road structure (Cheng et al. 2002, Suzuki and Yamauchi 2002, Yoshimura et al. 1996). Though vegetation, soil, and climate must be measured in the site, topography and road structure can be measured with high resolution DEM (e.g. from LIDAR). This is helpful to forecast where minor landslides occur and to design proper road locations. It is difficult and expensive to construct even low volume roads at many forest sites in Japan because of the steep slope (in some cases more than 30 degrees). In order to access the sites, monorails have been introduced to forestry (Nitami 2003). Many industrial monorails for agriculture Figure 2. Monorail with a crane. small monorail is about 15,000 yen/m and the cost of a large one is about 35,000 yen/m including rail material, labor, and cars (if more than 700 m long rails are constructed). All monorails can climb more than 40 degrees and their speed is about 40 m a minute. Monorails have rack and pinion traction mechanisms for slip-less movement and a dual brake system for safety, especially for downhill. In addition, monorails can be equipped with cranes and grapples to load logs (Jinkawa et al. 1998). In the end, it is important that monorails cause little disturbance to the overall forest environment. The Japanese Forestry Mechanization Society has developed autonomous monorails for transportation. FORESTRY VEHICLE The Japanese Forestry Agency has developed forestry machines to be used in mountainous forest regions. Recently, a semi-legged vehicle was introduced and modified for Japanese forestry (Aruga et al. 2001). In Europe, a harvester with four triangle shaped crawlers was produced by Valmet (Stampfer and Steinmulle 2001). These machines can maneuver in steep and rough terrain. Hence, forestry machines could be widely used on steep and rough terrains. When using a harvester and forwarder system, the forwarding cost represents approximately 10% of the forest industry’s raw material cost. Forwarders compact soil more than harvesters because forwarders move with many loaded logs. This operation’s efficiency can be improved by combining satellite navigation (GPS) and radio communication. After the harvester cuts and processes the trees, if the log positions could be transferred to a forwarder, the forwarder would not have to move around to find the trees. In this way productivity is improved because moving and judging time are shortened. The environmental impacts are also reduced because the number of vehicle passes decreases and trail areas are restricted. In addition, the use of a transport optimization algorithm for the forwarding operation would make this process even more efficient. Figure 1. Monorail with a passenger car. and civil engineering are used in Japan. Most notably, agriculture monorails are used at an orange grove on the steep terrain on Shikoku Island, Japan. Many companies produce industrial monorails. Maximum load ranges from 200 kg (Figure 1) to 5,000 kg (Figure 2). The construction cost of a 32 Figure 3. The examination of the speed sensor on national road. (The thick line indicates roads on which the speed sensor could be used.) Figure 4. The examination of GPS data communication system using cellular phones. (The thick line indicates roads on which communication was successful.) Though transportation efficiency has been improved with satellite navigation and radio communication in other industries, it is difficult to use GPS in the forest (Reutebuch et al. 1999). To counter this, a speed sensor, consisting of a GPS, yaw and pitch gyro, and an acceleration meter produced by Datatec Co., Ltd., was tested. The sensor measured vehicle positions in tunnels and on a forest road along a stream (Figure 3), locations unfavorable to the use of GPS (Figure 4). The sensor could not be used however to measure positions on strip roads where a considerable amount of slippage occurs. (The device was not produced for use on off-road vehicles.) (Figure 5, 6). As a result, it will be necessary to develop a new sensor with a GPS, gyro, an acceleration meter, and other meters for off-road vehicles (Imou et al. 2001, Mozuna and Yamaguchi 2003). As cellular phones are widely used in Japan, they could be instrumental in enhancing a vehicle navigation system. The transfer of GPS data by cellular phone was tested in the Tokyo University Forest in Chichibu. A car equipped with a GPS receiver (Trimble AgGPS124) traveled from the Tokyo University Forest office along R140 to the end of the Irikawa Forest Road. The distance between the office and the end of the Irikawa Forest Road is 25 km. GPS data obtained from the GPS receiver was downloaded to a computer in the car and transmitted from the computer to a computer in the office by cellular phone. It was demonstrated that GPS data can be transmitted anywhere except for sections with tunnels and the portion of the test forest road along a stream (Figure 4). A Japanese construction machinery company has started to equip construction equipment used as base-machines for forestry equipment with satellite communication systems. This system transfers the machine position obtained from GPS as well as operational information (such as working 33 Figure 5. The examination of the speed sensor on a forest road. Figure 6. The examination of the speed sensor on a forest strip road. 34 tion. J. Jpn. For. Eng. Soc. 13(1):3-14. (in Japanese with English summary) time and earthwork efficiency) so that a customer in an office decides when the machine needs to be maintained. This system is only used to transmit daily reports. Another Japanese forestry equipment company has developed a GPS data and message transmission system. This system transfers GPS position data (e.g. forestry workers or vehicles) to other workers, vehicles and office locations through satellite communication and Internet applications. Satellite communication systems are unique in that they can transfer information anywhere. Unfortunately, their continuous use is cost prohibitive. For this reason satellite communication systems associated with construction machines transfer information only once a day and the system developed by the Japanese forestry equipment company transfers only information during emergency situations. Certainly, if satellite communication system costs go down, we can expect its use will become widespread in the forestry industry. Imou, K., Okamoto, T., Kaizu, Y., and Yoshii, H. (2001) Ultrasonic Doppler Speed Sensor for Autonomous Vehicles. J. JSAM 63(2): 39-46. Jinkawa, M., Tsujii, T., Furukawa, K., and Fujii, T. (1998) Development and construction of the tram-car for slopes. J. Jpn. For. Eng. Soc. 13(3):183-192. (in Japanese with English summary) Mozuna, M. and Yamaguchi, H.(2003) A Study on an Autonomous Forwarder by Remote Brain Control System. Proceedings of Int. Seminar on New Roles of Plantation Forestry Requiring Appropriate Tending and Harvesting Operations: 474-479. Nitami, T. (2003) Network of Roads in the Forest with Compound Standards. Proceedings of Int. Seminar on New Roles of Plantation Forestry Requiring Appropriate Tending and Harvesting Operations: 91-95. CONCLUSIONS This paper describes the development of equipment required by more efficient and precise operations. More specifically, this equipment includes the forest survey tool, a forestry operation simulation tool, and remote controlled or autonomous machines. First, we have to gather background information such as tree location, property and topography with survey tools. Second, we must develop criteria for cost, productivity, energy consumption and environmental impact. Third, we have to evaluate equipment and systems based on these criteria using simulation tools. Once this is done, we can determine which are technically sound, economically efficient and environmentally acceptable. Finally, we must implement our decision and monitor its results. If current systems and equipment are inadequate, then modifications, improvements, or even new concepts must be investigated. Remote controlled or autonomous machines will be useful for forestry operations. However, it is most important that any forest operational activities protect the forest ecosystem. Reutebuch, S. E., Fridley, J. L., and Johnson, L. R. (1999) Integrating Realtime Forestry Machine Activity with GPS positional Data. ASAE Annual International Meeting: Paper No. 99-5037. Sakurai, R. (2001) The study on the development of mechanized logging operational system*. Ph.D. thesis, The University of Tokyo. 203pp (written in Japanese with a tentative translation by the author). Sasaki, S., and Kanzaki K. (1998) A computer simulation of yarding operation using an object-oriented model. J. Jpn. For. Eng. Soc. 13(1):1-8. (in Japanese with English summary) Stampfer, K., and Steinmulle, T. (2001) A New Approach to Derive a Productivity Model for the Harvester “Valmet 911 Snake”. Proceedings of The International Mountain Logging and 11th Pacific Northwest Skyline Symposium: 254262 (http://depts.washington.edu/sky2001/). LITERATURE CITED Suzuki, Y., and Yamauchi, K. (2000) Practical investigation on affiliate structures for prevention of disaster and degradation on low-standard forest roads. J. Jpn. For. Eng. Soc. 15(1):113124. (in Japanese with English summary) Andersen, H., Reutebuch, S., and Schreuder, G. F. (2002) Automated Individual Tree Measurement Through Morphological Analysis of a LIDAR-Based Canopy Surface Model. Proceedings of First International Precision Forestry Symposium: University of Washington, College of Forest Resources, June, 2001, pp.11-22. Yoshimura, T., Akabane, G., Miyazaki, H., and Kanzaki, K. (1996) The evaluation of potential slope failure of forest roads using the fuzzy integral -Testing the discriminant model. J. Jpn. For. Eng. Soc. 11(3):165-172. (in Japanese with English summary) Aruga, K., Iwaoka, M., Sakai, H., and Kobayashi, H. (2001) The Dynamic Analysis of Soil Deformation Caused by a Semilegged Vehicle. Proceedings of the Symposium IUFRO Group 3.11.00 at the XXI IUFRO World Congress: 1-7. Zhou, X. and Fujii, Y. (1995) Simulation of the yarding and logmaking operations system with a use of GPSS. J. Jpn. For. Eng. Soc. 10(2):243-252. (in Japanese with English summary) Cheng, P. F., Gotou, J., and Zhao, W. M. (2002) Assessing the stability of cut slopes by using soil profile pattern classifica- 35 36 Precision Forestry Applications: Use of DGPS Data to Evaluate Aerial Forest Operations JENNIE L. CORNELL, JOHN SESSIONS AND JOHN MATESKI Abstract: Aerial operations play an important role in efficient and cost-effective management of forestlands. The focus of this paper is on potential uses of precision forestry data for evaluation, planning and implementation of an aerial forest operation. Helicopters are used for aerial seeding of harvested or burned areas; application of herbicides, insecticides and fungicides; fertilization; timber harvesting; delivery of water and retardant in fire suppression efforts; transportation of crews, equipment and supplies; slash disposal; emergency medical evacuations; cone collection; tree pruning; insect and disease surveys; and for general reconnaissance. Planning and implementation of aerial forest operations with helicopters includes the safest and least-cost approach for personnel and the aircraft. A helicopter operation involved with the application of experimental minerals on stands of Douglas-fir in the central Coast Range of Oregon was used as a case study. The differential global positioning satellite data collected during application was used to evaluate empirical estimates for production and costs for one mineral ($/ton applied); to compare the influence of operational aspects on cycle time (e.g. heliport approach and departure pathways); and to develop a regression to estimate production based on operational parameters. The regression model derived from the differential global positioning satellite data collected on the operation validated the empirical estimates for helicopter production. This paper summarizes the data analyses and discusses some of the potential uses and limitations of differential global positioning satellite data for aerial forest operations. INTRODUCTION growth reductions in recent years due to the effects of Swiss needle cast disease (Filip et al. 2002). Preliminary results indicate application of specialized minerals have potential to offset growth reductions caused by Swiss needle cast. To facilitate incorporation of minerals into the soil with natural precipitation, the minerals were applied aerially during the winter and spring season to units in the project (Figure 1) (Gourley 2002). The experimental project covered nine application units. The units varied from 5 acres to 169 acres, with tree ages from 2 to 30 years. The minerals were applied from January 16, 2002 to February 1, 2002. The operation involved applications of up to six minerals at two different times. Dry material (granular and pelletized form) was applied in the winter using a bucket system and included up to five different minerals. Then, a two-stage liquid sulfur application with a spray boom configuration followed in late spring. The total 2002 acreage for the case study (dry material) application was 302 acres, with a combined amount of applied minerals of 306 tons. The mineral selected for production and cost estimates and the flight data analyses was doloprill, a pelletized dolomitic lime with a molasses binder consisting primarily of calcium with a 9% magnesium content. The application rate ranged from 1000 – 2273 lbs/acre. The application of large amounts of minerals during the winter season potentially increases road improvement and Aerial operations play an important role in efficient and cost-effective management of forestlands. Helicopters have historically been used in the forested environment for a variety of management activities from application and harvesting operations to fire suppression and reconnaissance. Planning for aerial forest operations has traditionally sought the use of the safest and least-cost approach for the helicopter, crew and support personnel. The focus of this paper is the use of precision forestry data collected for a case study of a mineral application operation. The case study served as the model for development of a planning approach to consider the economics of several options for the transportation and aerial application of specialized minerals to stands of Douglas-fir (Pseudotsuga menziesii) in the Coast Range of Oregon (Cornell 2003). CASE STUDY A forestland manager identified a practical need for an operations planning approach to minimize costs for transportation and aerial application of experimental minerals on private forestland. Stands of Douglas-fir regeneration in the Coast Range of Oregon have experienced significant 37 N 6 5 4 8 Newport 7 3 9 Highway 20 2 Corvallis 1 Philomath LEGEND Unit # Acres 1 2 3 4 5 6 7 8 9 20 12 37 12 169 10 5 15 22 Scale:1 inch Tons Applied 10 20 41.3 12 172.5 12.2 3.3 10 25 8 miles Figure 1: Vicinity map for case study mineral amendment project. eration/deceleration airspeed, a maximum ferry speed, a total ferry distance, a maximum payload and a calibrated application rate. Calculations for the model were in spreadsheet format. The cycle elements for the empirical model were the same for the flight data collection for the delay-free cycles. Details of the empirical model and the cycle components are discussed in another paper (Cornell 2003). The production rate for the helicopter was assumed to be the controlling factor for overall production on the operation. Production rates for the agricultural trucks were based on the lowest estimated production rate for a helicopter for a given heliport. maintenance costs for the landowner and thus can increase overall project costs. The increased road costs are a result of ground transportation of the large quantities of minerals to heliports. The landowner was interested in evaluating potential transportation and application scenarios to minimize the total project costs and facilitate future planning on a landscape scale for an aerial operation. The objective of the research was to develop a planning approach using mixed-integer linear programming techniques to evaluate a combination of heliports, aircraft, and transportation options to minimize overall project costs under the constraints of operational safety while meeting the forest landowner’s objectives. Key cost and production components for the operation were identified and formulated for the mathematical model on a common basis. The approach was suitable for the case study and could be applicable to other aerial forest operations involving helicopters as well. Empirical production estimates for application were used to estimate ferry and application costs for the Bell 47G3 (B47G3) helicopter in the modeled transportation network options. Data collection in the field application phase of the case study supported formulation of the production and cost estimate framework for the project. DATA COLLECTION Previous experience with collecting production data in logging operations has shown it to be challenging due to variability of the operating environment (Olsen and Kellogg 1983). The same is true for aerial forest operations because the activity is outdoors and performed under varying weather conditions and changing geographic locations. The overall objective for data collection was to obtain production and flight information for the B47G3 helicopter (Figure 2) applying the doloprill and to use that information to verify empirical production estimates and predict production and costs of future operations. The helicopter had a maximum hover-out-of-ground-effect1 payload of 1000 pounds for this operation. Detailed flight time data was collected using the Ag- EMPIRICAL PRODUCTION MODEL The empirical production model estimated a total cycle time for the helicopter based on an average forward accel38 Nav®22 differential global positioning satellite (DGPS) system installed on the B47G3 helicopter (Figure 3). The latitude, longitude received from Wide Area Augmentation System (WAAS) transmitters and used a WGS84 datum for map coordinates. The AgNav®2 system recorded a unit map with flight lines and application swaths flown by the helicopter (Figure 4). The flight and application data were cross-referenced with Figure 2: Bell 47G3 non-productive (non-aphelicopter reloading the plication) flight times and application bucket from the activities recorded with agricultural truck on the case the shift level information study application project. to identify delay-free cycles. The cycle data for selected samples from the Ag-Nav®2 DGPS system were converted from binary code data files to Figure 3: Ag-Nav®2 display screen and directional light bar installed in the Bell 47G3 helicopter. a spreadsheet format using the CROP2TXT software from Ag-Nav, Inc. The converted flight data were used to calculate total flight path distance per cycle; to estimate a maximum difference in elevation per cycle; to calculate average acceleration and deceleration; to determine reload time; and determine time and distance for the helicopter to transition and accelerate to ferry speed and decelerate to flare to load minerals (average airspeed below 25 mph). Unit boundary Flight path Application swath Heliport Figure 4: Ag-Nav®2 map with flight paths and application swaths for unit 3 of case study. 1 Hover-out-of-ground-effect (HOGE) payload is the maximum payload the helicopter can lift for a given air temperature, altitude, and wind velocity when the helicopter is at a vertical distance from the ground greater than one-half the rotor diameter. 2 The mention of commercial operators and trade names of commercial products, equipment and software in this paper does not constitute endorsement or recommendation by the authors or Oregon State University. 3 Wide Area Augmentation System 39 80 Empirical Model 70 DGPS Data 60 Airspeed (mph) 50 40 30 20 10 Helicopter turning to begin new swath 0 0 2000 4000 6000 8000 10000 12000 Total Flight Path Distance (feet) Figure 5: Empirical model estimate compared to 10 random samples of actual flight data for unit 6 on case study (Airspeed versus total flight path distance). 170 Unit 5 model output from y = 0.0088x + 70.557 R2 = .5822 n = 25 160 Total Cycle Time (seconds) 150 Unit 3 model output from y = 0.0094x + 59.832 R2 = .6131 n = 25 140 130 120 110 Unrestricted approach and departure flight path Restricted approach and departure flight path 100 Data Range Overlap (4540 ft - 6750 ft) 90 80 2500 3500 4500 5500 6500 7500 8500 9500 10500 11500 Total Flight Path Distance (feet) Figure 6: Effect on cycle time of restricted versus unrestricted heliport approach and departure flight path (Lines fitted from 25 random delay-free cycles for each unit). 40 Regression Line from Group 1 DGPS Data 250 95% Prediction Band Using Scheffe's Multiplier Group 2 DGPS Data Total Cycle Time (seconds) 200 Empirical Model Estimates 150 Regression line equation: Yest = 0.0093Xest + 66.9119 R2 = 0.7686 n1 = 50 100 50 Range of Validation (3900 ft - 10,500 ft) 0 0 2000 4000 6000 8000 10000 12000 14000 16000 Total Flight Path Distance (feet) Figure 7: Simple linear regression to predict a total cycle time as a function of the estimated total flight path distance with a 95% prediction band and range of validation data (Based on two random, exclusive, independent 50-cycle samples from entire case study). DATA ANALYSES the estimated total flight path distance of the helicopter per cycle. The empirical model estimates for helicopter cycle time were within the 95% prediction band for the simple linear regression. It was assumed the regression could be used as a surrogate for the empirical model to adequately predict a total cycle time for a similar operation with similar parameters. There were three objectives for the flight data analyses: first, to compare empirical model estimates for delay-free cycle time to actual flight data; second, to compare the effect of a restricted heliport approach and departure path on cycle time; and third, to derive a statistical relationship from the flight data to estimate cycle time for project planning and cost estimations. The first analysis compared empirical model estimates for selected field units of helicopter cycle times and production to actual flight data (Figure 5). The empirical model appeared to give a reasonable approximation of total cycle time as a function of total flight path distance. The second analysis compared the effect of restricted or obstructed helicopter approach and departure pathways on total cycle time for two heliports (Figure 6). Restricted heliports with obstacles and/or steep approach and departure paths can increase operational costs and risk to the pilot, aircraft and support personnel. The helicopter has to ascend loaded and/ or maneuver around to clear obstacles before acceleration to ferry or application speed, or deceleration to land or load. The heliport with the obstructed approach and departure pathway had an overall increase in total cycle time for the data sample. The third analysis developed a simple linear regression from a sample of recorded flight data to predict an average total cycle time for the helicopter (Figure 7). A second independent data sample was used to validate the regression. Through the use of an extra-sum-of-squares Ftest, the single most significant independent variable was USES OF PRECISION DATA Data Analysis for Case Study A comparison of empirical model estimates to actual flight data illustrated that the model gave a good representation of the actual flight pattern of the helicopter during a cycle for the given operational scenario. The flight data analysis indicated heliport approach and departure flight path access had an effect on the production and cost of the helicopter. Payload capability is perhaps the performance characteristic of greatest economic importance for some operations (Stevens and Clark 1974). In general, reduced payloads from restricted heliports decrease production and increase cost. Heliport access has a direct effect on risk management for an operation. Restricted heliports can reduce pilot visibility and increase the performance demands on the helicopter (Stevens and Clark 1974). If the straightline flight path gradient is greater than 29%, it is not safe to fly and this effect is exaggerated on short distances (O’Brien and Brooks 1996). The simple linear regression developed from the flight 41 25 Tons/hour = [(3600 secs/hr)/(Yest)] * [(payload in pounds)/(2000 lbs/ton)] Production (tons/hour) 20 15 10 Payloads 5 800 pounds 1000 pounds 0 0 2000 4000 6000 8000 10000 12000 Total Flight Path Distance (feet) Figure 8: Projected production trend for Bell 47G3 helicopter based on the regression model in Figure 7. tions and materials. Figure 8 illustrates the production trend for the B47G3 helicopter for two payloads over a range of total flight path distances using the regression to estimate total cycle time. Figure 9 illustrates the cost trend for the helicopter for two payloads over a range of total flight path distances based on the production estimates from Figure 8. data predicted a total cycle time in seconds as a function of the estimated total flight path distance in feet: Yest = 0.0093Xest + 66.9119 where: Yest = prediction estimate for average total cycle time (seconds) Xest = estimated average total flight path distance (feet) Other Potential Uses Aerial operations are well suited to the use of DGPS systems for data collection and analysis of the operation. Unlike most other forestry applications where this technology is used, the aerial operation is above the forest canopy where satellite signal reception is unimpeded and the system provides an abundant source of precise positional data throughout the entire operation. Compared to manual methods for time study data collection, this technology does not require constant visual contact with the helicopter. The maps from the DGPS system can serve as a visual record for an application or operation. The real-time headsup display of the operation can assist pilots with a consistent and even distribution of materials on a field unit, plan optimum flight patterns and approaches, and help delineate boundaries and buffers. Unit maps and coordinates can be downloaded into the navigation system ahead of time, reducing the flight time used to digitize boundaries and helping to identify the field units and heliports on the ground. The flight data in conjunction with field tests for application efficacy could assist in the validation of aerial productivity models previously developed for aerial application operations (Ghent 1999; Potter et al. 2002; Ray et al. 1999; Wu et al. 2002). In addition, flight data may be used to evaluate and provided a reliable estimation of average total cycle time within the data range indicated. The regression had a R2 = 0.7686, with a residual standard error of 12.64 on 48 degrees of freedom. Due to compound uncertainty in estimating several means simultaneously, the Scheffe’ method was used to construct the 95% prediction band (Ramsey and Schafer 1997). The regression had a good fit within the range of data in the samples. The estimated cycle time can be used to calculate helicopter production for an average payload of doloprill for a range of flight path distances within the parameter and operational limits of the project. Application of Regression If it is assumed that small changes in parameters, such as payload, do not substantially affect helicopter performance, the regression may be used to quickly generate general relationships and trends for helicopter production and costs over a range of flight path distances. These types of estimates may be useful to plan projects with similar condi42 60 $/ton = [$/hour (direct operating cost)] * [1/(tons/hour)] 50 Helicopter Cost ($/ton) 40 30 Payloads 20 800 pounds 1000 pounds 10 0 3000 4000 5000 6000 7000 8000 9000 10000 11000 Total Flight Path Distance (feet) Figure 9: Projected cost trend for Bell 47G3 based on production estimates in Figure 8. an increase in acceleration capacity. A reduced payload decreases the performance demands on the helicopter and can improve maneuverability of the aircraft, offsetting the potential production decrease when not flying with a maximum payload. Although the regression may be useful to help estimate production and costs for an operation, a knowledgeable person is still needed to evaluate the operation and identify operational limitations and hazards (e.g. access, heliports, power lines, pilot experience with an operation, etc.) that can influence helicopter productivity and project management. and validate production models for other types of aerial operations, such as heli-logging (Giles and Marsh 1994; Lyons et al. 1999; Sessions and Chung 1999). Flight data analysis may also assist managers and pilots with evaluation of safe and efficient use of the helicopter for an operation. Visualizing a similar operation prior to actual field application can help clarify procedures, processes and potential hazards for a pilot and crew unfamiliar with a certain type of operation. This method may also assist the experienced manager and pilot by providing a slightly different perspective to identify approaches for a new operation to improve safety and efficiency. Other Considerations LIMITATIONS Although the flight data collected with the DGPS system has been determined to be precise, it is the responsibility of the user to determine if the data and information are accurate. Additional record keeping on an operation (such as shift level production information) can be used to cross-reference loads, time, etc. and check for delays or other situations that may impact use of data for analysis or interpretation. On the case study, for each mineral applied on the field units and each cycle an observer recorded the time the bucket was loaded, application time, the payload and any delays. The unit map generated by the DGPS system may also have discrepancies between what is shown on the map and the actual application. For example, the bucket may be empty, but if the pilot does not release the switch the DGPS map will indicate the area that has been flown over while the switch is activated had received an application, when it had not. Case Study Data For the flight data analysis, additional cycle data are needed to check the reliability of the regression model beyond the limits and conditions of the data sets and case study. Data parameter limits included: regression data limits with a total flight path distance range from 1,900 feet to 10,500 feet; the second (validation) data set limits with a total flight path distance range from 3,900 feet to 14,250 feet; one type of mineral; the B47G3 helicopter with a skilled pilot; and the weather and operations conditions of the case study. An assumption for use of regression is all of the data set parameters are static as one variable of interest is changed over a projected range of conditions (e.g. equivalent helicopter performance while varying payload over a range of flight path distances). In a practical application, a pilot may adjust (increase) production for a reduced payload with 43 Another source of discrepancy can arise from boundaries that are digitized in flight. Although the digitized coordinate locations are precise, the mapped boundary consists of a series of straight lines between digitized points. Care must be taken to not cut off corners on the unit during the digitizing phase of the mapping operation. Gourley, M. 2002. Personal communication. Forester, Starker Forests, Inc., Corvallis, Oregon. Lyons, K., J. McNeel, J. Nelson, and R. Fight. 1999. Spatial modeling of helicopter logging in dispersed and aggregated partial cutting systems. In proceedings of the International Mountain Logging and 10th Pacific Northwest Skyline Symposium. March 28 – April 1. Eds. J. Sessions and W. Chung. Corvallis, Oregon. FUTURE CHALLENGES AND RESEARCH NEEDS O’Brien, S. and E.J. Brooks. 1996. A course filter method for determining the economic feasibility of helicopter yarding. Engineering Field Notes – Engineering Technical Information System. USDA Forest Service. Volume 28. 12 p. The equipment used for DGPS data collection on the case study has proven to be precise and efficient for gathering production information. However, initial capital investment in the system is considerable, and requires a substantial commitment by the operator of an additional investment in personnel training. Operation managers need to recognize the potential for added benefits of using this precision forestry tool to enhance overall operations safety and improve project efficiency where possible. With additional data collection and analysis, other regression equations and mathematical models could be developed for different pilots, helicopter types, materials, applications, flight path distances and operational conditions to estimate production and costs. Automated flight data recording systems, such as the Ag-Nav®2 DGPS system, could be used to evaluate flight characteristics and performances under varying circumstances to develop production relationships and cost estimates to assist in project planning. Olsen, E., and L.D. Kellogg. 1983. Comparison of time-study techniques for evaluating logging production. Transactions of the ASAE, Vol. 26, No. 6. 1665-1668, 1672. Potter, W.D., Ramyaa, J. Li, J. Ghent, D. Twardus, and H. Thistle. 2002. STP: an aerial spray treatment planning system. In proceedings IEEE SoutheastCon, 2002. pp. 300-305. Ramsey, F.L. and D.W. Schafer. 1997. The statistical sleuth, a course in methods data analysis. Duxbury Press. Wadsworth Publishing Company. Belmont, California. Ray, J.W., B. Richardson, W.C. Schou, M.E. Teske, A.L. Vanner, and G.C. Coker. 1999. Validation of SpraySafe Manager, an aerial herbicide application decision support system. Canadian Journal of Forest Research, 29: 875-882. REFERENCES Reynolds, R.D. 1999. Three GPS-based aerial navigation systems for forestry applications. Forest Engineering Institute of Canada. Field Note No.: Silviculture-118. October. Vancouver, British Columbia. Ag-Nav®2 by Ag-Nav, Inc. 1999. Newmarket, Ontario, Canada. Cornell, J. 2003. Aerial forest operations: mineral amendment project. M.For. paper, Oregon State University, Forest Engineering Department, Corvallis, Oregon. 259 p. Sessions, J. and W. Chung. 1999. Optimizing helicopter landing location – a preliminary model. In proceedings of the International Mountain Logging and 10th Pacific Northwest Skyline Symposium. March 28 – April 1. Eds. J. Sessions and W. Chung. Corvallis, Oregon. pp. 337 – 340. Filip, G., A. Kanaskie, K. Kavanagh, G. Johnson, R. Johnson, and D. Maguire. 2000. Silviculture and Swiss Needle Cast: research and recommendations. RC 30. OSU College of Forestry. Corvallis, Oregon. Stevens, P.M. and E.H. Clarke. 1974. Helicopters for logging, characteristics, operation, and safety considerations. USDA Forest Service General Technical Report PNW-20. Pacific Northwest Forest and Range Experiment Station. Portland, Oregon. Ghent, J. 1999. Development of an aerial productivity and efficiency model for large-scale aerial treatment programs. USDA Forest Service Research Proposal. R8-2000-02. Giles, R., F. Marsh.1994. How far can you fly and generate positive stumpage in helicopter salvage logging? Advanced Technology in Forest Operations: Applied Ecology in Action. Oregon State University. Portland and Corvallis, Oregon. pp. 231-236. Wu, L., W.D. Potter, K. Rasheed, J. Ghent, D. Twardus, H. Thistle, and M. Teske. 2002. Improving the genetic algorithm performance in aerial spray deposition management. In proceedings IEEE SoutheastCon, 2002. pp. 306 – 311. 44 Estimating Forest Structure Parameters on Fort Lewis Military Reservation using Airborne Laser Scanner (LIDAR) Data HANS-ERIK ANDERSEN, JEFFREY R. FOSTER, AND STEPHEN E. REUTEBUCH Abstract: Three-dimensional (3-D) forest structure information is critical to support a variety of ecosystem management objectives on the Fort Lewis Military Reservation, including habitat assessment, ecological restoration, fire management, and commercial timber harvest. In particular, the Forestry Program at Fort Lewis requires measurements of shrub, understory, and overstory canopy cover to monitor vegetation response to various management approaches. At present, these measurements are acquired through field-based procedures, which are relatively costly and time-consuming. The use of remotely sensed data, such as airborne laser scanning (LIDAR), has the potential to significantly reduce the cost of acquiring these types of measurements over large areas. As an active remote sensing technology, LIDAR provides direct, three-dimensional measurements of the forest canopy structure and underlying terrain surface. LIDAR-based cover measurements can be related to forest vegetation cover through a mathematical function based upon the Beer-Lambert law, which accounts for scanning geometry and vertical foliage density. This study was carried out to determine the utility of small-footprint, discrete-return LIDAR for estimation of forest canopy cover at Fort Lewis. LIDAR-based structural measures were compared to spatially-explicit field measurements acquired from inventory plots in five forest stands representative of the various forest types at Fort Lewis and a variety of terrain. Results indicate that LIDAR-based cover estimates for overstory and understory are generally related to field-based estimates. INTRODUCTION surements of vegetation cover, inventory costs could be significantly reduced through the use of remote sensing technology. In particular, actively-sensed airborne laser scanning (LIDAR) technology has the potential to provide information relating to spatial structure throughout the depth of the forest canopy and understory. Previous studies have shown that large-footprint, continuous- waveform LIDAR data can be used to characterize the vertical distribution of canopy foliage (Harding et al., 2001; Lefsky et al., 1999; Means et al., 1999). Researchers have related the vertical distribution of small-footprint, first-and multiple-return LIDAR data to empirical- and model-based estimates of leaf area distribution within Pacific Northwest forests (Magnussen and Boudewyn, 1998; Andersen, 2003). Other studies have shown that quantitative measures derived from the vertical distribution of small-footprint, discrete return LIDAR data are related to important stand parameters, such as volume, height, and biomass (Means et al., 2000). While LIDAR-derived measures of canopy cover have been used as independent variables in estimation of forest stand parameters (Means et al., 2000), the utility of LIDAR for differential characterization of canopy and subcanopy forest structure components has not been assessed. In this paper, a methodology for measurement of vegetation cover Forests are structured as complex systems in three-dimensional (3-D) space. The 3-D structural organization of forest canopies is the primary determinant of the understory light regime, micro-climate, and habitat structure. In the Pacific Northwest, the vertical distribution of canopy elements is one of the more important components describing the spatial structure of a forest stands. For example, the Forestry Program at Fort Lewis Military Reservation requires this structural information to guide an active silvicultural program designed to promote the development of forests with more diverse structures and composition, and to provide habitat for the northern spotted owl (Strix occidentalis caurina). Fort Lewis has implemented an inventory program to document and monitor the spatial structure of the installation’s forests. Three-dimensional forest structure is quantified by measuring vegetation cover of the overstory, understory, shrub, and ground layers. Vegetation cover, expressed as the proportion of the forest floor covered by the vertical projection of vegetation within a layer of the forest canopy, is a conventional measure of forest structure (Jennings et al., 1999). As the inventory program currently relies upon ocular, field-based mea45 LIDAR-Based Digital Terrain Models within discrete canopy layers using first return LIDAR data will be presented and evaluated. A filtering technique coded in IDL (Interactive Data Language version 5.5, Research Systems, Inc.) was used to identify the probable ground reflections within the last-return LIDAR data (Haugerud and Harding, 2001). An interpolation algorithm was used to generate a digital terrain model (DTM) on a grid, with a post spacing of 5×5 m, for the two areas covered by the LIDAR dataset. Figures 2a and 2b show hill-shade graphics of the DTMs. A comparison of LIDAR-estimated elevation from the DTMs to the elevations of 225 topographic survey points indicated a mean absolute error of -0.14 meters and a root mean square error (RMSE) of 1.00 meters for the southwestern Fort Lewis LIDAR DTM. A similar comparison for the northwestern Fort Lewis LIDAR DTM (244 topographic survey points) indicated a mean absolute error of 0.05 meters and an RMSE of 0.72 meters. STUDY AREA AND DATA Study Sites Within Fort Lewis, Washington Five stands considered to be representative of the variety of forest types present at Fort Lewis were selected as study areas for the project. The first two stands were located in the southwestern portion of Fort Lewis on an old recessional moraine of the Vashon Glaciation with hummocky topography and location variation in vertical relief of ca. 10 m. Area 1 was a 65-year-old mixed red alder/Douglas-fir stand, and Area 2 was a 75-year-old Douglas-fir stand. The other three stands were located on flat glacial outwash. Area 3, ca. 3 km southeast of Areas 1 and 2, was an 85-year-old mixed white oak/Douglas-fir stand. Areas 4 and 6 were 95year-old Douglas-fir stands in the northeastern portion of Fort Lewis. (Area 5 was in a prairie and therefore was not used in this study). Approximately 35 plots were located within each stand (169 total) to validate the remote sensing estimates (Figure 1). These plots were established in a systematic pattern of clusters to ensure a well-distributed sample within each stand type. Plot coordinates were established by a highly accurate total station topographic survey. An additional 300 topographic check points were established in these stands to assess the accuracy of the LIDAR digital terrain models. Field Cover Data To compare LIDAR-based measures of vegetation cover to conventional inventory metrics, field-based observations of vegetation cover were acquired at each of the 169 plots located in the five stands. Following the established field protocol of the Fort Lewis inventory program, ocular estimates of vegetation cover within the overstory, understory (1.8 m - base of overstory), and shrub (0.46 m – 1.8 m) were made at each plot (Figure 3). Overstory and understory cover were estimated for an 809 m2 circular plot, and shrub and ground cover for an 81 m2 circular plot. It should be noted that while the base of the upper and lower boundaries of the shrub and ground layers are at fixed heights in the inventory protocol, the height of the base of the overstory layer is a local characteristic of forest structure and was subjectively estimated at each plot. Cover was defined as the proportion of the total area “filled” by the two-dimensional (2-D) vertical projection of tree crowns and shrubs onto the ground. LIDAR Data LIDAR data were acquired over two 50-km2 areas on Fort Lewis in August, 2000 (fig. 1). These data were acquired with an Earthdata Aeroscan laser scanning system operating from a fixed-wing platform. System specifications and flight parameters are shown in Table 1. Table 1. System specifications for Earthdata Aeroscan LIDAR system. Pulse rate Ground spacing between pulses Laser wavelength Scan pattern Pulse length Attitude precision Range resolution Range accuracy 15,000 per second 1 meter (nominal) 1.064 µm Sinusoidal 12 ns 0.004 degrees 3 cm 2-4 cm Figure 3. Canopy layers used for cover estimation. 46 Area 4 Fort Lewis Military Reservation # ## # # # # # ## # ## # # ## # # # # # ## ### Are a 4 Fort Le wis Mil ita ry Rese rva tion # ## # ## # Ar ea 6 Ar ea 1 # # ## # ## # ## #### ## # ### # ## # # # ### # ## # # # ## # # # ## ## ## Are a 2 # # # ## Ar ea 3 Area 6 Area 1 Area 2 # ## # ## # # ## # ## # ## ## ## # # # ## # # ## ## ### ## # # # ## # ### ## # # # ## # # # ## ## # #### ## # # # ## # ## ### # ## #### # # Area 3 Figure 1. LIDAR coverage and study areas within Fort Lewis Military Reservation, Washington. (a) Southwest area (b) Northeast area Figure 2. LIDAR-based DTMs (5-m resolution) within Fort Lewis Military Reservation. 47 METHODS To convert LIDAR coordinate data into vegetation height data, the elevation of the underlying terrain (interpolated from the LIDAR DTM) was calculated for each LIDAR return; then, this terrain elevation was subtracted from the LIDAR return elevation to yield vegetation height. Vertical Point Quadrat Sampling Ground-based measures of canopy cover are typically based upon a sample of measurements acquired with a vertical sighting instrument. The percent cover is computed as the proportion of sample points where the sky is obscured by vegetation (Jennings et al., 1999). As Jennings noted, a limitation of this approach is that it is highly susceptible to sampling error. When the vertical heights from the ground to first contact with vegetation within each canopy layer are measured above each sample point, the canopy cover for a given canopy layer can be estimated as the ratio of the number of measured heights within a layer to the total number of sample points. This approach has been used to estimate vertical foliage distributions and is termed vertical point quadrat sampling (Ford and Newbould, 1971). When an optical range finding device, such as a laser, is used to measure the height to first leaf contact, this sampling technique is termed optical point quadrat sampling (MacArthur and Horn, 1969; Aber, 1979; Radtke and Bolstad, 2001). Figure 4. LIDAR-based cover estimation for overstory, understory, and shrub layers. Vertical lines represent LIDAR pulses. LIDAR-Based Cover Estimation If the geometry of the laser range-finding is inverted, an estimate of cover within a given layer of the canopy can be generated from LIDAR data, where cover within each layer is calculated as the ratio of the number of first return LIDAR reflections within a layer to the total number of LIDAR pulses entering the layer (Figure 4). A LIDAR-based cover estimate, based on first return data, was generated for overstory, under-story, and shrub layers for each of 169 plots Although field cover estimates for the shrub layer were based upon an 81 m2 plot, all LIDAR estimates were based upon an 809 m2 plot to maintain an adequate sample area. Estimates based upon the larger area will not bias the results. A single value for height was used to characterize the base of the overstory within each stand. A K-means clustering algorithm was applied to the LIDAR data within each forest stand to estimate the height that separated the understory and overstory layers (Mardia et al., 1995). This height was 10.0 meters for Area 1, 21.4 meters for Area 2, 7.5 meters for Area 3, 23.7 meters for Area 4, and 21.3 meters for Area 6. Figure 5 is a 3-D representation of a plot within Area 2 (75-year-old Douglas-fir stand) created using the Stand Visualization System (McGaughey, 1997). Figure 6 shows a 3-D perspective view of the distribution of first return LI- Figure 5. Visualization of 809 m2 plot within Douglas-fir stand. DAR data used to generate LIDAR-based cover estimates. If all LIDAR measurements were acquired at nadir, there is a linear relationship between the LIDAR-based cover measurement, generated using the theory developed in the previous section, and the field-based cover estimate for each layer. In practice, due to scanning geometry, LIDAR measurements are acquired at some off-nadir angle (-25 degrees to +25 degrees in the Aeroscan system used in this study). This off-nadir angle affects the probability of a laser pulse passing through a given layer of vegetation and influences the functional form of the relationship between a LIDARbased measurement of cover and the cover estimate observed in the field. 48 and shrub) was held fixed by randomly assigning a value of either 0 or 1 to each layer until the two-dimensional (2-D) projection of the “filled” area for each layer equaled the specified cover percentage. A specified density in the vertical dimension was also randomly allocated throughout the layer, while keeping the 2-D projection of cover fixed for each layer. The paths of LIDAR pulses traveling at some specified off-nadir angle through the 3-D array were calculated and the coordinates (x, y, height) for the “first vegetation contact” (i.e. first encounter with a cell with code “1”) were recorded as simulated first return LIDAR measurements. A simulated LIDAR-based cover estimate was then generated using the approach described in the previous section. These simulated LIDAR estimates were then compared to the specified, fixed (“simulated field”) cover values used in filling the 3-D array. Figures 7 and 8 show the influence of the off-nadir angle on simulated LIDAR-based cover estimates. Figure 7 shows the cover estimates with the off-nadir angle fixed at zero. Perhaps not surprisingly, there is a linear relationship between LIDAR- and field-based cover estimates. Figure 8 shows the relationship between LIDAR-based and field-based cover estimates when the off-nadir angle for each pulse is a random draw between -25 and +25 degrees and the foliage density within each layer is randomly chosen. Clearly the relationship is no longer linear, and appears to follow a relatively smooth curve. While this effect is apparent in the overstory and understory layers (Figures 8a Figure 6. Distribution of first return LIDAR data within 809 m2 plot within Douglas-fir stand. Simulation of LIDAR Data and Cover Estimates A simulation approach was used to investigate the effect of scanning geometry and foliage density on the relationship between field-based and LIDAR-based cover estimates. In this simulation, a 3-D array was used to simulate the 3-D “envelope” containing the canopy vegetation within a 100×100×60-meter area of forest. The cover within each layer of the canopy (i.e. overstory, understory (to 25 meters), 1.0 1.0 0.8 0.8 0.6 0.4 0.2 Field-based Shrub Cover Field-based Understory Cover Field-based Overstory Cover 1.0 0.6 0.4 0.2 0.3 0.5 0.7 0.9 0.1 LIDAR-based Overstory Cover 0.4 0.0 -0.2 0.1 0.6 0.2 0.0 0.0 0.8 0.3 0.5 0.7 0.1 0.9 0.3 (a) Overstory 0.5 0.7 0.9 LIDAR-based Shrub Cover LIDAR-based Understory Cover (b) Understory (c) Shrub 1.0 0.8 0.8 0.6 0.4 0.6 0.4 0.2 0.2 0.0 0.0 0.1 0.3 0.5 0.7 LIDAR-based Overstory Cover (a) Overstory 0.9 1.0 Field-based Shrub Cover 1.0 Field-based Understory Cover Field-based Overstory Cover Figure 7. Simulated cover estimates with off-nadir of 0 degrees. 0.8 0.6 0.4 0.2 0.0 0.1 0.3 0.5 0.7 LIDAR-based Understory Cover (b) Understory 0.9 0.1 0.3 0.5 0.7 LIDAR-based Shrub Cover (c) Shrub Figure 8. Simulated cover estimates with off-nadir angles ranging between -25 and +25 degrees. 49 0.9 vertical foliage density. Figures 9 and 10 show the influence of varying the vertical density of foliage. Figure 9 shows the relationship between simulated field- and LIDAR-based estimates of cover with a relatively high density of foliage in the vertical dimension, while Figure 10 shows this relationship for a low density of foliage, with fitted curves superimposed. These graphics indicate that both off-nadir angle and vertical foliage density will influence the relationship between field- and LIDAR-based estimates of cover. It also appears that the mathematical functional form can be adequately represented by the logarithmic model based upon the BeerLambert law given above, where the parameter of the function will represent the effects of scan angle and foliage density. and 8b) it is not evident in the shrub layer (Figure 8c), where the relationship exhibits a more linear form. The form of the mathematical function describing the relationship between LIDAR-based cover and field-based cover can be obtained from the principles of radiative transfer theory (Martens et al., 1993). Martens and others showed that the relationship between leaf area index and gap fraction, or the ratio of the amount of light beneath a canopy layer to the amount of light above a canopy layer, is given by the Beer-Lambert law, with the following form: LAI = (-1/k)ln(gap fraction), where k is the extinction coefficient governing the attenuation of light as it passes through the canopy. RESULTS In the context of cover estimation, the relationship between the field-based cover estimate (i.e., 2-D projection of vegetation onto terrain surface) and LIDAR-based cover can be estimated by the following function: LIDAR-based cover estimates obtained using first return LIDAR data were compared to the field-based cover estimates for the 169 plots on Fort Lewis. The relationship between LIDAR- and field-based estimates for each forest type was quantified using regression analysis, with the results shown in Figures 11-14. The coefficient of determination (r2) of each regression model is shown as well. The results indicate that LIDAR-based cover estimates for overstory and understory are generally related to fieldbased estimates. There does not appear to be a significant relationship between LIDAR- and field-based estimates of shrub cover for any forest types. Field cover = (a)ln[1/(1 – LIDAR cover)] 1.0 1.0 0.8 0.8 0.8 0.6 0.4 0.2 0.0 Field-based Shrub Cover 1.0 Field-based Understory Cover Field-based Overstory Cover where a is the parameter related to the extinction coefficient in Beer-Lambert law shown above, and cover values are expressed as fractions. Density of vegetation in the vertical dimension will also influence this relationship between field- and LIDAR-based measures of cover. The parameter a in the above function will represent the combined effect of off-nadir angle and 0.6 0.4 0.2 0.4 0.6 0.8 0.4 0.2 0.0 0.2 0.6 0.0 1.0 0.1 0.3 LIDAR-based Overstory Cover 0.5 0.7 0.9 0.1 0.3 LIDAR-based Understory Cover (a) Overstory 0.5 0.7 0.9 LIDAR-based Shrub Cover (b) Understory (c) Shrub Figure 9: Simulated cover estimates with high density of foliage in vertical dimension. 1.2 Field-based Understory Cover Field-based Overstory Cover Field-based Shrub Cover 1.2 1.0 0.8 0.6 0.4 0.8 1.0 0.8 0.6 0.4 0.4 0.2 0.2 0.0 0.0 0.0 0.1 0.3 0.5 LIDAR-based Overstory Cover 0.7 (a) Overstory 0.9 0.2 0.4 0.6 LIDAR-based Understory Cover (b) Understory 0.8 1.0 0.1 0.3 0.5 (c) Shrub Figure 10: Simulated cover estimates with low density of foliage in vertical dimension. 50 0.7 LIDAR-based Shrub Cover 0.9 (a) Overstory (r2 = 0.63) (b) Understory (r2 = 0.53) (c) Shrub (r2 = 0.15) Figure 11: Field-measured vs. predicted cover within 95-year-old Douglas fir stands (Areas 4 and 6). Dashed line shows 1:1 relationship. (a) Overstory (r2 = 0.63) (b) Understory (r2 = 0.53) (c) Shrub (r2 = 0.15) Figure 12: Field-measured vs. predicted cover within 65-year-old mixed red alder/Douglas-fir stand (Area 1). creasing the sample size and increasing the error of cover estimation. Results here indicate that sampling error is the primary source of variation in the estimation of shrub cover across all forest types. The results obtained from simulations and field data suggest that the interaction of off-nadir LIDAR scanning geometry and the vertical distribution of canopy foliage introduces a significant source of variability in LIDAR-based cover estimation. The geometry of LIDAR sensing leads to measurements of forest struc-ture that are more representative of 3-D canopy density than 2-D (i.e. orthogonal) canopy cover. However, if vertical structure is relatively constant over a forest stand, then vertical density can be modeled, allowing for a more accurate mapping of the LIDAR-based cover estimate to the 2-D vegetation cover. It should be noted that using a single height for separating overstory from understory layers adds a significant source of variability in LIDAR-based cover estimation. Again, error will be decreased when the stand is more structurally homogeneous. In stands exhibiting extremely complex vertical structure (e.g. the mixed red alder/Douglas-fir stand in Area 1), this classification error may lead to gross errors in cover estimation. Relationships between LIDAR- and field-based estimates are strongest in the mature Douglas-fir stand on flat terrain (Figure 11) and the mixed white oak/Douglas-fir prairie stands (Figure 14). It should be noted that the relationship for the overstory and understory layers in Area 3 did not exhibit a curvilinear form, so results for this area were based upon untransformed lidar cover values (Figure 14a and 14b). Relationships are still apparent, although less strong, within the Douglas-fir stand with hummocky topography (Figure 13). Relationships within the mixed red alder/Douglas-fir stand are extremely weak (Figure 12). DISCUSSION The graphical and quantitative results indicate that LIDAR has the potential to provide information relating to vegetation cover in multiple canopy layers within forest stands. The weaker relationship between field- and LIDARbased cover estimates for the shrub layer is most likely due to sampling error. Occlusion of subcanopy vegetation by overstory and understory foliage will reduce the number of first returns penetrating to the shrub layers, effectively de51 (a) Overstory (r2 = 0.42) (b) Understory (r2 = 0.38) (c) Shrub (r2 = 0.22) Figure 13. Field-measured vs. predicted cover within 75-year-old Douglas-fir stand in an area with varied, hummocky topography (Area 2). (a) Overstory (r2 = 0.79) (b) Understory (r2 = 0.38) (c) Shrub (r2 = 0.0005) Figure 14. Field-measured vs. predicted cover within 85-year-old mixed white oak/Douglas-fir stand (Area 3). CONCLUSIONS It should also be noted that there is no assumption in this study that the field-based, ocular estimate of vegetation cover represents a “true” measurement. Although attempts were made to “calibrate” the estimates through comparisons to the estimates of other observers, these measurements are inherently subjective and susceptible to bias. In the management context of Fort Lewis, however, it has been determined that ocular estimation remains the most economically viable approach to estimating the spatial characteristics of vegetation cover efficiently and quickly over the extent of the installation. The results of this study are, therefore, intended to show the correspondence between field-based estimates, acquired using established inventory protocol at Fort Lewis, and LIDAR-based estimates, and do not represent a true assessment of the accuracy of LIDAR-based cover estimates. Even though LIDAR-based cover estimation is subject to both systematic and random errors, due to the effects of sensing geometry, occlusion, and sampling rate discussed above, it provides for objective, spatially-explicit mapping of forest vegetation cover over extensive areas. LIDAR has the potential to be an extremely useful source of data for mapping of forest structure characteristics, including canopy cover within overstory and understory layers. The increased sampling error at greater depths in the canopy, due to the occlusion effect, limits the utility of LIDAR for estimation of cover within the shrub layer. The offnadir scanning geometry of LIDAR and the vertical foliage distribution can have significant effects on the functional relationship between LIDAR-based cover measurements and field-based observations of cover based upon the 2-D projection of tree crowns. In forest areas with homogeneous structural characteristics, these factors can be modeled using a mathematical function based upon radiative transfer theory. The methodology presented in this paper will be further developed and evaluated through comparison to intensive, objective field-based canopy cover estimates acquired at the same time as the LIDAR data. A possible extension of this research would be the use of 52 multiple-return or continuous-waveform (i.e. “single photon”), small footprint LIDAR data. The use of multiplereturn data with intensity information may allow for more sophisticated and accurate modeling of foliage density and vegetation cover. A follow-up project will use these results to model the spatial patchiness of canopy cover within Fort Lewis Military Reservation to support habitat monitoring and silvicultural programs. MacArthur, R. and H. Horn. 1969. Foliage profile by vertical measurements. Ecology 50:802-804. McGaughey, R. 1997. Visualizing forest stand dynamics using the stand visualization system. In: Proceedings of the 1997 ACSM/ASPRS Annual Convention and Exposition, vol. 4, pages 248-257, Bethesda, MD, April, 1997. American Society for Photogrammetry and Remote Sensing. Magnussen, S. and P. Boudewyn. 1998. Derivations of stand heights from airborne laser scanner data with canopy-based quantile estimators. Canadian Journal of Forest Research 28:1016-1031. LITERATURE CITED Aber, J. 1979. A method for estimating foliage height profiles in broad leaved forests. Journal of Ecology 67:35-40. Mardia, K., J. Kent, J. Bibby. 1995. Multivariate Analysis. Academic Press, London. Andersen, H.-E. 2003. Estimation of critical forest structure metrics through the spatial analysis of airborne laser scanner data. Unpublished Ph.D. dissertation, University of Washington, Seattle, WA. Martens, S., S. Ustin, and R. Rousseau. 1993. Estimation of tree canopy leaf area index by gap fraction analysis. Forest Ecology and Management 61:91-108. Means, J., S. Acker, D. Harding, J. Blair, M. Lefsky, W. Cohen, M. Harmon, and W. McKee. 1999. Use of large footprint scanning airborne lidar to estimate forest stand characteristics in the western Cascades of Oregon. Remote Sensing of the Environment 67:298-308. Ford, D. and P. Newbould. 1971. The leaf canopy of a coppiced deciduous woodland: I. development and structure. Journal of Ecology 59:843-862. Harding, D., M. Lefsky, G. Parker, J. Blair. 2001. Laser altimeter canopy height profiles: Methods and validation for closed-canopy, broadleaf forests. Remote Sensing of the Environment 76:283-297. Means, J., S. Acker, B. Fitt, M. Renslow, L. Emerson, C. Hendrix. 2000. Predicting forest stand characteristics with airborne scanning lidar. Photogrammetric Engineering and Remote Sensing 66(1):1367-1371. Haugerud, R. and D. Harding. 2001. Some algorithms for virtual deforestation (VDF) of lidar topographic survey data. International Archives of Photogrammetry and Remote Sensing, XXXIV-3/W4:211-217. Radtke, P. and P. Bolstad. 2001. Laser point-quadrat sampling for estimating foliage-height profiles in broad-leaved forests. Canadian Journal of Forest Research 31:410-418. Jennings, S.B., N. Brown, and D. Sheil. 1999. Assessing forest canopies and understory illumination: canopy closure, canopy cover and other measures. Forestry 72(1): 59-73. Support for this research was provided by Fort Lewis Military Reservation, the USDA Forest Service Pacific Northwest Research Station, and the Precision Forestry Cooperative within the University of Washington College of Forest Resources. Lefsky, M., W. Cohen, S. Acker, G. Parker, T. Spies, and D. Harding. 1999. Lidar remote sensing of the canopy structure and biophysical properties of Douglas-fir western hemlock forest. Remote Sensing of the Environment 70:339-361. 53 54 Developing “COM” Links for Implementing LIDAR Data in Geographic Information System (GIS) to Support Forest Inventory and Analysis ARNAB BHOWMICK, PETER P. SISKA AND ROSS F. NELSON Abstract: In the last decade the computerized technology made significant step forward in the data manipulation, storage, design and analysis. At the same time the acquisition of spatial data experienced significant changes in the natural resources. The field sampling methods, that originally represented the only source of spatial data, have been efficiently enhanced, in some cases even replaced, with modern remote sensing sensors. The airborne laser systems are promising tools for measuring heights of ground objects with high precision. In addition, LIDAR measurements can be linked with field sampling and regression models to provide estimates of ecosystem parameters such as biomass, leaf area index, carbon and other volumetric characteristics of vegetation structure. The objective of this project is to develop component object model (COM) for the direct transfer of raw laser measurements to GIS, perform fundamental analysis in a form of close coupling strategy and compute statistics on multiple return LIDAR data. This method allows GIS analyst and LIDAR users to effectively manipulate with laser data within GIS environment. COM compliant software supports flexible applications using objects and components from different sources. The close coupling of GIS with geospatial and statistical tools strengthens the role of GIS as an interdisciplinary science. In this paper coupling strategy involves modules that simultaneously manipulate software components from the GIS application and the data analysis application. The paper explores the significance of coupling strategies for natural resource management, engineering and the forest science applications. INTRODUCTION carbon. The exhaustive and tedious efforts to measure tree heights can be readily enhanced and even replaced by the laser systems that are capable of determining tree heights with high precision. Airborne laser profiling and scanning systems intensively sample forestlands, and these height and density measures can, using regression techniques, be used to infer stand characteristics (Nelson et al. 1988). The heights of trees and canopy densities are the first step in estimating more complex biometric parameters using the laser data. Hyyppä et al. (2001) estimated stem volume using high pulse rate laser scanner data, segmentation method and regression equations with a 10.5% error. Nelson et al. (1988) developed and tested two logarithmic equations in conjunction with six laser based canopy measurements. The results indicated that the mean total tree volume can be predicted with a 2.6% accuracy of the ground value and a mean biomass within a 2% accuracy based on 38 ground plots. The Scanning Lidar Imager of Canopies by Echo Recovery (SLICER) was also frequently used to collect data in the deciduous forests on the North American continent (Lefsky et al. 1999). The quadratic mean canopy heights explained the 70% variation in the stand basal area and the 80% variation in the above ground biomass. The crucial segment in estimating the above mentioned biometric parameters either with pro- Over the past two decades LIDAR (Light Detection and Ranging) data have been extensively used in natural resources and engineering. Efforts to determine the depth of the ocean floor with high accuracy appeared to be one of the first applications involving laser systems (Hoge et al. 1980). Realizing that vegetation canopy “depth,” i.e., height and structure, might also be measured using this bathymetric LIDAR technology, Link and Collins (1981), Arp et al. (1982), and Hoge et al. (1982) reported some of the first airborne laser studies of terrestrial targets in the western hemisphere. Roughly parallel, more theoretically based, terrestrial LIDAR investigations were ongoing in the United Soviet Socialist Republic at the time (Solodukhin et al. 1977a, b; 1979; 1985; Stolyarov and Solodukhin 1987), though the Cold War precluded scientific cooperation. Over the past 20+ years, numerous researchers have demonstrated the capabilities and limitations of a wide variety of airborne laser systems for forest, rangeland, geologic, and topographic measurement and mapping. With respect to forestry, airborne LIDAR research has centered on the assessment of forest canopy heights, crown closure, internal canopy structure, and the use of these laser measures to estimate stems, basal area, wood volume, forest biomass, and 55 filing, SLICER, or other imaging laser systems is in developing regression equations that relate measured parameters on the ground with the airborne laser measurements. The newest results in Delaware (Nelson et al. 2003) indicate that biometric estimates of biomass and volume in four land cover categories (deciduous, mixedwood, conifer, and wetlands) are within 16% of U.S. Forest Service estimates statewide. The geographic information systems (GIS) have provided, until now, limited service to laser based projects. Such applications include data storage and display. More sophisticated application are associated with overlay operations. In particular, classified vegetation maps can be overlaid in a GIS with scanning laser data or transect lines (profiling lasers) in order to calculate stratified estimates of biomass and other biometric parameters. Therefore, the spatial distribution of vegetation and its covariation with laser measurements across the study area plays a significant role in the final evaluation of biometric parameters. Additional overlays with soil cover can also be useful in relating laser data to natural resource information. In this project the close coupling procedure was developed to support the automated transfer of raw laser measurements to GIS from the ArcMap module for further analysis. The linkage continued back to excel spreadsheet for statistical analysis using moving window strategy and the process ended again in GIS environment. This task was accomplished using one of the most recent approaches in computerized technology - component object modeling (COM). GIS tools. The computation of the dissectivity parameter outside of the GIS environment includes the “moving windows” strategy that was developed in this project. The visual basic module is capable of “dropping” a window of selected size around each point sample value and computing the statistical parameter from the samples that falls within the window size. Since LIDAR data represent a highly dense clustered data, the program computed statistics several thousands windows. The purpose of this project was to develop a module based on the component object model (COM) for the direct transfer of raw laser measurements to GIS, perform statistical analysis outside of a GIS environment and then complete surface analysis using the first and last return LIDAR value. This close coupling procedure allows the GIS analyst and LIDAR’s users to manipulate more effectively with laser data directly from the GIS environment. COM compliant software supports flexible applications using objects and components from different sources. Currently, a number of programs have COM compatibility and therefore intelligent modules to perform simple or sophisticated analysis on the spatial data can link them. All spatial data sets were managed, viewed, queried, and manipulated in an ArcGIS environment. Component Object Modules (COM) was developed and embedded in an ArcGIS environment using Visual Basic Applications (VBA) to facilitate the automation of input, processing, generation, management and representation of data. The Visual Basic.net version would also be used for the web interface programming. The resultant GIS environment enables multiple users to access and manipulate digital map and tabular data layers and files. In general, the database handles three generic types of data – 1) laser and ground transect ASCII files, 2) laser coverages (i.e., GIS data layers), and 3) satellite-based land cover vector layers and geostatistical vector layers. PROCEDURES AND RESULTS A fully integrated computerized system provides dynamic links between spatial analysis tools such as GIS and statistical software packages. The newest development in computerized technology has introduced object oriented database management systems and object component models. The GIS packages such as ArcInfo 8.0 and newer versions and IDRISI have developed their own component object system that can be linked with other COM compliant software. This integration increases the flexibility of analysis in spatial analysis. Linking GIS with other statistical and analytical packages via component object modeling can assist in the pursuit of new and creative research ideas. Remodeling ArcGIS Interface The raw ACSII LIDAR data (Figure 1) are not easily transported to analytical packages for viewing, storage and manipulation. Using the COM technology, VBA codes were written to manipulate simultaneously objects in ArcGIS and MS Office (Excel). These objects were embedded in a GIS environment using a visual basic module such as “menu” or “toolbar” items. In addition, the buttons were designed to make them operational. This technology endows the user to leverage the functionality of statistical calculations, the data management attributes in MS Excel and the spatial functions of GIS without even leaving the GIS interface. The VBA codes work in the background to couple the GIS and statistical/database software, perform the destined procedures, and then import the data back to GIS for further analysis. As can be seen in Figure 2, this interface converts the raw data file into a *.csv data format. It also sorts the data according to laser return numbers. This file can be opened in any database or simple ASCII formats for layer inputs (according to return numbers) into GIS and other tools for further geospatial analysis. Database Management Systems The integration of GIS and spatial analysis tools is commonly known as a coupling strategy. Ungerer and Goodchild (2001) described four levels of coupling: isolated, loose, close and integrated. The scope of this work falls under a close coupling strategy whereby the actual task is taken from GIS, and the spatial data is manipulated outside of GIS in an excel package using a visual basic editor. After the computation of a desired task is completed, the results are automatically imported back via established COM links into the GIS environment for further processing and analysis using 56 LIDAR COVERAGE FILES Individual pulse locations from laser instruments along the actual flight path are stored as ArcGIS point layers, along with GMT time tags. After the data raw laser data are input to GIS, they are stored as strips of point layers that can be manipulated individually or merged together in one layer. This, however, requires highly efficient computer power (CRANE) due to the extremely high data density from laser measurements. The processing of all the individual layers in GIS is done using the VBA-COM. The Terrain DTM is first developed taking the lowest return after sufficient filtering of noise (Figure 5). After the digital terrain model (DTM) was generated from the last laser return value using TIN data structure in GIS, the canopy layers were also generated and superimposed over the DTM (Figure 6). Figure 1. The raw multi-return laser data. Figure 2. The new ArcGIS Interface. Figure 4. Lidar Data Import in GIS. Figure 5. Digital elevation models generated using one strip of laser data (the lowest return) and TIN method in GIS. Figure 3. Output of Raw Data Conversion. 57 Spatial Analysis The output of this program consists of parameters, attribute values of the interest that were calculated and tabulated by the program. New fields were appended to the database, which in turn automatically updates the attribute tables in the GIS layers. The moving window technique and the computation of basic geo-spatial parameters and surface properties are being programmed in the VBA-COM as another object. The same input file can be used to calculate spatial parameters using 2D and 3D moving windows. The 2D window deals with surface parameters like dissectivity (Siska and Hung, 2003) while the 3D window permits computation of volume based parameters. Therefore, the 3D window data could be linked with the regression equations to calculate the volume, biomass and carbon content in the forest. The program is designed to capture surface characteristics that will yield to better understanding of the map content. As it was indicated earlier COM system play increasingly important role in current technology. This COM module will be also linked to the GIS platform and integrated with MS Office products for computations and further analysis. The program is designed for a point data file and it computes map characteristics. 2D and 3D Moving Windows The first step in surface analysis began with developing a module in VBA-COM for reading *.csv files that represented an output from previously transformed LiDAR raw data. Surface analysis of ground DTM or tops of the canopy is important for assessing the surface diversity and assessing the variability of statistical parameters, including the uncertainty of estimated values. The user defined moving windows are programmed around each data point and the previously mentioned statistical parameters are computed from each data value inside the moving window. In the next paragraph the example of computation of the surface dissectivity parameter (Di) is discussed. The formula for this surface parameter is as follows: Dissectivity (Di) = ( (z_max- z_min) ) *100 d Where, z_max = maximum elevation within window z_min = minimum elevation within window d = distance between the points The example of building flexible links between COM compliant software packages for the purpose of utilizing the advantages of each package involved and how newly created parameters can be implemented into these links is the computation of surface dissectivity parameter in moving windows. Di is a simple statistical parameter that captures the surface gradient change. Computations of a number of similar statistical parameters are to a large extent performed manually by using individual statistical packages not linked with GIS. This is laborious and extremely time consuming. In this project, the computation of this parameter is fully embedded in GIS using COM linkages. Figure 6. Superimposition of Ground and Top of Canopy DTMs. Figure 8. Ground layer import for surface analysis in 2D windows. Figure 7. Output from the VB based surface analysis program. 58 a close coupling strategy. The results of this work will assist users of LIDAR data in forestry and natural resource management sciences. The authors plan to continue developing more sophisticated COM based links that will significantly enhance the power of spatially oriented projects. For example, further coupling of this system with the Gradient program (Meyer 2001) that computes true surface gradients from irregular data, based on finite difference and the directional derivative method will be a great asset to geospatial, engineering and natural resource management application. The link will combine the strength of the method developed here with the true gradient approximation at any point of the surface that originated from irregularly spaced data sets such as LIDAR. Application in forest inventory and analysis include also developing regression equations for calculating 3D density parameters. Another example of application includes determining the height of airborne laser and spacing of flight lines for profiling laser in order to develop stable, reliable, precise estimates of forest volume and biomass at the county and state level. The intention of authors is to perform statistical sampling tests and developing algorithms for optimization of the grid distance in flight paths. This would further economize the cost of repetitive flights for forest inventory assessment. Figure 9. Multiple layer import for density analysis in 3D windows. DISCUSSION The above-described processes include the integration of COM compliant software such as Microsoft Office and ArcGIS through state-of–the-art technology known as close coupling. This ensures that the developed software program gives the user-friendly GUI, and the VBA runs the algorithms and integrates LiDAR datasets, statistical programs and MSOffice in the background without coming out of the interface known to the user (GIS). Multiple linear regression procedures will then be used to relate computer simulated, top of canopy measurements to estimate volume and biomass. Forest canopy simulation techniques would be used to develop the regression equations that will predict volume or biomass as a function of airborne laser measurements. Independently, neural networks may also be used to predict volume and biomass as a function of canopy height or density measures. Forest Inventory, which has to be regularly monitored and repetitively measured, is a major application field for this kind of technology as was presented in this project. The moving windows technique is extremely useful in determining the local variation of studied properties that might be significantly different from global estimates. This particular 2D-window style that computed the Di parameter indicates the spatial variability of forest canopy as measured by the imaging laser system. If the dissectivity in a certain window at the top of canopy were high, it would mean a sharp difference in tree heights in that zone. This finding may influence a decision making process in managing the renewable resources. The 3D-window system that will be implemented in a follow-up project will significantly improve spatial analysis of the forest inventory parameters such as biomass, tree volume, carbon estimates, etc. using regression parameters that can be programmed as a VBA module and linked with GIS. REFERENCES Arp, H., J.C. Griesbach, and J.P. Burns. 1982. Mapping in Tropical Forests: A New Approach Using the Laser APR. Photogrammetric Engineering and Remote Sensing 48(1): 91-100. ESRI. 2002. ArcGIS User’s Guide. ESRI Inc. Redlands, CA. Goodchild, M.F., R. Haining, and S. Wise. 1992. Integrating GIS and spatial data analysis: problems and possibilities. International Journal of Geographic Information Systems 6(5):407-423. Hoge, F.E., R.N. Swift, and E.B. Frederick. 1980. Water depth measurements using airborne pulsed neon laser system. Applied Optics 19(6):871-883. Hoge, F.E., R.N. Swift, and J.K. Yungel. 1982. Feasibility of airborne detection of laser-induced fluorescence emissions from green terrestrial plants. Applied Optics 22(19): 29913000. Hyyppä, J., O. Kelle, M. Lehikoinen, and M. Inkinen. 2001. A segmentation-based method to retrieve stem volume estimates from 3-D tree height models produced by laser scanners. Transactions of Geoscience and Remote Sensing 39(5):969-925. CONCLUSION Lefsky, M.A., D. Harding, W.B. Cohen, G. Parker, and H.H. Shugart. 1999. Surface Lidar remote sensing of basal area and biomass in deciduous forests of eastern Maryland. Remote Sensing of Environment 67:83-98. The primary goal of this project was to develop a module capable of linking GIS and statistical/database platforms in 59 Link, L.E., and J.G. Collins. 1981. Airborne Laser Systems Use in Terrain Mapping. Proceedings 15th International Symp. on Remote Sensing of Environment, ERIM, Ann Arbor, MI., Vol I: 95-110. Solodukhin, V.I., A.Ya. Zhukov, I.N. Mazhugin, T.K. Bokova, and V.M. Polezhai. 1977b. Vozmozhnosti lazernoi aeros”emki profilei lesa (Possibilities of laser aerial photography of forest profiles). Lesnoe Khozyaistvo No. 10: 53-58. Nelson, R.F., W. Krabill, and J. Tonelli. 1988a. Estimating forest biomass and volume using airborne laser data. Remote Sensing of Environment 24:247-287. Solodukhin, V.I., I.N. Mazhugin, A.Ya. Zhukov, V.I. Narkevich, Yu.V. Popov, A.G. Kulyasov, L.E. Marasin, and S.A. Sokolov. 1979. Lazernaya aeros”emka profilei lesa (Laser aerial profiling of forests). Lesnoe Khozyaistvo No. 10: 43-45. Nelson, R.F., R. Swift, and W. Krabill. 1988b. Using Airborne Lasers to Estimate Forest Canopy and Stand Characteristics. Journal of Forestry 34:3-38. Solodukhin, V.I., A.V. Zheludov, I.N. Mazhugin, T.K. Bokova, and K.V. Shevchenko. 1985. Lesotaksacionnaya obrabotka lazernykh profilogram (The processing of laser profilograms for forest mensuration). Lesnoe Khozyaistvo No.12: 35-37. Nelson, R.F., M.A. Valenti, A.Short, and C. Keller. 2003. A Multiple Resource Inventory of Delaware Using Airborne Laser Data. BioScience, accepted for publication. Stolyarov, D.P., and V.I. Solodukhin. 1987. O lazernoj taksacii lesa (Laser forest survey). Lesnoi Zhurnal No.5: 8-15. Solodukhin, V.I., A.G. Kulyasov, B.I. Utenkov, A.Ya. Zhukov, I.N. Mazhugin, V.P. Emel’yanov, and I.A. Korolev. 1977a. S”emka profilya krony dereva s pomoshch’ yu lazernogo dal’nomera (Drawing the crown profile of a tree with the aid of a laser). Lesnoe Khozyaistvo No. 2: 71-73. Ungerer, M.J., and M.F. Goodchild. 2002. Intergrating spatial data analysis and GIS a new implementation using Component Object Model (COM). International Journal of Geographic Information Systems 16(1):41-53. 60 Large Scale Photography Meets Rigorous Statistical Design for Monitoring Riparian Buffers and LWD RICHARD A. GROTEFENDT AND DOUGLAS J. MARTIN Abstract: Large scale photography (LSP) proved to be a cost-effective and accurate method for examining the effects of buffer zones on timber stand composition and wood recruitment to streams in Southeast Alaska. Rigorous statistical design requirements were met for the comparison of riparian stand characteristics between a large photo population of logged and unlogged units. The creation of the photo sample population (1,700 photo pairs from 52 km of streams) from 3,700 sq km of remote terrain was facilitated by a fixed-base camera system that was mounted underneath a helicopter. Large scale photography was the only medium that could fulfill this design because 3D vision was required to see details as fine as twigs on down trees. Visual classification of sample units by landform, stream direction, stand type, density, and treatment provided a stratified population from which 62 paired, unbiased samples were selected. Photo digitization on an analytical stereoplotter facilitated accurate measurements of key stand characteristics (tree density, height, and type; down tree density, length, position relative to stream, and decay class; stream length, area, and average bankfull width; and a stem map) that were evaluated by the analysis. Large scale photography provided a cost effective means to gather a large sample population and provided the medium for accurate measurement of a wide variety of ecosystem metrics. This study provided the largest known database of riparian buffer characteristics in Southeast Alaska and the photography allows for future re-measurement and monitoring. IINTRODUCTION acteristics of the forest stand. Large scale photography (LSP; >1:2,500) must be collected because standard aerial photography (scales of 1:12,000 to 1:62,000) has insufficient image detail. The most common method of LSP collection is from a fixed-wing aircraft that collects sequential, overlapping images that may be viewed in 3D and measured by using scale derived either from known ground coordinates or a global positioning system (GPS) / inertial measurement unit (IMU). Interpretation of riparian buffer variables also requires the ability to see in between the tree crowns to the forest floor. LSP collected by fixed-wing aircraft have large inter-photo distances that reduce the forest crown penetration and rugged topography and cloud conditions often prevent flights. Because of this a fixed base camera system was used to collect the riparian buffer imagery and perform a a retrospective study to determine: (1) the effects of the standard buffer treatment on change in stand density; (2) the post-logging mortality rates; (3) the relationship of stand density change to location within the buffer; (4) the windthrow effects compared to other stand mortality processes; (5) the importance of physical factors; and (6) the effect of windthrow on wood recruitment to streams. This met the statistical design requirements. The Alaska Forest Resources and Practices Act requires that 20 m wide buffer zones be retained along streams on private timberlands with anadromous salmonids in Southeast Alaska (ADNR 1990). A key function of the buffer zones is to supply large woody debris (LWD) that is important for the formation of fish habitat and influences other ecological processes that support fish production (Bisson et al. 1987). Because buffer zone rules were only implemented in the early 1990s, their effectiveness to provide LWD to streams has not been evaluated in Southeast Alaska and information from other regions including the Pacific Northwest is limited (Murphy 1995). The effects of wind on buffer zone survival and LWD recruitment are a major concern in Southeast Alaska because wind disturbance is the dominant environmental force shaping forest composition and structure in the region (Harris and Farr 1974). Remote sensing was determined to be the most cost-effective approach to evaluate buffer zone effectiveness. The standard transect survey approach (Dunham and Collotzi 1975) was simply too costly given the sample intensity and travel logistics that were needed to accurately characterize buffer stand composition from remote areas. Transect data also give different results from re-sampling in different locations due to the high variability of riparian buffer stands. Imagery was the optimal remote sensing medium for this study since 3D vision is required to interpret detailed char- STUDY AREA The study was conducted on Prince of Wales Island and Revillagigedo Island in southern Southeast Alaska (Figure 61 ## # (1,12) # # ## ## # ## # # #### ## #### ## # (1,0) (2,0) # ### # # ## ## ## # ## ## # ## ## # ## # ## # #### ## ## ## (4,0) ## ## ####### (2,3) ##### # # # (2,0) (3,0) ####### (1,0) ## # # ## ###### (1,0) ######### ########### ### ##### #### ## ## ## ## # # # ## # # ## ## # ## (0,2) KARTA ## ##### ### # ###### ## (0,5) # #### ## # ## # ### ###### ## ## (4,7) TOLSTOI (2,6) # # ### ### ## ## # # ## # (1,0) KLAWOCK (4,1) (4,0) (3,4) ### ###### ######## ###### ## # ## # ### ## ## ## ######### ###### # # ##### # ### # # # # # # ### ######## # # # # #### ## ## # ## # (1,0) ## # # ### (5,2) ######## ## ## # # ## ###### ## # (2,0) ### ###### ######## #### ## (0,3) (0,1) #### (1,0) ## # # # SMITH COVE ####### # ## ##### # ### ###### (7,0) # ## ## ### ## ## (1,3) # # ### ## ### ## # SULZER ### ## ## ############### ##### ## (4,8) ################ # ##### ### (1,0) ########## ### # (0,5) KETCHIKAN Study area ##### ## (4,8) N SEGMENTS PHOTOGRAPHED 2 0 (5,0) 2 4 6 Miles (N LOGGED, N UNLOGGED) Figure 1. Location of stream segments photographed and study units sampled by geographic area in southern Southeast Alaska. unpublished data) were used to determine the minimum sample size required to achieve an acceptable level of statistical power. Based on the range of density changes measured in these data, and the need to attain a statistical power of 80%, the study required a sample of 62 buffer units. The challenge of this study was to collect detailed stand compostion data (e.g., density, height, position, mortality agent, and decay class), as well as associated environmental variables (e.g., channel confinement, channel width, length, and aspect). The LSP of the riparian zone and stream channel was taken prior to leaf-out at a scale of 1:1,900 using a fixed-base camera system (Grotefendt, et al. 1996) to provide the necessary resolution. Dual 70 mm Rolleiflex 6006 metric cameras with 80 mm planar lenses were mounted on a 12 m long boom and carried transversely underneath a helicopter. Each large scale photo pair covered approximately 1 ha and was taken every 30 m as the helicopter flew at 3 knots along the centerline of the channel. This provided three different stereo views of the same stream location. To assist in classification of the LSP by physical factors and ensure the unlogged LSP was free of logging related disturbance a flight at 457 m collected smaller scale 1) and covered approximately 3,700 km2. Conifers dominate the temperate rain forest, which are predominantly composed of old-growth western hemlock Tsuga heterophylla and Sitka spruce Picea sitchensis in the uplands, with mountain hemlock Tsuga mertensiana, western red cedar Thuja plicata, and Alaska cedar Chamaecyparis nootkatensis on wetlands. Deciduous trees (red alder Alnus rubra and Sitka alder A. sinuata) are moderately abundant along streams. The buffer zones studied were from lands managed for timber harvest either by Native American corporations or by the Tongass National Forest. METHODS The population of buffer zone sample units was identified from reviews of timber harvest type maps, landowner information, and reconnaissance aerial photography. Only buffers that were 4 to 11 years old and that met Forest Resources and Practices Regulations (ADNR 1993) were included in the sample population. Data on buffer stand density from a pilot study that used LSP (Martin and Grotefendt, 62 aerial photos (1:5,712) with a metric camera. A radar altimeter was used to maintain consistent altitude during all flights and a global positioning system (GPS) recorded photo positions. An analytical stereoplotter with a measurement precision of 7 to 15 microns (AP190, Carto Instruments, Inc.) was used to measure and interpret riparian stand characteristics from the LSP. Counts of standing trees, stumps, and down trees within the buffer zone were used to calculate the proportional change in stand density (PCID). Decay class was determined from the presence/absence of conifer needles and terminal branches on all down trees. These data were used to identify recently fallen trees, which indicated post-logging mortality. The distance of each standing tree, stump, and down tree from the stream bank was measured from LSP local coordinates and was used to evaluate the effect of location (inner [0-10 m], and outer, [10-20 m]) on the change in stand density within the buffer. The cause of mortality (windthrow, bank erosion, and other) was detected from the LSP to compare windthrow effects to other stand mortality processes. The buffer units were classified by aspect, confinement, and stand density to detect any effect from physical factors. Initial confinement classification was done with the small scale photography (1:5,712) and final unit classification was completed with the LSP. Aspect was determined for each photo pair from GPS position, the small scale photography, LSP, and detailed maps produced from a geographic information system (GIS). The stand density category (low, medium, and high) was visually estimated from the LSP by a forester with extensive timber cruising experience. The logged and unlogged buffer units were stratified by geographic area and within those areas matched by aspect, confinement, and stand density category to reduce variability in PCID. The LSP was used to interpret the down trees’ positions relative to the stream (in stream, over stream, or away) so that the proportion of the stand that was recruited to the stream could be computed. Tree heights were measured from the LSP to determine whether the tree could potentially supply LWD. A tree needs to be at least 6 m tall to produce the minimum sized log (10 cm diameter and 2 m length) to qualify as LWD given the tree taper. lytical stereoplotter, as well as the 3D view, aided in identification. The fixed base camera system facilitated larger stereo views of the ground and down trees through the tree crowns. Three stereo pairs provided three different views of each same stream location. When the middle pair was measured, the two neighboring stereo pairs could be examined and this reduced the number of missed down and standing trees and stumps as well as over counting multiple top trees. Even though the ground was not visible by every tree, visible ground points could be used to develop a digital terrain model that was then used for tree height measurement. The use of decay class to define post-logging stand mortality was sufficient to detect a difference in the PCID for recently down trees in logged and unlogged units. The fine green branches, twigs, and bark texture and color could be seen on the LSP and enabled the determination of recent versus old windthrow. Also, reddish duff indicative of rotten logs was visible. Exposure of the LSP film during overcast skies increased the contrast and improved interpretability. The down tree details were visible even in shadowed areas when the diapositives were properly illuminated. The stream bank edge was delineated with the LSP by direct viewing or by interpolation when overhanging, tree crowns obscured the bank. Bank location was necessary for the GIS program to determine the position (i.e., distance from the stream edge) of all standing trees, down trees, and stumps in the buffer zone. This information was used to stratify the buffer zone into two sub-zones; 0 to 10 m and 10 to 20 m from the stream edge. Data analysis by subzones showed that the effects of windthrow on stand mortality depended on location within the buffer zone. The PCID was not statistically significant within the inner zone (0-10 m) but was in the outer zone (10-20 m). Although we stratified the buffer zone into two sub-zones, the LSP data could easily be formed into finer categories. The cause of tree mortality was discernable for most of the down trees. We used the presence of upturned root wads to indicate windthrow and the tipping out of trees on the stream bank to indicate bank erosion. Other forms of mortality were lumped together. Mortality as a result of gnawing by beaver was an example of stand mortality processes that were visible on the LSP. Channel confinement, aspect, and buffer density categories were reliably determined from interpretation of the LSP by an experienced stream ecologist or forester. This classification enabled us to stratify the sample units, which facilitated a more powerful analysis. For example, we found that all three strata influenced the PCID. We also found the measured stand density (trees/ha) results corroborated the density categories that were assigned during the photo stratification process. The position of down trees relative to the streams could be seen on the LSP. The analytical stereoplotter enabled us to see the images in high resolution to determine if the down trees were located in, over, or away from the stream. This RESULTS The stand density data (i.e., counts of standing trees, down trees, and stumps) were sufficient to detect a 5% difference in the PCID between logged and unlogged buffer zones with a statistical power of 81%. The quality of the color professional film (Kodak Portra) and the scale (1:1,900) of the LSP enabled discernment of details as fine as tiny, western hemlock, drooping leaders. Stumps were often similar in color to the dried branches, tree tops, and logs remaining after timber harvest. In these cases, the image enlargement provided by the zoom optics of the ana63 and analysis of the large population of LSP could occur without more fieldwork given additional funding. information was important for evaluating the effect of windthrow on wood recruitment to the stream. The fixed base method facilitated the collection of 1,700 photo pairs (scale 1:1,900) from 42 stream segments (29 km length from logged areas and 23 km length from unlogged areas) on 34 different streams. The same section of stream was usually viewable on 3 separate pairs due to photo pairs taken every 30 m. Ground objects that were obscured on one pair, could thus be viewed on a subsequent pair for interpretation and measurement. Over 15,000 trees, down logs, and stumps were measured and located with the analytical stereoplotter. This instrumentation and the camera system yielded horizontal errors ranging from 0.20% to 1.76% and vertical errors ranging from 1.16% to 2.61% which is comparable or better than field methods (Grotefendt, et al. 1996). Although the riparian buffers are highly variable, the large, unbiased sample number and size of each sample unit (0.2 ha) enabled us to detect effects that are patchy, such as windthrow. REFERENCES ADNR. 1990. Alaska forest resources and practices act. Alaska Department of Natural Resources, Division of Forestry, Juneau, AK. ADNR. 1993. Alaska forest resources and practices regulations. Alaska Department of Natural Resources, Division of Forestry, Juneau, AK. Bisson, P. A., Bilby, R. E., Bryant, M. D., Dolloff, C. A., Grette, G. B., House, R. A., Murphy, M. L., Koski, K. V., and Sedell, J. R.. 1987. Large woody debris in forested streams in the Pacific Northwest past, present, and future. Pages 143-190 in E.O. Salo and T.W. Cundy, editors. Streamside management, forestry and fishery interactions. University of Washington Press, Seattle, WA. CONCLUSION Dunham, D. K. and Collotzi, A. 1975. The transect method of stream habitat inventory: guidelines and applications. Ogden, Utah. United States Forest Service, Intermountain Region. LSP proved to be a cost effective and accurate method for examining the effects of buffer zones on timber stand composition and wood recruitment to streams. LSP facilitated detailed measurements of stand conditions (e.g, tree height, counts, and down tree lengths) as well as defining the environmental characteristics of buffer zones. Objects in shadowed areas could be seen and interpreted with extra illumination of the diapositives. The rigorous statistical design requirements were met by the LSP. The reliability of the inferences and conclusions were improved because all visible objects were reliably measured rather than subsampled. A larger sample size and increased amount of data per sample were possible with LSP for less cost than by field sampling methods. The LSP data are comparable in accuracy to field methods except for obscured objects that are missed. The fixed base method of LSP collection overcame the limitations of other methods by providing scale without the collection of ground control or direct georeferencing, operating in all types of rugged topography and non-optimal weather conditions, even rain, and providing stereo vision of the forest floor through the canopy. Future additional measurement Grotefendt, R.A., B. Wilson, N.P. Peterson, R.L. Fairbanks, D.J. Rugh, D.E. Withrow, S.A. Veress, and D.J. Martin. 1996. Fixed-base large scale aerial photography applied to individual tree dimensions, forest plot volumes, riparian buffer strips, and marine mammals. Proceedings of the Sixth Forest Service Remote Sensing Applications Conference: Remote Sensing; People in Partnership with Technology. April 29-May 3, 1996, ASPRS, Bethesda, MD. Harris, A. S. and W.A. Farr. 1974. The forest ecosystem of southeast Alaska. USDA Forest Service, Gen.Tech. Rep. PNW-25, Portland, OR. Murphy, M.L. 1995. Forestry impacts on freshwater habitat of anadromous salmonids in the Pacific Northwest and Alaska—requirements for protection and restoration. NOAA Coastal Ocean Program, Decision Analysis Series No. 7, NOAA Coastal Ocean Office, Silver Spring, MD. 64 Forest Canopy Models Derived from LIDAR and INSAR Data in a Pacific Northwest Conifer Forest HANS-ERIK ANDERSEN, ROBERT J. MCGAUGHEY, WARD W. CARSON, STEPHEN E. REUTEBUCH, BRYAN MERCER, AND JEREMY ALLAN ABSTRACT: Active remote sensing technologies, including interferometric radar (INSAR) and airborne laser scanning (LIDAR) have the potential to provide accurate information relating to three-dimensional forest canopy structure over extensive areas of the landscape. In order to assess the capabilities of these alternative systems for characterizing the forest canopy dimensions, canopy- and terrain-level elevation models derived from multi-frequency INSAR and high-density LIDAR data were compared to photogrammetric forest canopy measurements acquired within a Douglas -fir forest near Olympia, WA. Canopy and terrain surface elevations were measured on large scale photographs along two representative profiles within this forest area, and these elevations were compared to corresponding elevations extracted from canopy models generated from Xband INSAR and high-density LIDAR data. In addition, the elevations derived from INSAR and LIDAR canopy models were compared to photogrammetric canopy elevations acquired at distinct spot elevations throughout the study area. Results generally indicate that both technologies can provide valuable measurement s of gross canopy dimensions. In general, LIDAR elevation models acquired from high-density data more accurately represent the complex morphology of the canopy surface, while INSAR models provide a generalized, less-detailed characterization of canopy structure. The biases observed in the INSAR and LIDAR canopy surface models relative to the photogrammetric measurements are likely due to the different physical processes and geometric principles underlying elevation measurement with these active sensing systems. 65 66 Enhancing Precision in Assessing Forest Acreage Changes with Remotely Sensed Data GUOFAN SHAO, ANDREI KIRILENKO AND BRETT MARTIN Abstract: The acreage of forest cover constantly changes over time as a result of natural and/or human-induced changes. Remote sensing technology is an effective tool for detecting these changes over time. A commonly used remote sensing technique is the post-classification change detection. In this case, classification accuracy of any individual-date data can affect the accuracy of the change assessment. Various statistics are available for quantifying classification accuracy but they are not developed for assessing the accuracy of the area of cover types. To assure accurately detect forest cover change, it is essential to accurately quantify the area of forest cover from individual-date remote sensing data. In this study, we demonstrated how to increase the precision of forest change detection with a combined accuracy index, which was derived for assessing areal accuracy of cover classes. It was found that this new approach was effective in improving the accuracy of forest change detection whereas conventional accuracy statistics normally over-estimate the accuracy of forest change detection. We examined and explained several possible situations with actual remotely sensed data and hypothetical examples. The proposed technique has practical significance in decision making that is based on forest acreage changes. INTRODUCTION not have statistical relations with the errors in areal changes for individual land cover types. In other words, the errors in areal changes cannot be readily corrected with conventional accuracy assessment methods. Data processing and analysis involve errors, which propagate from one stage to the next and up to the end users. Because change detections are made by comparing data between two time periods, the errors associated with information on changes includes all the accumulated errors from both data sets used. On one hand, classification errors from a single time cannot explain the total errors of change detections; on the other hand, errors in change detections are higher than the errors involved in data from each time period. The overall effects of error propagations determine that the detected changes in cover class areas may not reflect the actual changes on the ground. The corrections of areal errors prior to change detection can help reduce the errors in areal changes over time. This paper will demonstrate the effectiveness of areal corrections with a combined accuracy index developed by Shao et al. (2003) for accurately quantifying changes in forest acreage over time. The study of change usually increases our understanding about the natural and human-induced processes at work in the landscape (Jensen 2000). Forest management activities generally lead to changes in forest area over time. Forest clearing for agriculture, urbanization, and other land uses results in deficits in forest area; afforestation, on the other hand, increases forest area. Reliable information on forest change over time reflects the overall forest management efforts and is particularly useful to understand wildlife populations, habitat, forest biodiversity, and forest productivity (Franklin et al. 2000). If a forest is the home of rare and endangered species, its areal change over time indicates how well the habitat is protected or managed. As can be seen, forest change information can yield many types of useful data, and therefore needs to be performed in a precise and accurate manner. Remotely sensed data are commonly used for forest cover change detection (e.g. Hayes and Sader 2001, Rogan et al. 2002, Turner et al. 2001). Both pre-classification and postclassification methods can be used to determine these changes (Franklin et al. 2000). In the latter approach, two dates of imagery are independently classified and registered. The accuracy of such procedures depends upon the accuracy of each of the independent classifications used in the analysis (Lillesand and Kiefer 1999). Congalton and Green (1999) demonstrated a matrix technique to assess errors in changes between two time periods. However, the errors in changes among land cover types do METHODS We conducted the study on forest cover change in a forested landscape on the eastern Eurasian Continent (128o E and 42o N) (Fig. 1). The study area was covered mainly with old-growth broadleaved-coniferous mixed forest 67 (Barnes et al. 1993), one of typical vegetation zones in the eastern Eurasian Continent (Nakashizuka and Iida 1995). Extensive logging in this area did not start until 1970s when state owned forestry enterprises were founded throughout forested regions in China (Shao et al. 1996). Forests were largely cut with a so-called small-area clear cutting method. The average size of a cutting area or field was about 15 ha (Shao and Zhao 1998). Following forest cutting, cleared fields were planted with ginseng, larch or pine seedlings, or left for natural regeneration. It took 5-10 years of natural regeneration for secondary forests to develop into closed canopy forest. The secondary forests were composed mainly of birch and aspen (Shao et al. 1994). It was common that the remaining forests between cutting fields were damaged by selectively cutting valuable trees during logging processes. The 14 pairs of thematic maps. The acreage of forest classes was computed for each thematic map. Changes in forest acreage between any of 14 1985 thematic maps and any of 14 1997 thematic maps were computed, resulting in 196 image pairs. Manually digitized thematic maps for the two areas, sized 7 by 12 km and 4 by 8 km, respectively, were used as reference data for accuracy assessment. We assume that a manual digitization is correct because the fragmentation of the homogenous forestland has some regular patterns and manual digitizing is more capable to trace the actual pattern than computer-aided classifications. A total of 1,300 points (pixels) were randomly selected from the two areas and used to build an error matrix for assessing classification accuracy of each thematic map (Congalton and Green 1999). Producer’s, user’s, and overall accuracy were computed with Fig. 1. The location of study site and a display of the TM image used in the study. dimension of the study area was defined by a quarter scene of Landsat Thematic Mapper (TM) imagery. Except for cutting areas and roads, there were no other major human disturbances within the study areas. TM data of path 116 and row 31 were acquired from May 12, 1985 and September 4, 1997 (Fig. 1). The 1997 data were rectified into a 30m resolution image in the UTM coordinate system by referring to 1:50,000 topographic maps. The 1985 data were rectified against the 1997 data and the RMS errors were controlled within 0.5 pixels. A composite data set was made by stacking the 1985 and 1997 image data. The bi-temporal data contain richer information about forest change than a single-temporal data (Wu and Shao 2003). The image data classifications were performed by seven student analysts. Each temporal data set was classified with supervised and unsupervised algorithms available from the computer program Erdas Imagine (http://gis.leicageosystems.com/Products/). After initial classification, spectral classes were grouped into two information classes: forest and clear cut. The classification experiment resulted in each error matrix. The relative error of area (REA) for the forest class was also computed (Shao and Wu 2003) as follows:  1 1  × 100 REA f =  − (1)  UA f PA f   ï£ where, UAf is user’s accuracy and PAf is producer’s accuracy for forest class. The area in percent for forest class from each thematic map was corrected with the following formula (Shao and Wu 2003): A f ,c = A f , m − n ff N × REA f × 100 (2) where, nff is the number of points checked by both reference and classification in the error matrix, N is the total of points (N = 1,300 in this study), Af,m is forest area in percent derived from a thematic map (referred to as “original forest 68 same between the two data sets. Both the corrected and the original data followed normal distributions (Fig. 3). area” in this paper), and Af,c is corrected forest area in percent. Changes in forest acreage between one of 14 1985 thematic maps and one of 14 1997 thematic maps were computed as follows: ∆= A1997 − A1985 × 100 A1985 DISCUSSION AND CONCLUSIONS The range or variation of forest acreage change is an indication of the uncertainty in forest change detection. Table 3 shows that the original data had four times higher uncertainty in assessments of deforestation than the corrected data. The correction filter shown in Eq. 2 proved significantly effective on increasing the certainty of forest area change assessment. In contrast, selectively using maps with higher overall accuracy was not as effective as the filtering process for reducing the uncertainty of forest acreage change assessment. This is because the area of a land cover class had no close relationships with the overall accuracy (Shao and Wu 2003). The extremely low values of forest area change are found in the lower right corner of the plot of scattered data of the two data sets (Fig. 4). They resulted from combinations of the 1997 forest area in map #13, which overestimated forest area, and the 1985 data from other maps. The algorithm successfully corrected the extremely low values (left part of the group being discussed), but over-corrected the values previously closer to average. The extreme variability of raw data (5 to 20% deforestation) was successfully reduced to 26.5 – 31%, and average estimates of forest land reduction for the sample (12.4% for raw data and 28.6% for corrected data) have come closer to the overall mean of 24.8%. It is concluded that the suggested algorithm provides reasonably good corrections for assessing forest area change. After the correction, the range and variation were significantly reduced, the mean value remained the same, and data distribution stayed normal. Yet there was a notable propensity to slightly over-correct the samples that included extreme values. This was caused mainly by sampling errors involved in building an error matrix because REA was derived under the assumption that the distribution of errors in the error matrix is representative of the types misclassification made in the entire area classified. The thematic maps used in this study, particularly the 1985 maps, have classification accuracy that is acceptable in real remote sensing applications. This does not mean that the computations of forest acreage change with these maps provide reasonably reliable estimations. If the mean of forest acreage change, which is 24.8%, is used as standard, the errors of forest acreage change derived from the original data can be between -60% ((10.0-24.8)/24.8*100 = -60%) and 61% ((40.0-24.8)/24.8*100 = 61%). It is obviously too risky to use uncorrected areas to quantify changes in area. If the 1985 maps had as low classification accuracy as the 1997 maps, the estimations of forest acreage change would be misleading in decision making that is based on forest acreage changes. The areal correction technique is especially effective to make the estimations of forest acreage change more reliable and meaningful. 3) where ⌬ is forest area change in percent and A1985 and A1997 are forest acreage in 1985 and 1997, respectively. Forest area change is computed with both original and corrected forest areas. In each case, there are 196 combinations. RESULTS Classification accuracy for forest class from the 1985 TM data is consistently higher than that from the 1997 TM data (Fig. 2). The former ranges between 93.4 – 95.9% for overall accuracy, 92.2 – 97.5% for user’s accuracy, and 94.1 – 99.6% for producer’s accuracy; the latter ranges between 75.8 – 88.2% for overall accuracy , 71.4 – 90.7% for user’s accuracy, and 70.3 – 94.0% for producer’s accuracy. The different ranges of the percentages indicate that the 1997 maps have much lower classification accuracy but higher variations in classification accuracy than the 1885 maps. As the changes are derived from the data of both years, the 1997 data limit the accuracy of change detections more than the 1985 data in this study. Similar to the variations in classification accuracy, the variations in forest acreage from the 1985 data set were much smaller than those from the 1997 data set (Fig. 2). The area of forest is between 304,442 and 360,514 ha in 1985 and declined to a range of 206,519 and 289,143 ha in 1997 depending on which thematic map was used to compute forest acreage. The ranges of forest acreage change were between 5 and 43 percent when the original forest area data were used (Table 1). There were 189 (out of 196) combinations that resulted in forest acreage change between 10 and 40%. Based on the overall accuracy of the 1997 maps, seven higher-accuracy maps were selected from the 14 maps (Table 1). When these better maps were used, the range of forest acreage change was between 13 and 40%. After area correction with Eq. 2, the ranges of forest areas were reduced to between 28,399 and 30,155 ha for the 1985 maps and between 20,853 and 22,462 ha for the 1997 data (Fig. 2). With the corrected forest areas, the changes in forest acreage were between 21 and 31 percent (Table 2). When only the better maps were used, the range of the changes dropped slightly to between 22 and 29 percent (Table 2). The major differences in forest acreage change between the original and corrected data sets were the range and variation (Table 3). The mean value of the change was about the 69 1985 1 1 2 2 3 3 4 4 5 5 6 Overall Accuracy 7 User's Accuracy 8 6 8 Producer's Accuracy 9 10 11 11 12 12 13 13 14 14 60 70 80 90 Corrected Area 9 10 50 Original Area 7 15000 100 20000 25000 30000 35000 1997 1 1 2 2 3 3 4 4 5 5 6 Overall Accuracy 7 User's Accuracy 8 Producer's Accuracy 9 6 8 10 11 11 12 12 13 13 14 14 60 70 80 90 10000 100 Corrected Area 9 10 50 Original Area 7 15000 20000 25000 30000 Fig. 2. Classification accuracy in percent (left) and area in ha (right) of forest class in 14 thematic maps derived from the 1985 TM data (above) and 1997 TM data (below). Figure 3. Distribution of the raw (a) and corrected (b) estimates of forest area change. 70 Table 1: Changes in forest area by percent among 14 maps between 1985 and 1997 using original forest area values. Map# 1 2 3 4 5 6 7 8 9 10 11 12 13 14 1 30.5 32.4 34.8 24.0 19.0 17.6 26.9 34.7 38.1 20.5 29.8 18.1 13.4 19.1 2 29.6 31.5 34.0 23.0 17.9 16.5 26.0 33.9 37.3 19.4 28.9 17.0 12.3 18.0 3 26.6 28.5 31.1 19.7 14.4 12.9 22.8 31.0 34.6 15.9 25.8 13.4 8.5 14.5 4 31.6 33.4 35.8 25.2 20.2 18.9 28.0 35.7 39.1 21.7 30.9 19.3 14.7 20.3 5 32.2 34.0 36.3 25.8 20.9 19.5 28.6 36.3 39.6 22.3 31.5 20.0 15.4 21.0 6 27.6 29.5 32.0 20.8 15.5 14.1 23.8 31.9 35.5 17.1 26.8 14.6 9.7 15.6 7 23.8 25.8 28.5 16.7 11.1 9.6 19.9 28.4 32.2 12.8 23.0 10.2 5.0 11.2 8 29.9 31.7 34.2 23.3 18.2 16.8 26.2 34.1 37.6 19.7 29.1 17.3 12.6 18.3 9 32.8 34.5 36.9 26.4 21.6 20.3 29.3 36.8 40.1 23.0 32.1 20.7 16.2 21.7 10 35.7 37.4 39.6 29.6 25.0 23.7 32.3 39.6 42.7 26.3 35.0 24.2 19.8 25.0 11 30.4 32.3 34.7 23.9 18.9 17.5 26.8 34.6 38.1 20.4 29.7 18.0 13.3 18.9 12 31.6 33.4 35.8 25.2 20.2 18.9 28.0 35.7 39.1 21.7 30.9 19.3 14.7 20.3 13 29.6 31.5 34.0 23.0 17.9 16.6 26.0 33.9 37.4 19.5 28.9 17.1 12.3 18.0 14 24.7 26.7 29.3 17.6 12.1 10.6 20.8 29.2 32.9 13.7 23.9 11.2 6.1 12.2 Note: Bold numbers indicate higher-accuracy maps in 1997. 71 Table 2: Changes in forest area by percent among 14 maps between 1985 and 1997 using corrected forest area values. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 1 26.4 25.0 26.8 26.4 23.6 23.5 24.6 25.2 27.7 23.4 25.6 24.3 28.9 25.0 2 26.0 24.6 26.4 26.0 23.2 23.1 24.2 24.8 27.3 23.0 25.2 23.9 28.5 24.6 3 24.7 23.3 25.1 24.8 21.9 21.7 22.9 23.5 26.1 21.7 24.0 22.6 27.3 23.4 4 26.6 25.2 27.0 26.7 23.9 23.7 24.9 25.5 27.9 23.7 25.9 24.6 29.1 25.3 5 27.4 26.0 27.7 27.4 24.7 24.5 25.6 26.2 28.6 24.4 26.6 25.3 29.8 26.0 6 24.8 23.4 25.2 24.8 22.0 21.8 23.0 23.6 26.1 21.7 24.0 22.6 27.3 23.4 7 24.0 22.6 24.4 24.0 21.2 21.0 22.2 22.8 25.3 20.9 23.2 21.8 26.6 22.6 8 26.2 24.8 26.6 26.2 23.5 23.3 24.4 25.0 27.5 23.2 25.4 24.1 28.7 24.9 9 27.2 25.8 27.6 27.2 24.5 24.3 25.4 26.0 28.5 24.2 26.4 25.1 29.6 25.9 10 28.4 27.1 28.8 28.5 25.8 25.6 26.7 27.3 29.7 25.5 27.7 26.4 30.9 27.1 11 27.0 25.6 27.4 27.0 24.3 24.1 25.3 25.8 28.3 24.0 26.3 24.9 29.5 25.7 12 26.6 25.2 27.0 26.7 23.9 23.7 24.9 25.5 27.9 23.7 25.9 24.6 29.1 25.3 13 26.5 25.1 26.9 26.6 23.8 23.6 24.8 25.3 27.8 23.6 25.8 24.4 29.0 25.2 14 24.1 22.7 24.5 24.2 21.3 21.1 22.3 22.9 25.5 21.0 23.4 21.9 26.7 22.7 Map# Note: Bold numbers indicate higher-accuracy maps in 1997. Table 3. A comparison of forest area change between the original data and the corrected data. Minimum Maximum Mean Range Standard Deviation Variance Skewness Standard Error of Skewness Kurtosis Standard Error of Kurtosis Original Data 5.03 42.72 24.80 37.69 8.36 69.82 -.046 .174 -.935 .346 72 Corrected Data 20.90 30.85 25.21 9.95 2.02 4.08 .159 .174 -.279 .346 Corrected data (%) vegetation regrowth in a time series. Photogrammetric Engineering and Remote Sensing 67: 1067-1075. 45 Jensen, J.R. 2000. Remote Sensing of the Environment: An Earth Resource Perspective. Prentice Hall, Upper Saddle River, NJ. 544 p. 35 Lillesand, T.W. and R.W. Kiefer. 1999. Remote Sensing and Image Interpretation (4th Edition). John Wiley & Sons, New York. 736 p. 25 Nakashizuka, T. and S. Iida. 1995. Composition, dynamics and disturbance regime of temperate deciduous forests in Monsoon Asia. Vegetatio 121: 23-30. 15 Rogan, J., J. Franklin and D.A. Roberts. 2002. A comparison of methods for monitoring multitemporal vegetation change using Thematic Mapper imagery. Remote Sensing of Environment 80: 143-156. 5 5 15 25 35 45 Raw data (%) Shao, G., P. Schall, and J.F. Weishampel. 1994. Dynamic simulations of mixed broadleaved-Pinus koraiensis forests in the Changbaishan Biosphere Reserve of China. For. Ecol. Manage. 70: 169-181. Figure 4. Forest area change, %: raw data vs. corrected data. Notice the 13 lower right corner points—all of those were generated using 1997 forest areas estimated by one of the experts. Shao, G. and G. Zhao. 1998. Protecting versus harvesting of oldgrowth forests on the Changbai Mountain (China and North Korea): A remote sensing application. Natural Areas Journal 18: 334-341. REFERENCE: Barnes, B.V., Z. Xu and S. Zhao. 1993. Forest ecosystems in an old-growth pine-mixed hardwood forest of the Changbai Shan Preserve in northeastern China. Canadian Journal of Forest Research 22: 144-160. Shao, G., S. Zhao, and H.H. Shugart. 1996. Forest Dynamics Modeling: Preliminary Explanations of Optimizing Management of Korean Pine Forests. China Forestry Publishing House, Beijing (in Chinese). 159 p. Congalton, R.G. and K. Green. 1999. Assessing the Accuracy of Remotely Sensed Data: Principles and Practices. Lewis Publishers, New York. 137 p. Shao, G., W. Wu, G. Wu, X. Zhou, and J. Wu. 2003. An explicit index for assessing the accuracy of cover class areas. Photogrammetric Engineering & Remote Sensing 69: 907-913. Franklin, S.E., E.E. Dickson, D.R. Farr, M.J. Hansen, and L.M. Moskal. 2000. Quantification of landscape change from satellite remote sensing. Forestry Chronicle 76: 877-886. Turner, B.L., S.C. Villar, D. Foster, J. Geoghegan, E. Keys, P. Klepeis, D. Lawrence, P.M. Mendoza, S. Manson, Y. Ogneva-Himmelberger, A.B. Plotkin, D.P. Salicrup, R.R. Chowdhury, B. Savitsky, L. Schneider, B. Schmook, C. Vance. 2001. Forest Ecology and Management 154: 353370. Hayes, D.J. and S.A. Sader. 2001. Comparison of change-detection techniques for monitoring tropical forest clearing and 73 74 Automatic Extraction of Trees from Height Data Using Scale Space and SNAKES BERND-M. STRAUB Abstract: An approach is presented for the automatic extraction of trees and the boundaries of treecrowns. It is based on a multi-scale representation of an orthoimage and a surface model in Linear Scale Space. The segmentation of the surface model is performed using a watershed transformation. Finally the boundary of every crown is measured with Snakes (Active Contour Models). The approach was tested with data from laser scanner (1 m) and image matching (0.25 m). INTRODUCTION strategy. In the last section some exemplary results are shown. The paper closes with a short summary and an outlook. In this paper we present a new approach for the automatic extraction of individual trees using a true orthoimage and a surface model as input data. The surface model is used as main source of information for the extraction of individual trees. Additional colour information from the orthoimage is used to differentiate between vegetation and other objects in the scene. The aim of the presented approach is to detect every tree in the observed area of the real world and to measure the boundary of its crown. Originally, the method was developed for the automatic extraction of trees in settlement areas using height data from image matching. The type of surface model used has a ground sampling of 0.25 m, it was produced by the French company ISTAR using 1:5000 color infrared aerial images. An example of such a data set, acquired over Grangemouth, Scotland in summer 2000 is depicted in Figure 1, refer (Straub and Heipke 2001) for details. In order to demonstrate the potential of the approach, we applied it on a test site in a forest in the Austrian alps. The main species in this test site is spruce (94%). A surface model was used in the investigations. The laser scanner flight with a Toposys I Scanner was carried out in August 1999 in Austria, close to Hohentauern. The flying height was approximatley 800 m above ground, leading to 4-5 points per square meter, refer (Baltsavias 1999) for an overview about airborne laser scanner. The data were provided by Joanneum Research in Graz, Austria for this investigation. In the next section of the paper a short overview is given on the related work in the field of automatic extraction. In the main section of the paper the approach is described in detail, it is divided in two subsections: The first one depicts the object model for trees and the second one the processing RELATED WORK The first trial to utilize an aerial image for forest purposes was performed in 1897 (Hildebrandt 1987). Since that time the scientific forest community has worked on methods for the extraction of tree parameters from aerial images. Early work was carried out on the manual interpretation of images for forest inventory (Schneider 1974), (Lillesand and Kiefer 1994). Pioneering work in the field of the automated, individual tree extraction from images emerged about one and a half decades ago (Haenel and Eckstein 1986), (Gougeon and Moore 1988), (Pinz 1989). Recent work in the field was published by Pollock (1996), Brandtberg and Walter (1998), Larsen (1999), Andersen et al. (2002), Persson et al. (2002), Schardt et al. (2002). An excellent state-of-the-art overview is given by Hill and Leckie (1999). Some of the recent publications are described in detail in the following section. A common element of most approaches is the geometric model of a tree as it was proposed by Pollock (1994). In the following, this surface description is referred to as the Pollock-Model. The geometric part of the Pollock-Model can be depicted as follows: n 2 2 2 zn ( x + y ) + =1 an bn (1) The parameter a corresponds to the height, and b to the radius of the crown, n is a shape parameter. Two examples of surfaces which can be described with Equation 1 are de75 Figure 1: Orthoimage, surface model and 3D vizualization of automatically extracted trees in settlement areas. DESCRIPTION OF THE APPROACH picted in Figure 2. The surface of a real tree is of course very noisy in comparison to the Pollock-Model. This “noise” is not caused by the measurement of the surface, it is simply a consequence of using such a model for a complex shape like the real crown of a tree. But the main shape of the crown is well modelled with this surface description. Another common element in the most approaches is the application of the Linear Scale-Space in the early processing stages (refer to (Dralle and Rudemo 1996), (Brandtberg and Walter 1998), (Schardt et al. 2002), and (Persson et al. 2002)). In (Andersen et al. 2001) a Morphological ScaleSpace is used for the extraction of tree positions. A basic idea of the Linear Scale-Space is to construct a multi-scale representation of an image, which only depends on one parameter and has the property of causality: that means it has to be insured, that features in coarse scale have always a reason in fine scale (Koenderink 1984). One can show, that a multi-scale representation based on a Gaussian function used as a low pass filter fulfils this requirement. In practice, r the original signal f ( x ) is convolved with a Gaussian kernel with a different scale parameter s; the result of the convolution operation is assigned as . Small values of σ correspond to a fine scale level, large values to a coarse scale. An extensive investigation and mathematical reasoning including technical instructions can be found in (Lindeberg 1994). One of the crucial problems is the estimation of the scale parameter σ , i.e., the selection of the scale level for the extraction of the low-level features. In (Schardt et al. 2002) it was proposed to use a scale selection mechanism, refer to (Lindeberg 1998) for details, based on the maximum response after Scale-Space transformation. In our approach the scale selection is applied on a higher level, i.e. after the segmentation of the image, and not before, as it was proposed in (Schardt et al. 2002). This allows an internal evaluation of the segments on a semantic level, which is an important possibility if it is necessary to distinguish between trees and other objects. The idea of our approach is to create a multi-scale representation of the surface model similar to (Persson et al. 2002). The selection of the scale level is of crucial importance for the extraction of trees, the reasons are: (1) The correct scale level depends mainly on the size of the objects one is looking for. In the case of trees this size can neither be assumed to be known nor is it constant for all trees in one image. The size of trees depends on the age, the habitat, the species and many more parameters, which cannot be modelled in advance. (2) The correct scale is of crucial importance for the segmentation. The small structures of the crown are very difficult to model and – except for small structures - the crown has a relatively elementary shape. The image is segmented in a wide range of scales, bounded by reasonable values for the minimum and maximum diameter of a tree’s crown. In (Gong et al. 2002) the typical range for the diameter is proposed to be minimal 2.5 m up to 15 m covering all species of trees. This section is subdivided into two parts. In the first part the model for trees, which constitutes the basis for the strat- Figure 2: 3D visualisation of the Pollock-Model. Left: Surface model of a typical deciduous tree: a=7, b=3.5, n=1.2. Right: Coniferous tree: a=20.0; b=5.0; n=1.2, different scales of the horizontal and vertical axis. 76 In the case of real data this model is only valid in a convenient scale level. A height profile from real data is used to explain the term “convenient” in this context. Two different Scale-Space representations of the surface model with σ values of 0.5 m and 8 m are depicted in Figure 4. One can see that more and more fine structures disappear and the coarse structure is enhanced with the increase of the scale parameter σ. The height profile along the tree tops is measured along the dotted line which is superimposed to the surface model in Figure 4. The left height profile measured in the original surface model is noisy compared to the profile of the synthetic trees. As a result of this noise the Laplacian is oscillating close to zero. In the “correct” scale level for this small group of trees the assumptions regarding the Laplacian are fulfilled quite well. The Laplacian is negative for trees and positive for the valleys between them, just like the profile of the synthetic Pollock-Trees (Figure 3). The coarse structure of the crown is enhanced, and as a result of this the properties of the Pollock-Model are valid also for the surface model of real trees in this scale level. egy of extraction, is described. The strategy is explained in detail in the second part of the paper. MODEL FOR TREES The geometric part of the model of an individual tree simplifies the crown to a 2.5D surface, see the Pollock-Model (Equation 1). The parameter n can be used to define the shape of a broad-leafed tree with a typical range of values from 1.0 to 1.8, and also for conifers with a typical range for n from 1.5 to 2.5. These numerical values are based on an investigation described in (Gong et al. 2002). Based on the Pollock-Model the following features for the extraction from the surface model can be derived: the projection of the model into the xy-plane is a circle with a diameter in given range. Furthermore, the 3D shape of the surface is always convex. The image processing is based on differential geometric properties. A profile is used along four tree tops to study the geometrical properties of the surface if the trees stand close together. In the left part of Figure 3 four Pollock-Trees computed with a=6 m, b=2 m, and n=2.0 (1 m is equivalent to 10 pixels respectively grey values) are depicted. The profile is plotted in dark grey in Figure 3. PROCESSING STRATEGY In general, there are two possibilities to build a strategy for the automatic extraction of trees from the image data. The first possibility is to model the crown in detail: one could try to detect and group the fine structures in order to reconstruct the individual crowns. The second possibility is to remove the fine structures from the data with the aim to create a surface which has the character of the PollockModel. In the literature examples for both strategies can be found: Brandtberg (1999) proposed to use the typical fine structure of deciduous trees in optical images for the detection of individual trees. In (Andersen et al. 2002) the fine structure of the crown is modelled as a stochastic process with the aim to detect the underlying coarse structure of the crown. The other strategy, the removal of noise, was proposed by Schardt et al. (2002) and by Persson et al. (2002). The main problem of this type of approach is the determination of an optimal low pass filter for every single tree in the image. This is kind of a chicken-and-egg problem, because the optimal low pass filter depends mainly on the diameter of the individual tree one is looking for, which is not known in advance. The basis of our approach is the Linear Scale-Space Theory. The watershed transformation is used as a segmentation technique, Fuzzy Sets for the evaluation of the segments, and Snakes for the refinement of the crowns outline. The basic ideas of the Linear Scale-Space Theory were originally proposed by Koenderink (1984), and were worked out by Lindeberg (1994). The watershed transformation for the segmentation of images was introduced by Beucher and Lantéjoul (1979). Details about the watershed transformation can be found in (Soille 1999). Fuzzy Sets (Zadeh 1965) are used, because they are a “very natural and intuitively Figure 3: Profile of the surface model of four PollockTrees, the location of the profile is depicted in the upper left corner. One can see that the “valley” between the trees decreases from the left to the right. The absolute value of the gradient (black line in Figure 3) decreases also. Obviously this is a consequence of the decreasing distance between the trees, and of the crown shapes. The surface at the tree tops has a convex shape in both directions, along and across the profile. Therefore, the sum of the second partial derivations is always negative for the whole crown (refer to the light grey line in Figure 3). At a point on the profile between two trees the partial, second derivative is smaller than zero along the profile and larger than zero perpendicular to the profile. Therefore, the Laplacian of the surface model at points like this is normally higher than at points on the crown, because both partial second derivatives are smaller than zero at the tree tops. These characteristics lead to local maxima between the crowns in . 77 r Figure 4: Representation of the surface model H ( x ) at two different scale levels, left: σ=0.5 m, right: σ=8 m. The height profiles below are measured along the dotted lines in the images. r Bσ (a ) with a feature vector ar of four fuzzy membership values. (3) Selection of valid hypothesis: Every tree hypothesis r Bσ (a ) is first evaluated based on the feature vector. In some cases this is leading to valid hypothesis from different scale levels which are covering each other in the xy-plane. These covering segments have to be detected and the best one acr cording its membership value, is selected as Treeσ (a ) . r (4) The outline of the crown of every selected Treeσ (a ) is measured using Snakes. plausible way to formulate and solve various problems in pattern recognition.” (Bezdek 1992). Snakes were introduced by Kass et al. (1988) as a mid-level tool for the extraction of image features. They “look on nearby edges, localizing them accurately” (Kass et al. 1988). These tools were combined into a strategy, whose main steps are depicted in Figure 5. The aim is to detect individual trees first and reconstruct the outline of the crown in a second step. As mentioned above a multi-scale representation of the image in the Linear Scale-Space is used as a basis for the approach: Segmentation of the Surface Model of the sur(1) Segmentation: Every scale level face model is subdivided into segments using the watershed transformation. The resulting segments are the Basins of the watershed transformation, where indicates the scale level. (2) Computation of membership values: Membership values were assigned to every segment , which are partly derived from both segments (size and circularity), or the r area belonging to the appropriate scale level H ( x , σ ) of the r surface model (curvature), and the image I ( x , σ ) ( ⇔ vegetation index or texture). This results in hypothesis for trees The segmentation of the surface model is the part of the approach which depends heavily on the scale. As mentioned before the segmentation of the surface model is performed in many scales. The segmentation procedure itself should be (1) free of parameters and (2) operate only in the image space, not in the feature space. The reason is that a feature space has to be independent from the scale level. The watershed transformation fulfils these requirements. Additionally it is well suited for the segmentation of height data because the key idea of the watershed transformation is a segmentation Figure 5: Processing strategy for the extraction of trees. 78 of an image by means of a flooding simulation (Soille 1999). Basins are the domains of the image, which are filled up first if a water level increases from the lowest grey value in the image, Watersheds are embankments between the basins. This segmentation technique is also used in (Schardt et al. 2002) and a quite similar technique in (Persson et al. 2002) with the aim of detecting individual trees. If the watershed procedure is applied to extract trees from height data, the surface model has to be transformed in such a way, that the trees itself are basins. The easiest way to do this is to invert the surface model, as proposed in (Schardt et al. 2002). In forest areas there are usually narrow valleys between the individual crowns. In other areas the situation may change, for example if trees occur in small groups (e.g. in settlement areas), or if a way or a road occurs in forest areas. If these valleys are wide, the outlines of the basins are quite poor approximations of the crowns. In the general case it leads to better results to use the first or the second order derivatives of the surface model as the segmentation function, in closed stands as well as in open areas with individual trees. Good results have been found using the Laplacian as the segmentation funcsquared tion in our experiments. The following break points are used to define the membership function (Figure 6, upper right) for the size of a tree: The lower border is 20 m² for a minimum diameter of 2.5 m and the upper border is 700 m² (maximum 15 m diameter). For larger values the membership value decreases, the largest possible diameter is assumed to be 35 m (3850 m²). These typical values for diameters cover all tree species, they can be found in (Gong et al. 2002). The feature vitality is derived from an optical image, used to discriminate between vegetation and non-vegetation areas. In the settlement example the Normalized Difference Vegetation Index (NDVI) is used for the vitality (Figure 6 lower left) of a segment. A membership function with increasing membership values for positive NDVI values is used with a break point at (0.5, 0.8). This breakpoint is set empirically, motivated by the fact that the NDVI values measured at healthy trees are usually higher than the values for other vegetation types as bushes or lawn. Selection of Valid Hypothesis The classification of the segments is subdivided into two steps. First, valid segments are selected according to their membership values. A tree is an object with a defined size, circularity, convexity and vitality. Consequently the minimum value of the feature vector is the value which defines if a hypothesis is a or not. In some cases a valid hypothesis can occur at a more or less identical spatial position in the scene, but at different scale levels. Some examples can be found in Figure 7, the left image shows the valid hypothesis for trees at a scale level of = 2 pixels (according to 0.5 m), the middle at = 4 pixel, and the right at = 8 pixel in the foreground. All Basins of the watershed transformation are depicted in the background superimposed on the surface model in the corresponding scale level. One can see, that valid tree hypothesis occur in more than one scale. In some cases the segments are quite similar in both depicted scale levels, and in some other cases the segments are subdivided in the finer scale level. The trivial case – a segment in just one scale – is rather an exception. These different situations of every segment have to be analyzed. Hence, the type of the topological relation between the segments of different scale levels has to be classified. If the type is known, the best hypothesis for a tree can be selected for a given spatial position. The classification of the topological relations between the valid segments is performed as proposed by Winter (2000). In general, eight different topological relations exist in 2D space: disjoint, touch, overlap, equal, covers, contains, contained by, and covered by (Egenhofer and Herring 1991). These topological relations can be subdivided into two clusters C1 and C2, where the C1 cluster includes the relations disjoint, touch and C2 includes the other types. The overlap relation is between these two clusters, it can be divided into weak-overlap (C1) and strong-overlap (C2) (Winter 2000). The motive behind this partitioning is that the relations in C1 are similar to disjoint, and in C2 to equal. Computation of Membership Values Four membership functions are used to transform the values of circularity, convexity, size and vitality into membership values. Circularity of a segment is expressed as (Figure 6 upper left), where Area is the aera covered by the segment and is the maximum distance from the center of the region to the border. A sensible lower border is close to the value of 0.7 (circularity of square) and a the upper border is 1 (circularity of a circle, the largest possible value). The sign of the Laplacian of the surface model is used to discriminate between convex surfaces as trees and non-convex surfaces. For example, the surfaces of buildings and most ground surfaces are planes, whereas the crown of a tree is a convex surface. Thus, a negative mean value of the Laplacian within the covered area of a segment leads to a membership value of 1, and in the case of a positive mean value, the membership value is 0 (Figure 6, lower right). Figure 6: Membership functions, upper left: size, upper right: circularity, lower left: convexity, lower right: vitality. 79 Figure 7: Upper row: Basins of the watershed transformation on three scale levels. Lower row: Hypotheses for trees in the corresponding scale levels. Left: σ = 2 pixels, middle: σ = 4 pix, right: σ = 8 pixels. r We postulate that all the segments BA (a ) which have a r topological relation in C2 to another segment BB (a ) , A ≠ B from another scale level are potential hypothesis of the same tree in the real world. The best hypothesis - the one with the r highest membership value - is selected as a Treeσ (a ) instance. Accordingly, both investigated hypothesis are assumed to be valid, if the relation between the two segments is part of C1. The final selected hypotheses are depicted in Figure 9. But even if the trees in the scene were detected correctly, the boundaries are often poor approximations for the outline of the individual crowns. This problem leads to the last processing step, where the outlines of the crowns will be refined with Snakes. or an image, or the edges of an image. Snakes were originally introduced by Kass et al. (1988) as mid-level algorithms which combine geometric and/or topologic constraints with the extraction of low-level features from images. The principal idea is to define a contour with the help of mechanic properties like elasticity and rigidity, then to initialize this contour close to the boundary of the object one is looking for, and finally to let the contour move in the direction of the boundary of the object. In general, there are two main drawbacks to the application of Snakes as a measurement tool. The first one is that the Snake has to be initialized very close to the features one is looking for. The second one is the challenging tuning of the parameters, primarily the weighting between internal and external forces and the selection of the external force field itself. In our approach the Snake is used only for fine measurement in the last stage, the coarse shape of the crown is more or less known. Furthermore, the approximation is often too small. Based on these constraints one can built a Snake which is quite stable under these special conditions: the geometry r of the Snake is initialized for every Treeσ (a ) as circular shaped closed polygon at the gravity center of the appropriate Basin Bσ . The snake could also be initialized with the outline obtained from the watershed transformation. The idea of using a circle instead of that is to make the Snake Measurement of the Crown’s Outline Up to now the geometry of the segments stems from different scale levels, as the Basins Bσ were extracted in different scales. But the outline of the crown is an object without a changing scale, as distinct from the crown itself. The outline of the crown is measured in the fine scale with the help of Snakes. A Snake is a kind of a virtual rubber cord which can be used to detect valleys in a hilly landscape with the help of gravity. This landscape may be a surface model, 80 were extracted manually. This is looked upon as the reference data set for this evaluation (left of Figure 9). It should be noted, that these manually extracted data are a kind of an optimal result of what the approach should deliver from the developer’s point of view. The relationship between the manually extracted reference trees and the trees in the real world is not discussed here. An automatically extracted tree is assigned as a True Positive (TP), if it has a topological relation from the C2 cluster with a tree in the manually extracted reference data set, otherwise it is assigned as a False Positive (FP). Those trees in the reference data set with a C1 relation to an automatically extracted tree are assigned as False Negatives (FP). Based on these numbers, the Completeness and the Correctness2 of the extraction result can be computed: Figure 8: Example for the measurement of a crown outline with a Snake, five different optimization steps are depicted. Completeness = TP optimization a bit more independent from the geometry of the segment stemming from the watershed transformation. The radius of the circular closed polygon is computed r by Area ( Bσ ( a ) ) / π . This initialization stage is depicted in Figure 8 as the black circle in the right image. The parameters for the internal energies were tuned in the following way: the length of the contour is weighted low, and the curvature is weighted high. Without external forces, a Snake which is tuned in such a way converges to a circle with a trend to decrease its length1. As the approximation is often too small, an additional force is added which makes the Snake behave like a balloon, which is inflated (Cohen 1991). With this additional force the contour moves towards the outline of the crown if no external forces influence the movement. The sum of the gradients over all scale levels is used as external force. r Finally, the membership values of every Treeσ (a ) has to be computed again because the outlines have changed. Also the topological relations between all tree hypothesis are no longer valid and have to be computed again. A changing of r the topology occurs, if two or more segments Bσ (a ) are parts of the same crown in the real world. In these cases, the Snake usually converges to the correct solution, i.e. the topological relation changes from the C1 cluster (similar to disjoint) to the C2 cluster (similar to equal). As these updated membership values are quite independent from the pre-processing in the different scale levels, these values are used as an internal evaluation of the tree hypothesis. TP + FN Correctness = TP 2 TP + FP In order to characterize the accuracy of the correct automatically extracted trees, the mean value and the standard deviation of the mean value were computed for the distance between the centers of gravity and the radii between reference tree and automatically extracted tree. The results are depicted in Table 1. One can see that the internal evaluation, which is performed after the measurement of the crown’s outline, leads to a significant degradation for the completeness. As expected the Correctness is enhanced, 97% of the extracted trees are correct. The accuracy measures are nearly equivalent. This is a little bit surprising, because it was expected that the outline of the crown would be much more precisely delineated by the Snake than by the watershed transformation. Similar experiences were made with other datasets. SUMMARY AND OUTLOOK In this paper an approach for the automatic extraction of trees is presented. The object model and the processing strategy are illustrated in detail, as well as some exemplary results. The approach is free of assumptions about the scale level, because the segmentation is performed in a wide range of different scale levels. The classification of the tree hypothesis is based only four parameters: size, circularity, convexity, and vitality. Of these four parameters only the vitality is dependent on using image data, the others are geometric object properties. It should be noted that the values for the size of the crowns stems from an independent investigation (Gong et al. 2002), and the convexity is always positive. Only the breakpoints in the circularity membership function are empirical values. The measurement of the crown’s outline is performed with a Snake algorithm. The adjustment of the parameters for the Snake is a quite difficult task. But once adjusted, the algorithm is stable as a measurement tool for this task with- RESULTS The described approach was applied to a small subset of the Hohentauern dataset as mentioned in the introduction of this paper. The selection of the subset was mainly motivated by the fact that ground truth is available for a part of this scene. The LIDAR first return data were transformed into a 0.25 m raster. In order to get an initial idea of the performance of this approach in forest areas the trees in a slightly smoothed (σ =1.0 pixel) version the surface model 81 Figure 9: Left: Manually extracted trees superimposed to the surface model. Middle: Selected tree hypotheses, different gray values correspond to different scale levels. Right: Results of the approach, final selected trees after internal evaluation. est or dense settlement areas. This is possible, because the approach is free of assumptions about the terrain and the height of the trees is not used for the detection. out changing these settings for different scenes. Unfortunately, the accuracy of the results, namely the position and the radius of trees, did not increase. This should be investigated in detail, to deterime if this applies only for the center of gravity and the radius or for the whole outline. The approach was tested with synthetic data (refer Figure 3), high resolution data in settlement areas (Completeness 68%, Correctness 82%), and a small dataset of a forest (Completeness 70%, Correctness 86%). In the forest case it is necessary to evaluate the results with more reliable reference data. Furthermore, it is planned to use the information about the outline of the individual trees for a detailed classification. For example, we will investigate how the curvature is correlated with the shape parameter of the Pollock-Model, which can be used as a feature for the classification of the species. Another idea is the use of the approach in a combined strategy for the extraction of the ground surface in for- ACKNOWLEDGEMENT Parts of this work were funded by the European Commission under the contract IST-1999-10510. The surface model and the true othoimages were produced by the French company ISTAR. All aerial images, digital elevation models, and true orthoimages are copyrighted by ISTAR, Sophia Antipolis, France. Many thanks go to Alix Marc and Frank Bignone for their valuable cooperation. The Hohentauern dataset was provided by Joanneum Research in Graz, Austria. Many thanks go to Matthias Schardt and Roland Wack for their trustful and open discussion in Graz, April 2003. Table 1: Quality measures and accuracy approximations (in pixels) for the Hohentauern dataset (1 pixel = 0.25 m). The numbers are given after the step “Selection of the valid hypotheses” in the first row, and after “Measurement of the crowns outline” in the second row in the table. Step Selection of valid hypothesis Measurement of the crowns outline Quality Completeness Correctness 70% 86% 45% 97% 82 Position Mean Difference RMS 4.4 3.6 4.6 3.5 Radius Mean Difference RMS -3.0 3.6 -2.9 3.6 REFERENCES Geoscience and Remote Sensing Symposium, 2. IEEE. Haenel, S., and Eckstein, W. 1986. Ein Arbeitsplatz zur automatischen Luftbildanalyse. P. 38-42, Springer, Berlin, Deutschland. Hildebrandt, G., 1987. 100 Jahre forstliche Luftbildaufnahme Zwei Dokumente aus den Anfängen der forstlichen Luftbildinterpretation. Bildmessung und Luftbildwesen, (1987)55:221-224. Andersen, H., Reutebuch, S.E., and Schreuder, G.F. 2001. Automated Individual Tree Measurement through Morphological Analysis of a LIDAR-based Canopy Surface Model. P. 11-22, in Proceedings of the First International Precision Forestry Cooperative Symposium, Seattle, USA, June 1720. Andersen, H., Reutebuch, S.E., and Schreuder, G.F. 2002. Bayesian Object Recognition for the Analysis of Complex Forest Scenes in Airborne Laser Scanner Data. P. 35-41, in The International Archives of the Photogrammetry, Remote Sensing and Spatial Information Sciences, XXXIV. 3A. ISPRS, Graz, Austria. Hill, D.A., and Leckie, D.G. (eds.), 1999. International forum: automated interpretation of high spatial resolution digital imagery for forestry, February 10-12, 1998. Natural Resources Canada, Canadian Forest Service, Pacific Forestry Centre, Victoria, British Columbia, 395p. Baltsavias, E., 1999. Airborne laser scanning: existing systems and firms and other resources. ISPRS Journal of Photogrammetry and Remote Sensing, (1999)54:164-198. Kass, M., Witkin, A., and Terzopoulus, D., 1988. Snakes: Active Contour Models. International Journal of Conputer Vision, (1988)1:321-331. Beucher, S., and Lantéjoul, C. 1979. Use of Watersheds in Contour Detection. in International Workshop on Image Processing, Rennes, France. Koenderink, J., 1984. The Structure of Images. Biological Cybernetics, (50):363-370. Bezdek, J.C., 1992. Computing with Uncertainty. IEEE Communications Magazine, (1992)September:24-36. Larsen, M. 1999. Individual tree top position estimation by template voting. P. 8, in 21st Canadian Symposium on Remote Sensing, Ottawa, 21-24 June. Brandtberg, T., and Walter, F., 1998. Automated delineation of individual tree crowns in high spatial resolution aerial images by multiple scale analysis. Machine Vision and Applications, (1998)11:64-73. Lillesand, T.M., and Kiefer, R.W., 1994. Remote Sensing and Image Interpretation. John Wiley and Sons, Inc., New York Chichester Brisbane Toronto Singapore, 750p. Lindeberg, T., 1994. Scale-Space Theory in Computer Vision. Kluwer. Academic Publishers, Boston, USA, 423p. Brandtberg, T. 1999. Structure-based classification of tree species in high spatial resolution aerial images using a fuzzy clustering technique. P. 165-172, in The 11th Scandinavian Conference on Image Analysis, Kangerlussueq, Greenland, June 7-11. Lindeberg, T., 1998. Feature Detection with Automatic Scale Selection. International Journal of Computer Vision, (30)2:79-116. Cohen, L., 1991. On active contour models and balloons. CVGIP: Image Understanding, (53)2:211-218. Persson, A., Holmgren, J., and Söderman, U., 2002. Detecting and Measuring Individual Trees Using an Airborne Laser Scanner. Photogrammetric Engineering and Remote Sensing, (68)9:925-932. Dralle, K., and Rudemo, M., 1996. Stem Number Estimation by Kernel Smoothing of Aerial Photos. Canadian Journal of Forest Research, (1996)26:1228-1236. Pinz, A. 1989. Final Results of the Vision Expert System VES: Finding Trees in Aerial Photographs. P. 90-111, in Proceedings ÖAGM 13. Workshop of the Austrian Association for Pattern Recognition, Oldenbourg OCG Schriftenreihe. Egenhofer, M.J., and Herring, J.R., 1991. Categorizing Binary Topological Relations Between Regions, Lines, and Points in Geographic Databases. University of Maine, National Center for Geographic Information and Analysis, Orono, 28p. Pollock, R.J., 1996. The Automatic recognition of Individual trees in Aerial Images of Forests Based on a Synthetic Tree Crown Image Model. The University of British Columbia, Vancouver, Canada, June 1996. Gong, P., Sheng, Y., and Biging, G., 2002. 3D Model Based Tree Measurement from High-Resolution Aerial Imagery. Photogrammetric Engineering and Remote Sensing, (68)11:1203-1212. Pollock, R.J.. 1994. A model-based approach to automatically locating tree crowns in high spatial resolution images. P. 526-537, in Image and Signal Processing for Remote Sensing, 2315. SPIE. Gougeon, F., and Moore, T. 1988. Individual Tree Classification Using Meis-II Imagery. P. 927 -927, in IGARSS ’88 83 Schardt, M., Ziegler, M., Wimmer, A., Wack, R., and Hyyppä, R. 2002. Assessment of Forest Parameter by Means of Laser Scanning. P. 302-309, in The International Archives of the Photogrammetry, Remote Sensing and Spatial Information Sciences, XXXIV. 3A. ISPRS, Graz, Austria. Straub, B., and Heipke, C., 2001. Automatic Extraction of Trees for 3D-City Models from Images and Height Data. P. 267277. in Automatic Extraction of Man-Made Objects from Aerial and Space Images. Vol. 3. A.A.Balkema Publishers. Lisse/Abingdon/Exton(PA)/Tokio. Schneider, S., 1974. Luftbild und Luftbildinterpretation. de Gruyter, Berlin New York, 530p. Winter, S., 2000. Uncertain Topological Relations between Imprecise Regions. International Journal of Geographic Information Science, (14)5:411-430. Soille, P., 1999. Morphological Image Analysis: Principles and Applications. Springer, Berlin Heidelberg NewYork, 316p. Zadeh, L., 1965. Fuzzy Sets. Information Control, (8):338-353. 84 A Tree Tour with Radio Frequency Identification (RFID) and a Personal Digital Assistant (PDA) SEAN HOYT, DOUG ST. JOHN, DENISE WILSON AND LINDA BUSHNELL Abstract: A popular tree tour at the University of Washington campus has been automated via RFID and a PDA. The previous 81-tree hardcopy tour has also been updated to include more information on each tree, including digital photos. A survey conducted demonstrates the updated, electronic tree tour is easier to navigate, full of better visuals, and results in less false identifications. 85 86 Value Maximization Software – Extracting the Most from the Forest Resource HAMISH MARSHALL AND GRAHAM WEST Abstract: Global competition is encouraging all forest owners to manage their forested lands in more integrated manner and extract more value from the resource. ATLAS is a new suite of forest management software tools developed by Forest Research. The goal of the ATLAS concept is to have a suite of fully integrated software applications covering all forest management decisions from planting through to sawmilling. Three major applications have been developed so far: ATLAS Cruiser – a state-of-the-art forest inventory application, ATLAS GeoMaster – an advanced stand record systems and ATLAS Market Supply – one of the first weekly market supply planning optimization systems. In the future further applications covering growth modeling, saw mill optimization, strategic and tactical planning will be developed. This presentation will give a brief overview of the ATLAS system and highlights its key attributes 87 88 Costs and Benefits of Four Procedures for Scanning on Mechanical Processors GLEN E. MURPHY AND HAMISH MARSHALL Summary: Four simulated procedures for scanning and bucking Douglas fir, ponderosa pine and radiata pine trees were evaluated on the basis of productivity, costs, and value recovery. The procedures evaluated were: (a) a conventional operating procedure where quality changes and bucking decisions were input by the machine operator, (b) an automated scan of the full stem prior to optimisation and bucking, (c) a 6 m automated scan with 6.2 m forecast ahead, and (d) a 4.7 m automated scan with 7.5 m forecast ahead before optimal bucking took place. After subtracting costs, net value recovery for the automated scanning methods was 4 to 9% higher than for a conventional procedure. Breakeven capital investment costs for new scanning and optimisation equipment were dependent on tree species and size, markets and scanning procedure and could range between US$0 and US$1,400,000. 89 90 Evaluation of Small-Diameter Timber for Value-Added Manufacturing – A Stress Wave Approach XIPING WANG, ROBERT J. ROSS, JOHN PUNCHES, R. JAMES BARBOUR, JOHN W. FORSMAN AND JOHN R. ERICKSON Abstract- The objective of this research was to investigate the use of a stress wave technology to evaluate the structural quality of small-diameter timber before harvest. One hundred and ninety-two Douglas-fir and ponderosa pine trees were sampled from four stands in southwestern Oregon and subjected to stress wave tests in the field. Twelve of the trees, six Douglas-fir and six ponderosa pine, were harvested and sawn into logs and lumber. The mechanical properties of wood were then assessed by both stress wave and static bending techniques in the laboratory. Results of this study indicated a significant difference in stress wave time (SWT) between Douglas-fir and ponderosa pine trees and between two stands of each species. SWT of Douglas-fir trees increased slightly as tree diameter at breast height (DBH) increased; whereas, SWT of ponderosa pine trees decreased significantly as DBH increased. The statistical analysis also revealed good relationships between SWT of trees and modulus of elasticity (MOE) of logs and lumber produced from the trees as the two species were combined. However, the strength of the relationships was reduced within the species because of small sample size and narrow property range. INTRODUCTION This study is part of the project “Evaluation of smalldiameter timbers for value-added manufacturing: An integrated approach” conducted jointly by Oregon State University, USDA Forest Service Forest Products Laboratory, and USDA Forest Service PNW Research Station. The overall goal of the project was to design, construct, and deliver a system by which communities and forest industries may efficiently recognize value-added wood products potential in small diameter trees. The specific objective of this study was to investigate the use of a stress wave nondestructive evaluation technique to assess the potential structural quality of small-diameter timbers before timber harvest. Throughout the United States, past management practices have created thousands of acres of forest densely stocked with small-diameter trees. These stands are at increased risk of insect and disease attack and have higher catastrophic fire potential. Increased management emphasis on forest health and bio-diversity has forced land managers to seek economically viable stand treatments such as thinning to improve the stand condition. Economical and value-added uses for removed small-diameter timber can help offset forest management cost, provide economic opportunities for many small, forest-based communities, and avoid future loss caused by catastrophic wildfires. However, the variability and lack of predictability of the strength and stiffness of standing timber cause problems in engineering applications. It is essential to develop cost-effective technologies for evaluating the potential structural quality of such materials. The traditional log-to-product manufacturing process fails to recognize a tree’s full value. The process occurs in a series of mostly independent steps (trees, to logs, to lumber, to parts), each optimized for its own outputs. The ultimate end use is rarely a consideration during intermediate processing stages. By identifying final product potential before timber harvest, we hope to 1) enhance resource utilization efficiency, 2) make it economically viable for secondary wood products manufacturers to utilize small-diameter timber, and 3) facilitate stand management activities by identifying smalldiameter timber value. MATERIAL AND METHODS A total of one hundred and ninety-two Douglas-fir (Pseudotsuga menziesii) and ponderosa pine (Pinus ponderosa) trees were sampled for stress wave evaluation at four different stands in southwestern Oregon. The stands were located in the Applegate Ranger District on the Rogue River National Forest. Stand A (Yale Twin) was a 70 year old evenaged stand consisting primarily of Douglas-fir with some madrone and a small compliment of ponderosa pine. The stand had a mean diameter of 6.4 inches (16.3 cm) and a quadratic mean diameter of 7.4 inches (18.8 cm). Stand B (Toe Top) consisted of a sparse stand of 90-year-old trees (primarily ponderosa pine) with a 65 year old under-story of Douglas-fir, smaller ponderosa pine, madrone, and an occasional incense cedar. It had a mean diameter of 6.0 91 inches (15.2 cm) and a quadratic mean diameter of 7.8 inches (19.8 cm). Both stand A and B were slow grown and stagnant, and the trees marked for thinning and testing had small branches. Stand C (Squaw Ridge) was a 40-year-old evenaged ponderosa pine stand with a minor compliment of Douglas-fir. The trees were vigorous and fast-growing, with large crowns and large branch diameters. The stand had a mean diameter of 8.7 inches (22.1 cm) and a quadratic mean diameter of 9.4 inches (23.9 cm). Stand D (No Name) was a mixture of Douglas-fir and ponderosa pine, with some madrone peristing in the understory. Tree age ranged from 35 to 40 years. The stand had a mean diameter of 7.2 inches (18.3 cm) and a quadratic mean diameter of 8.0 inches (20.3 cm). All sampled trees were subjected to stress wave tests in the field. Douglas-fir trees were evaluated in stands A and B, and ponderosa pine trees were evaluated in stands C and D. Trees of each stand were classified into six diameter classes that had a mean diameter at breast height (DBH, measured outside bark) of 5, 6, 7, 8, 9, and 10 inches (12.7, 15.2, 17.8, 20.3, 22.9, and 25.4 cm) respectively. A random sample consisting of eight trees per diameter class was subjected to stress wave tests in each of the four stands. A recently developed stress wave technique was used to conduct in-situ tests on sampled trees (Wang 1999, Wang et al 2001). The testing system consisted of two accelerometers, two spikes, a hand-held hammer, and a portable scopemeter (Figure 1). Two spikes were imbedded in the trunk at 45o to the trunk surface, one spike at each end of the section to be assessed with a span of 4 ft (1.2 m). The spikes were pounded into the stem about one inch (2.5 cm), which was deep enough for the tips to penetrate the bark and reach the sapwood. The Accelerometers were mounted on the spikes using two specially designed clamps. A stress wave was introduced into the tree in the longitudinal direction by impacting the lower spike with a hammer. The resulting signals were received by start and stop accelerometers and recorded on the scopemeter as waveforms. The stress wave time (SWT, the time for a stress wave to travel through the distance between two spikes) was determined by locating the two leading edges of the waveforms on the scopemeter (Wang et al 2001). Six measurements were obtained on each tree, three on each of two sides. After field tests, one tree per diameter class was felled in stands B (Toe Top) and C (Squaw Ridge), resulting in a sample of six Douglas-fir and six ponderosa pine trees ranging from 5 to 10 inches (12.7 to 25.4 cm) in DBH. These felled trees were then bucked into 10-foot (3.0 m) long logs and transported to Michigan Technological University in Houghton, Michigan for laboratory tests. For each log, the green weight and diameters (at two ends and the middle of the log) were measured and the green density was determined accordingly. All logs were then evaluated using longitudinal stress wave and static bending methods to obtain stress wave time and static modulus of elasticity (MOE) of the logs. A detailed description of the instrumentation and analysis procedures for log tests is given by Wang et al. (2002). Standing tree Accelerometer L Oscilloscope Figure 1. Schematic of experimental setup used in field test (L = test span). To validate the stress wave analysis of trees and logs, all logs were sawn into 2- by 4-in. (51 by 102-mm) and 2- by 6in. (51- by 152-mm) dimension lumber on a portable horizontal band sawmill for further assessment in terms of structural quality. Sawing pattern for each log was diagrammed so that the location of each piece of lumber within each log could be tracked. Each piece of lumber received a unique identification number associating it with its location within the log and tree from which it was sawn. The lumber was stickered and stacked for air-drying until they reach the moisture content of approximately 15 percent. When dry, the lumber was planed to industry standard thickness and width for surfaced dry lumber. Longitudinal stress wave and static bending tests were also conducted on lumber at both green and dry conditions. RESULTS AND DISCUSSION Stress Wave Time in Standing Trees The stress wave time in standing trees was the average value of six measurements from each tree and was reported on the unit per length basis (time/length). Lower stress wave time corresponds to higher stress wave speed (length/time). The descriptive statistics of tree measurements (SWT and DBH) from all tree samples are given in Table 1. Figure 2 shows histograms of stress wave time distribution for four different stands. The difference between Douglas-fir and ponderosa pine can be easily distinguished in terms of stress wave time. The mean SWT of ponderosa pine trees is about 27 percent higher than that of Douglas-fir trees, which means stress waves travel much slower in ponderosa pine than in Douglas-fir trees. In general, this result is in agreement with the strength and stiffness difference between the two species as given in the Wood Handbook (FPL 1999), which states the modulus of rupture (MOR) and modulus of elasticity of ponderosa pine are about 34 percent lower than those of Douglas-fir (green condition). 92 14 (a) Douglas-f ir Stand A:Yale Tw in Stand B: Toe Top 12 Frequency (%) The SWT of ponderosa pine trees also shows much higher variation than the SWT of Douglas-fir trees. The standard deviation of SWT is 4.50 ms/ft (14.8 ms/m) for Douglas-fir (stand A and stand B combined), and 16.17 ms/ft (53.0 ms/ m) for ponderosa pine (stand C and stand D combined). This might suggest a larger variation in strength and stiffness properties of ponderosa pine compared to those of Douglas-fir. The statistical comparison analysis showed significant SWT differences between two stands of each species, which imply a potential difference in strength and stiffness between the stands. But this could not be substantiated due to the lack of mechanical property data of all tested standing trees. The relationship between SWT and DBH of standing trees is shown in Figure 3. For better illustration, stress wave times in trees were analyzed in terms of diameter classes. The data points are mean values of SWT for eight trees in each class, and the error bar indicates the standard deviations (±1 standard deviation). The SWT in Douglas-fir trees increased slightly as DBH of the trees increased. The trend is more evident in stand A (Yale Twin) than in stand B (Toe Top). The SWT for stand A increased about 12 percent as DBH changed from 5 in. to 10 in. (12.7 to 25.4 cm). The SWT-DBH relationship for ponderosa pine trees was quite different from Douglas-fir. As shown in Figure 3(b), the SWT in ponderosa pine trees decreased significantly as DBH of the trees increased, especially in stand C (Squaw Ridge) where the SWT dropped 24 percent when the DBH increased from 5 in. to 10 in. (12.7 to 25.4 cm). The causes for the different functional relationships between SWT and DBH for Douglas fir and ponderosa pine trees are not fully understood yet. Huang (2000) reported that, for the same age trees, stress wave time is lower for trees with slower growth rate or narrower rings. This might explain the SWT-DBH trend found in Douglas 10 8 6 4 2 0 60 62 64 66 68 70 72 74 76 78 80 82 84 86 88 90 Stress w ave time in trees (µs/ft) 14 (b) Ponderosa pine Stand C: Squaw Ridge Stand D: No name Frequency (%) 12 10 8 6 4 2 0 60 70 80 90 100 110 120 130 140 150 Stress w ave time in trees (µs/ft) Figure 2. Histograms of stress wave time (SWT) distribution for Douglas-fir and ponderosa pine trees. Table 1. Diameter at breast height and stress wave time of standing trees. a Sample Species Douglas-fir Ponderosa pine a SWT (µs/ft) DBH (in.) Stand No. Mean Min Max SD Mean Min. Max. SD A 48 7.4 4.7 10.3 1.71 75.2 67.3 87.0 4.79 B 48 7.5 4.6 10.2 1.76 72.8 60.8 79.2 3.83 C 48 7.6 4.8 10.3 1.76 98.9 77.3 150.0 14.94 D 48 7.6 4.7 10.1 1.77 89.2 71.3 134.3 16.12 1 in. = 2.54 cm, 1 µs/ft = 3.28 µs/m. DBH, diameter at breast height. SWT, stress wave time. SD, standard deviation. 93 fir trees. For ponderosa pine trees, the opposite SWT-DBH trend could be more related to other factors such as the characteristics of tree forms (size and frequency of branches), proportion of mature and juvenile wood in the cross section as well as moisture content. 100 (a) Doudlas f ir Stand A SWT in trees ( s/ft) 90 Relationship Between Stress Wave Time in Trees and Log Properties Stress wave time in standing trees was measured in the lower part of the stem, which tracks to the butt log after harvesting and cutting. In this study, a total of 42 10-ft. (3.0-m) long logs were obtained from 12 harvested trees. The number of the logs produced from each tree varied from 3 to 5 for Douglas-fir and from 1 to 4 for ponderosa pine as a result of the difference in tree height. The diameter of the logs (average value of diameters measured at two ends and the middle) ranged from 4.3 to 10.0 in. (10.9 to 25.4 cm) for Douglas-fir and from 4.4 to 9.8 in. (11.2 to 24.5 cm) for ponderosa pine. The physical and mechanical properties (density, stress wave time, and static MOE) of logs are summarized in Table 2. Note that all these properties were determined in green and un-debarked logs. Figure 4 shows the relationship between SWT of trees and SWT of the butt logs cut from the trees. A linear regression analysis indicated a strong correlation (R2 = 0.95) when two species were considered as a single population. The strength of the relationship was weakened when the two species were considered separately (R2 = 0.61 for Douglasfir, R2 = 0.85 for ponderosa pine). This was presumably due to the small sample size (n=6) and limited property range for samples of each species. It was found that SWT measured in standing trees was about 10 and 22 percent lower than SWT of logs for Douglas-fir and ponderosa pine, respectively. This could be a systematic difference caused by different stress wave approaches. It has been reported that the stress wave speed measured in trees could be dominantly controlled by the mature wood (outer wood in the crosssection) since both wave generation and sensing occurred on the surface of the stem (Wang 1999, Huang 2000, Ikeda Stand B 80 70 60 50 40 4 5 6 7 8 9 10 11 DBH of standing trees (in.) SWT in trees ( s/ft) 160 150 (b) Ponderosa pine Stand C 140 130 120 Stand D 110 100 90 80 70 60 50 40 4 5 6 7 8 9 10 11 DBH of standing trees (in.) Figure 3. Relationship between stress wave time (SWT) and tree diameter at breast height (DBH). Table 2. Physical and mechanical properties of logs. a Species logs Douglas-fir Ponderosa pine a Stress wave time (µs/ft) 3 No. of Density (lb/ft ) Mean Min. Max. SD Mean Min. Max. SD Mean Min. Max. 2 70.4 84.1 SD 25 41.45 35.17 48.45 3.539 76.1 3.68 0.99 0.52 1.33 0.213 17 51.75 43.12 57.55 4.310 116.3 106.0 134.7 8.84 0.57 0.33 0.79 0.149 1 lb/ft = 16.02 kg/m , 1 µs/ft = 3.28 ms/m, 1 lb/in = 6895 Pa. 3 6 MOE (10 lb/in ) 3 2 MOE, modulus of elasticity determined by static bending method. SD, standard deviation 94 et al. 2000, and Wang et al. 2001), whereas in logs the waves were introduced into the stem from one end and sensed at the other end (Wang et al. 2002). The relationships between SWT of trees and the average MOE of logs are shown in Figures 5. Regression analysis indicated a linear relationship between SWT of trees and MOE of logs as all samples combined. The coefficient of determination (R2) was found to be 0.74. Again, the strength of the relationships was reduced significantly as two species were analyzed separately. 1.2 MOEs of logs (Mpsi) 1 Ponderosa pine 20 40 60 80 100 120 SWT in trees (µs/ft) 140 160 Figure 5. Relationship between SWT in trees and average MOE of logs. 3 Average MOEs of lumber (106 lb/in2) SWT SWT in in butt butt logs ((µs/ft) s/ft) Douglas-fir 0 Douglas-fir Ponderosa pine 130 0.4 0 A total of 81 pieces dimension lumber (2 by 4s and 2 by 6s), 49 Douglas-fir and 32 ponderosa pine, were obtained from the logs. Stress wave and static bending tests were performed on lumber in both rough-cut and dry conditions (air dried and 4-side surfaced). The moisture content (MC) of rough-cut lumber (designated as green lumber) ranged 140 0.6 0.2 Relationship between Stress Wave Time in Trees and Lumber MOE 150 0.8 120 110 100 90 80 70 60 2.5 2 1.5 1 Douglas-fir (Green) Douglas-fir (Dry) Ponderosa pine (Green) Ponderosa pine (Dry) 0.5 0 50 50 60 70 0 80 90 100 110 120 130 140 150 SWT in trees (µs/ft) 20 40 60 80 100 120 140 SWT in trees (µs/ft) Figure 6. Relationships between SWT in trees and average MOE of lumber produced from the trees. Figure 4. Relationship between SWT in trees and SWT of butt logs. Table 3. Stress wave and static bending properties of lumber Rough-cut lumber (MC = 24%) Species Number SWT MOE of lumber (? s/ft) Douglas-fir 49 Ponderosa pine 32 a 2 Dry lumber (MC = 9%) SWT MOE (10 lb/in ) (? s/ft) (10 lb/in ) 67.8 (4.0) 2.14 (12.4) 59.1 (4.3) 2.60 (11.8) 115.3 (11.1) 1.06 (14.2) 78.1 (11.4) 1.33 (14.5) 6 2 1 ? s/ft = 3.28 ? s/m, 1 lb/in = 6895 Pa. SWT, stress wave time. MOE, modulus of elasticity determined by static bending method. COV, coefficient of variation (%). Data in parenthese represents coefficients of variation (%). 95 6 2 from 19 to 26 percent for Douglas-fir with an average of 24 percent and 30 to 42 percent for ponderosa pine with an average of 36 percent. The MC of dry lumber was 8 to 10 percent with an average of 9 percent for both species, which was actually lower than target MC. The averages and coefficients of variation (COV) for stress wave and static bending properties of lumber are summarized in Table 3. The mean comparison results indicated a significant difference between SWT in trees and SWT in lumber. For Douglas-fir, the mean SWT in rough-cut and dry lumber decreased about 7 and 17 percent respectively compared to the mean SWT in trees. The low SWT in lumber is mainly due to the low moisture content (the MC was below fiber saturation point for both rough cut and dried lumber). For ponderosa pine, however, the mean SWT in green lumber (rough cut) increased about 19 percent compared to that in trees. This could be caused by the different wave propagation mechanisms associated with the testing approaches used in tree and lumber measurements. As mentioned earlier, the SWT measured in trees is more controlled by the mature wood (outer wood in the cross-section) compared to the SWT measured in logs. The same interpretation could be reached for lumber. The expectation is that, given the same moisture condition, the SWT in trees would be lower than the SWT in lumber. In terms of moisture effect, since the MC of green ponderosa pine lumber was well above the FSP, the moisture has less effect on the SWT compared to Douglas-fir lumber. Therefore, the high SWT in ponderosa pine green lumber might be mainly due to the different wave propagation mechanism. In the case of dried ponderosa pine lumber (the MC was far below the FSP), the mean SWT decreased about 19 percent compared to that in trees because the moisture effect played a more important role compared to wave propagation mechanism. The relationships between SWT in trees and average MOE of lumber produced from the trees are shown in Figure 6. In the case of Douglas-fir, both tree and lumber property range was very small, and no statistical relationship was found between SWT of trees and average MOE of lumber. In the case of ponderosa pine, the data points had a wider property range (tree and lumber) and shown a linear relationship between SWT of trees and average lumber MOE (R2 = 0.39 – 0.63). When the two species were combined, the statistical analysis resulted in a good correlation between SWT of trees and average MOE of lumber. The coefficients of determination (R2) were found to be 0.88 for green lumber and 0.86 for dry lumber. difference in stress wave time between Douglas-fir and ponderosa pine trees. Stress wave time ranged from 60.8 to 87.0 ms/ft (199 to 285 ms/m) for Douglas-fir trees and 71.3 to 150 ms/ft (234 to 492 ms/m) for ponderosa pine trees. Statistical comparison analysis between stands suggested a potential difference in wood stiffness between the two stands of each species. It was found that stress wave time in Douglas-fir trees increased slightly as tree diameter at breast height increased; whereas, stress wave time in ponderosa pine trees decreased significantly as tree diameter at breast height increased. The statistical analysis resulted in good relationships between stress wave time of trees and modulus of elasticity of logs and lumber when the two species were combined. However, the statistical significance was reduced as the two species were considered separately because of small sample size and narrow property range within each species. The data colleted for this study illustrates the potential of the stress wave technique for assessing the structural quality of small-diameter timbers in the field. Further studies are planned to develop a broader database of SWT-MOE relationship with sufficient samples for each species, and examine if species has an effect on the relationship. LITERATURE CITED Forest Products Laboratory. 1999. Wood Handbook – Wood as an engineering material. Gen. Tech. Rep. FPL-GTR-113. Madison, WI: U.S. Department of Agriculture, Forest Service, Forest Products Laboratory. 463 p. . Huang, Chih-lin. 2000. Predicting lumber stiffness of standing trees. In: Proceedings of the 12th international symposium on nondestructive testing of wood; 2000 September 13-15; Western Hungary, Sopron, University of Western Hungary, Sopron: 173-180. Ikeda, K., S. Oomori, and T. Arima. 2000. Quality evaluation of standing trees by a stress-wave propagation method and its application III: Application to sugi (Cryptomeria japonica) standing plus trees. Mokuzai Gakkaishi. 46(6): 558-565. Wang, X. 1999. Stress wave-based nondestructive evaluation (NDE) methods for wood quality of standing trees. Ph.D. diss. Michigan Technological Univ., Houghton, MI. CONCLUSIONS Wang, X., R. J. Ross, M. McClellan, R. J. Barbour, J. R. Erickson, J. W. Forsman, and G. D. McGinnis. 2001. Nondestructive evaluation of standing trees with a stress wave method. Wood and Fiber Science, 33(4): 522-533. A stress wave technique was used to evaluate the structural potential of small-diameter Douglas-fir and ponderosa pine trees. The results of the study indicated a significant Wang, X., R.J. Ross, J.A. Mattson, J.R. Erickson, J.W. Forsman, E.A. Geske, M.A. Wehr. 2002. Nondestructive evaluation techniques for assessing modulus of elasticity and stiffness of small-diameter logs. Forest Prod. J. 52(2): 79-85. 96 Early Experience with Aroma Tagging and Electronic Nose Technology for Log and Forest Products Tracking GLEN MURPHY Abstract: Worldwide the movement of logs from forest to customer can be conservatively estimated at over 5 billion logs per annum. There is increasing interest in being able to track the movement of individual logs from stump to mill or at least determine the chain-of-custody of groups of logs back to individual stands. Some segments of industry would ideally like to be able to track wood products from the standing tree through to the ultimate product – “from seedling to rocking chair”. Barcoding and radio frequency identification, although not ideal, are the dominant technologies for tagging and tracking of forest products. This presentation will cover early experience with a novel technology, aroma-tagging and an electronic nose, for tracking logs from the forest through the mill and out of the drying kilns. Trials indicate that this novel technology is most likely to be successful in chain of custody applications from forest to mill door. Development of new tools or further development of new procedures could eventually result in the ability to track of individual logs. 97 98 Modeling Steep Terrain Harvesting Risks Using GIS JEFFREY D. ADAMS, RIEN J.M. VISSER, AND STEPHEN P. PRISLEY Abstract: When preparing to harvest timber on steep terrain, it is necessary to assess a variety of risks, including slope failure, excessive erosion, residual stand damage, and job-related injury. A number of the risks associated with steep terrain harvesting can be modeled using terrain and soil characteristics such as slope gradient, slope form, soil strength, and soil erodibility. Once assessed, these risks can often be mitigated through detailed harvest planning, an important part of which is the selection of an appropriate harvesting system. This paper describes the development of a steep terrain harvesting risk assessment model using ArcObjectsä. The model operates within the Visual Basic for Applicationsä (VBA) environment embedded in ArcMapä, and accepts soil and digital elevation data as inputs into a decision matrix containing key steep terrain harvest system parameters. Model outputs include maps depicting debris slide hazard, soil strength hazard, soil erosion hazard, and harvest system recommendations. The intended use of the model is to serve as a decision support system in the strategic planning phase of forest management, facilitating the identification of high-risk areas and long-term harvesting system requirements. An application of the model is demonstrated on approximately 500 hectares of mountainous terrain in southwest Virginia. and personal injury can often lead to significant direct and indirect costs for companies and injured parties. The factors that contribute to the existence of the abovementioned hazards are often unalterable features of the terrain. However, many of the adverse impacts associated with the hazards can be mitigated through informed planning. To properly assess the severity and extent of the hazards, it is often necessary to conduct detailed field investigations in which site-specific data is collected and analyzed. When properly assessed, one of the more effective ways to mitigate the identified hazards is to select and apply an appropriate harvesting system. For the purposes of this research effort, harvesting system will refer specifically to the equipment and techniques used to move felled trees from the stump to the landing. Harvesting systems commonly used in mountainous terrain include wheeled skidder, track skidder, cable, and helicopter systems. The objective of this research was to design a GIS model that could serve as a decision-support tool during both the strategic (long-term) and tactical (short- and medium-term) planning phases of forest management planning. During the strategic phase, when forest-level management concerns are being addressed, the model can be used to assess longterm harvesting system requirements. The model provides estimates of the proportions of a land base that might be appropriate for the different harvesting systems, which can help forest managers and planners refine projected harvesting costs and determine whether the necessary equipment or an adequate supply of harvesting contractors is available. Model outputs also include the relative location, severity, and geographic extent of the environmental hazards associ- INTRODUCTION In many mountainous regions, planning forest management activities can be complicated by a variety of terrain factors (slope gradient, slope form, topographic complexity, etc.) and host of soil characteristics (strength, erodibility, etc.). This is particularly true in southwest Virginia, where the topography is extremely diverse due to the convergence of the Appalachian Plateau, Ridge and Valley, and Blue Ridge physiographic provinces. In many locations throughout the region, it is necessary to assess a number of potential environmental hazards when planning timber harvesting operations. The more prominent hazards associated with conducting timber harvesting operations on mountainous terrain include soil erosion, soil compaction, and debris slides. Depending on the severity and extent of the hazard, each can potentially lead to significant adverse environmental and economic impacts if not properly assessed and managed. Soil compaction can retard the growth of regeneration as well as lead to increased soil erosion (Martin 1988). Soil erosion, a common byproduct of timber harvesting on steep terrain, can lead to decreases in forest site productivity, water quality, and stream habitat (Rice, et al. 1972). Debris slides can rapidly deliver sediment and woody debris to waterways resulting in high turbidity, bank scouring, channel aggradation, and potential damage to roads and other improvements in their paths (Washington State Forest Practices Board 2000). In addition, steep terrain harvesting operations carry a greater risk of equipment damage and personal injury than operations conducted on flat terrain. Equipment damage 99 and can range from a few years to decades (Martin 1988, Schnepf 2002). Quite often, debris slides represent the dominant erosional process in steep mountainous terrain (Wu and Sidle 1995). Debris slides are mass failures in which the internal strength of soil is exceeded by a variety of stressors, including gravity, soil pore pressure, and material weight (Dietrich, et al. 1986, Shaw and Johnson 1995). They commonly occur in convergent topography, where water, sediment, and organic debris become concentrated (Dietrich, et al. 1986). Areas prone to debris slides will infrequently experience recurrent activity, usually triggered by intense rainfall events. While debris slides are a natural process, certain forest management activities are believed to increase the frequency and severity of debris slide activity. As with surface erosion, the management features commonly associated with debris slide activity are poorly located or constructed roads. In addition to environmental damage, conducting poorly planned timber harvest operations in steep terrain can result in equipment damage and worker injury. Logging is one of the most hazardous occupations, with a rate of occupational death, illness, or injury approximately 3 times greater than the average incident rate for all private industries. As slope gradient increases, so too does the potential for injury and accident. Most ground-based harvesting equipment such as wheeled and track skidders possess relatively high centers of gravity and can overturn in steep or uneven terrain (Conway 1982). The majority of groundbased and aerial systems (cable and helicopter) require manual felling. Falling materials (i.e. trees, snags, and branches) and poor felling practices are common causes of injury and death for tree fellers. This is especially true in locations characterized by complex stand structures and steep terrain, such as the mixed hardwood stands of the Appalachians. The high-tension cables used in cable yarding operations pose additional threats to workers on the ground. Lastly, helicopter operations can be extremely dangerous, with crashes leading to severe injury or death to both pilots and loggers (Manwaring and Conway 2001). ated with steep terrain harvesting. During the tactical phase of management planning, these hazard assessments can be used to prioritize field investigation activities. To maximize the model’s operability and accessibility, data requirements were limited to widely distributed, publicly available spatial data. To provide examples of model output, an analysis was conducted on approximately 500 hectares of mountainous terrain that serves as a teaching and demonstration forest for Virginia Polytechnic Institute and State University. STEEP TERRAIN HARVESTING RISKS When conducting timber harvest operations in steep terrain, it is necessary to mitigate a number of risks. The sedimentation of waterways resulting from increased surface erosion is often cited as the primary concern associated with forest management activity in steep terrain. Many of the streams originating in or flowing through steep forested terrain provide important habitat for aquatic species and represent important sources for water supplies, recreation, and a number of other uses. Sedimentation of these streams can have adverse impacts on water quality and aquatic habitat, as well as lead to increased flood potential (Virginia Department of Forestry 2002). As a result, many states have established Best Management Practices (BMP) for forest management activities. BMPs identify forest management activities that mitigate increased erosion. Management activities that are commonly identified as potential contributors to increased surface erosion include logging operations, road construction, grazing, and site preparations associated with planting and fire (Toy, et al. 2002, Virginia Department of Forestry 2002). Of the above listed activities, road construction is widely recognized as the biggest potential contributor to increased surface erosion. Although some degree of increased erosion may be unavoidable, measures can be taken to minimize the severity and extent of erosion (Rice, et al. 1972). Another concern associated with steep terrain harvesting is the compaction of soil caused by the ground pressure exerted by heavy harvesting equipment. Soil compaction alters the physical properties of a soil by reducing the amount of macropore space and increasing density. While soil compaction is a hazard that should be assessed for any harvesting operation, the amount of ground pressure exerted by harvesting equipment is greater when operating on uneven or sloping terrain (Adams 1998). The physical changes brought about by compaction can have significant adverse impacts, including restricted rooting depths for regeneration, restricted water and nutrient cycling, increased water runoff, and increased surface erosion hazard (Adams 1998, Krag, et al. 1986, Martin 1988, Miller and Sirois 1986, Rice, et al. 1972, Schnepf 2002). Compacted soils can be restored given an adequate period of time and the proper environmental conditions. The amount of time required to restore compacted soils depends on the severity of the disturbance, HARVESTING SYSTEMS Harvesting systems commonly used throughout the Appalachians and other mountainous regions include wheeled skidders, track skidders, cable yarders, and helicopters. Under a broad range of conditions, the wheeled skidder system represents the most efficient ground-based alternative. Wheeled skidders are rubber-tired vehicles specially outfitted to transport felled timber. They require a relatively low initial capital investment, are relatively inexpensive to maintain, and can move a given quantity of wood from the stump to the landing up to twice as fast as their tracked counterparts (Conway 1982). Wheeled skidders travel through harvested areas on a network of skid roads and skid trails. Skid roads, which are the primary routes from the harvested area to the landing, are often systematically located throughout 100 the harvested area and experience heavy use during a harvesting operation. In steep terrain operations, skid roads are often located on cut-and-fill slopes. Skid trails are secondary routes established while accessing felled timber and can be somewhat random in location. Skid roads and skid trails can be major sources of erosion in steep terrain (Gibson and Biller 1975, Krag, et al. 1986, Rice, et al. 1972). Track skidders, often referred to as crawler tractors, are specially outfitted tracked vehicles used to transport felled timber. While slower and more expensive than their wheeled counterparts, track skidders can be much more versatile. They are capable of transporting larger payloads and can be used to construct roads and landings (Conway 1982). In some situations, soil disturbance impacts can be mitigated by switching from wheeled to track vehicles (Martin 1988). Track skidders spread their weight over a much larger area, which can significantly reduce the severity of soil compaction and rutting. This is particularly true for operations conducted on wetter sites, where wheeled skidders can also suffer significant decreases in pulling power (Conway 1982). Aerial systems such as cable yarders and helicopters are commonly used in locations possessing gradients too steep for the safe and productive implementation of ground-based systems. In cable harvesting systems, felled trees are rigged to a suspended cable and pulled to the landing with winch systems called yarders. Depending upon the configuration of the system being used, felled trees are suspended either partially or fully off the ground. In general, the soil distur- bance associated with cable systems is less severe and widespread than the disturbance caused by ground-based systems, due in most part to the lack of skid roads and trails (Krag, et al. 1986, Miller and Sirois 1986). A necessary feature of any cable system configuration is deflection, which is sag in the suspended skyline cable. In general, a minimum deflection of 5% is required for a skyline to possess an acceptable load-carrying capability. Cable operations are typically conducted on terrain characterized by concave ground profiles, which allow for adequate deflection. Helicopter systems are the most expensive alternative and applied when all other systems are deemed inappropriate. For the most part, the use of helicopter systems is relegated to remote locations that are very sensitive to adverse environmental impacts. Trees are felled manually and then transported to the landing using a helicopter. The use of helicopters eliminates skid road construction, soil rutting associated with skid trails, and corridor damage associated with cable systems. However, large landings with access roads capable of heavy transport traffic are required, typically within a 3-mile distance of the harvested area (Sloan 2001). METHODS In order to provide an automated spatial assessment of the risks associated with terrain and soil conditions, a GISbased model was developed. The model operates within the Figure 1. Screen capture of the user interface, which contains a set of tabbed pages on which the user identifies the model input, selects output options, and can adjust model parameters for the different hazards assessed. 101 respect to the intensity and scale at which the soil units are mapped, with SSURGO being the most detailed. The model is designed to accept either SSURGO or STATSGO data. The soil units in SSURGO datasets are mapped at scales ranging from 1:12,000 to 1:63,000 and can contain up to three different soil components. The availability of SSURGO datasets, while increasing, is currently limited to select locations throughout the conterminous United States, Alaska, Hawaii, and Puerto Rico. STATSGO datasets are available for the entire conterminous United States, Alaska, Hawaii, and Puerto Rico. STATSGO soil units can contain up to 27 different soil components, and with the exception of Alaska (1:1,000,000), are mapped at a scale of 1:250,000. Visual Basic for Applicationstm (VBA) environment embedded in ArcMaptm, and accepts soil and digital elevation data as inputs into a decision matrix containing key steep terrain harvest system parameters. The interface of the model contains a set of tabbed pages on which the user identifies the model input, selects output options, and can adjust model parameters for the different hazards assessed (Figure 1). Default parameter values are provided, however, adjustments can be made to suit local conditions or knowledge. Model outputs include tabular and spatial output depicting soil erosion hazard, soil compaction hazard, debris slide hazard, and harvest system allocation. STUDY AREA SOIL EROSION HAZARD MODELING The study area selected to illustrate model operation is the Fishburn Forest, a teaching and demonstration forest owned by Virginia Polytechnic Institute and State University. The forest is situated on an isolated, east-west trending ridge in the Valley and Ridge province of southwest Virginia and is comprised of approximately 500 hectares of Appalachian hardwood and mixed pine-hardwood cover types. Elevations range from approximately 550–730 meters above sea level with a mean and standard deviation of 629 and 39, respectively. Slope gradients in the forest range from 0-112%, with a mean and standard deviation of 28 and 15, respectively. Within the boundaries of the forest, the following soil series are represented: Berks, Caneyville, Craigsville, Duffield, Groseclose, Jefferson, McGary, and Weaver series. Soil erosion hazard is modeled using a combination of slope gradient classes and Kffact. Kffact is an experimentally determined value that quantifies the susceptibility of soil particles to detachment and movement by water (Natural Resources Conservation Service 1995). Kffact values can range from 0 to 1, higher values indicating greater erosion potential. In both SSURGO and STATSGO datasets, each map unit can contain multiple soil components and each component is typically comprised of multiple layers, each of which is assigned a Kffact value. To characterize soil erosion hazard, the model required that each map unit be represented by only one Kffact value. For each soil component within a particular map unit, the relevant Kffact value for the modeling of surface erosion is the Kffact value associated with the soil layer constituting the thickest mineral horizon in the upper 15 cm of the component (Natural Resources Conservation Service 1998). As such, each map unit contained multiple soil components represented by the Kffact value attributed to the soil layer meeting the above-described conditions. To provide the most conservative estimate of soil erosion hazard, the highest Kffact value from the set of soil components contained within the map unit was attributed to the particular map unit. The representative Kffact value and slope gradient were then combined to characterize relative soil erosion hazard. The default soil erosion hazard classification criteria (Table 1) offered by the model is adapted from interpretive DATA REQUIREMENTS The data requirements for the model include elevation and soil data, both of which represent important data sources for GIS applications in a variety of disciplines, including engineering, ecology, hydrology, natural resource management and geomorphology. With respect to elevation data, the model is designed to accept grid-based data with either 30-meter or 10-meter horizontal resolution. The United States Geological Survey (USGS) produces both 30-meter and 10-meter grid-based digital elevation models as part of the National Mapping Program (U.S. Geological Survey 1987). While the availability of 10-meter elevation data is still somewhat limited, 30-meter data is available to the public for a majority of the conterminous United States, Hawaii, and Puerto Rico. With respect to soil data requirements, the United States Department of Agriculture’s (USDA) Natural Resources Conservation Service (NRCS) distributes three spatial soil databases, including the Soil Survey Geographic (SSURGO), State Soil Geographic (STATSGO), and National Soil Geographic (NATSGO) databases. The databases consist of mapped soil units (polygons) and a collection of relational tables containing associated physical properties, chemical properties, and interpretations. The databases differ with Table 1. Default slope gradient classes and Kffact values used to characterize relative soil erosion hazard. Soil Erosion Hazard Lower Moderate Higher 102 Kffact < 0.35 0 - 25% 25 - 45% > 45% Kffact ≥ 0.35 0 - 17% 17 - 35% > 35% criteria used by the NRCS to rate potential off-road/off-trail erosion hazard (Natural Resources Conservation Service 1998). on the criteria used by the NRCS to rate log landing suitability, natural surface road suitability, and harvest equipment operability (Natural Resources Conservation Service 1998). Where slope gradient exceeded 20%, lower and moderate ratings were shifted to moderate and higher ratings, respectively. SOIL COMPACTION HAZARD MODELING DEBRIS SLIDE HAZARD MODELING Soil compaction hazard is modeled using a combination of Unified Classification soil group designations and slope gradient. The Unified Classification System was developed by the Army Corps of Engineers in 1952 and classifies soils into groups based on a number of characteristics, including grain size, gradation, liquid limit, and plasticity index (Cernica 1995). Unified Classification designations are used in a number of NRCS interpretive ratings as an indicator of soil strength for forestry-related activities. Up to four different Unified Classification group designations are provided for each soil layer in a soil component. For each soil component, the relevant Unified Classification designations with respect to the modeling of soil compaction are the group designations attributed to soil layers located in the upper 15 cm of the component that are ≥ 7cm in thickness. For the purposes of modeling protocol, each map unit can only be represented by a single Unified Classification designation. As with the soil erosion hazard modeling described above, the algorithm used to obtain a map unit’s representative Unified Classification group designation was designed to provide the most conservative estimate of soil compaction hazard. This was achieved by first selecting the most limiting of the multiple designations attributed to each layer located in the upper 15 cm of the component that were ≥ 7 cm in thickness. This designation was subsequently attributed to the component to which the layer belonged. The most limiting designation was then selected from the set of designations corresponding to the soil components in the map unit. The representative group designation was assigned to the map unit, and used to characterize the relative soil compaction hazard. The default classification scheme (Table 2) used by the model is based Debris slide hazard is modeled using slope gradient and slope form. The protocol to produce hazard ratings is adapted from a slope morphology model developed by the Washington Department of Natural Resources (Shaw and Johnson 1995). Slope gradient is calculated from the elevation data and classified into low, moderate, steep, and very steep classes (Table 3). Slope form is captured spatially using planform surface curvature, which proved to be very effective in the identification of the landforms commonly associated with debris slide occurrences. Planform surface curvature is also calculated from the elevation data and classified into convex, planar, and concave classes (Table 4). The combination of the slope gradient and slope form classes provide a matrix from which debris slide hazard classes are derived. The default matrix used by the model to rate debris slide hazard from the slope gradient and slope form classes is provided in Table 5. Table 3. Slope gradient classification parameters used in the modeling of debris slide hazard. Slope Gradient Class Low Moderate Steep Very Steep Table 4. Slope form classification parameters used in the modeling of debris slide hazard. Table 2. Default classification scheme used to characterize relative soil compaction hazard. Soil Compaction Hazard Lower1 Moderate1 Higher Slope Gradient (%) 0 - 25 25 - 45 45 - 65 >65 Unified Classification Group Other CL, CH, CL-ML, ML, MH OL, OH, PT Slope Form Class Convex Planar Concave Planform Curvature1 > -0.1 -0.1 - -0.4 < -0.4 1 the unit of measure in which planform curvature is expressed is 1 over 100 units hazard ratings shift to one class more limiting on slopes >20% 1 103 Table 5. Debris slide hazard matrix. Slope Form Class Convex Planar Concave Slope Gradient Class Moderate Steep Low Lower Hazard Lower Hazard Moderate Hazard Lower Hazard Lower Hazard Higher Hazard Very Steep Lower Hazard Moderate Hazard Higher Hazard Moderate Hazard Higher Hazard Higher Hazard Table 6. Classification scheme used to allocate harvest systems. System Helicopter Cable Track Skidder Wheeled Skidder Equipment Operability (Slope Gradient) Min Max (%) (%) 0 150 15 150 Hazard Tolerance Soil Erosion Higher Higher Soil Compaction Higher Higher 0 45 Lower Moderate 0 30 Lower Moderate Debris Slide Higher Higher Moderate Lower Table 7. Area in hectares by relative hazard category for the Fishburn Forest. Relative Hazard Lower Moderate Higher Soil Erosion 223.2 211.7 69.5 Soil Compaction 159.8 343.5 1.1 Table 8. Harvest system allocation for the Fishburn Forest. Harvest System Wheeled Skidder Track Skidder Cable Helicopter Area (ha) 223.1 173.7 104.4 3.2 104 Debris Slide 436.0 53.8 14.6 HARVEST SYSTEM ALLOCATION rithms used to compute the relative ratings for the different hazards and harvesting system allocations are coarse representations of complex systems. For example, debris slides and excessive erosion are often initiated by intense or prolonged rainfall events, which are dynamic, localized features that are difficult to model spatially. Features such as canopy cover and time of year can have an impact on the all of hazards assessed. As such, care needs to be taken to avoid overstepping the intended utility of the model output. Appropriate inferences that can be drawn from the output of the Fishburn analysis include the following: Harvest system allocation is dictated primarily by slope gradient and tolerance to the aforementioned environmental hazards. Slope gradient limitations on ground-based equipment are imposed based on a combination of production, environmental, and safety reasons (Conway 1982). For the aerial systems, maximum operable slopes are imposed predominantly for the safety of forest workers. Maximum tolerable ratings for soil erosion, soil compaction, and debris slide hazards are imposed based on the potential for adverse impacts associated with the different harvesting systems. The default classification scheme used by the model is contained in Table 6. When two or more systems are deemed appropriate, the model defaults to the least expensive alternative. For the purposes of this modeling effort, the wheeled skidder system is considered the least expensive alternative, followed by the track skidder, cable, then helicopter systems. In addition to slope gradient and hazard tolerance, yarding distance and deflection are also factored into cable system allocation. While a number of different cable system configurations exist, the model assesses the suitability of a single span system with a default maximum yarding distance of approximately 450 meters. To ensure adequate loadcarrying capacity, the algorithm for cable system suitability requires that a minimum mid-span deflection of at least 5% is attainable given the shape of the terrain and a yarder tower and tailhold of 18 meters and 2 meters, respectively. 1. Efforts to mitigate soil erosion and soil compaction will have to be considered for over 50% of the forest. 2. The hazard of debris slide occurrence is low for most of the forest; however, a few locations will require detailed field investigation. 3. A majority of the forest can be harvested using ground-based systems, however, approximately 20% will most likely require the use of a cable system. The coarseness of the algorithms is a function of the model’s intended use and its reliance on datasets readily available to the public. The intended use of the model output is to supplement the planning of timber harvests at the strategic and tactical levels. The model is not intended to serve as an operational, site-specific guide for forest management activities. For example, it would be inappropriate to use the hazard and harvesting system allocation maps to delineate harvesting or site treatment boundaries without conducting detailed field analyses. With respect to data requirements, the model was designed to widely distributed datasets that were readily available to the public. As such, parameter selection is limited to variables that can be obtained from these readily available datasets. Though limited to the strategic and tactical phases, the model provides a quick first approximation of harvesting system requirements and can assist planners and managers in the prioritization of detailed hazard inspection. The value of any model, spatial or nonspatial, is often assessed through verification and validation. Verification is a subjective assessment of the internal logic used by a model, given its intended purpose (Brady and Whysong 1999). With respect to verification, the protocol and default parameter values used by the model are based primarily on published research. Given the intended use and scale of model application, the protocol, algorithms, and data used by the model are believed to be more than adequate. Validation is an objective test of model behavior and performance. Because the hazard RESULTS AND DISCUSSION The analysis on the Fishburn Forest was conducted using elevation data obtained from the Blacksburg and Radford North 10-meter USGS 7.5-minute DEMs and soils data from the Montgomery County, VA SSURGO dataset. Tables 7 and 8 contain tabular results pertaining to the relative hazard assessments and harvest system allocation, respectively. Figure 2 contains spatial output depicting soil erosion hazard, soil compaction hazard, debris slide hazard, and harvest system allocation. Even with the conservative approach taken by the model, only a small portion of the forest was assigned Kffact values indicative of greater potential erosion. Specifically, 24 hectares were assigned a Kffact > 0.35 and were subjected to the more restrictive slope gradient ranges described in the erosion hazard assessment protocol outlined in Table 1. With respect to soil compaction hazard, all but 5 hectares were observed to have higher soil strengths as dictated by their Unified Soil Group designations. However, due to the influence of slope gradient, a good portion of the higher strength soils was assigned a relative soil compaction hazard of moderate. Although the model generates relatively precise tabular and spatial information, care must be taken in the interpretation and use of output. The purpose of the model is to serve as decision support tool during the strategic and tactical phases of forest management planning, and the algo105 Figure 2. Model output depicting relative soil erosion hazard, relative soil compaction hazard, relative debris slide hazard, and harvest system allocation for the Fishburn Forest (classification schemes in black-and-white reproductions of model output are difficult to discern due to the hillshade effect used to convey topographic information). 106 assessments are qualitative (lower, moderate and higher hazard), validation will most likely take the form of sensitivity analyses, the results of which could vary significantly depending on the terrain characteristics of the study area. The flexibility built into the design of the model with respect to the ability to manipulate key parameter values and select datasets of varying scale and resolution greatly facilitates the user’s ability to conduct sensitivity analyses. Analyses can easily be conducted to determine the sensitivity of the hazard assessments to perturbations in parameters values and to the use of datasets possessing different scales and resolutions. Similar types of sensitivity analyses could be conducted on the harvesting system allocation component of the model. Sensing and the American Congress on Surveying and Mapping. San Francisco, CA. Brady, W.W. and G.L. Whysong. 1999. Modeling. P. 293-324 in GIS solutions in natural resource management: Balancing the technical-political equation. S. Morain (ed.). OnWord Press, Santa Fe, NM. Cernica, J.N. 1995. Geotechnical engineering: Soil mechanics. John Wiley & Sons, Inc, New York, NY. 453 p. Conway, S. 1982. Logging practices: Principles of timber harvesting systems. Miller Freeman Publications, Inc., San Francisco, CA. 416 p. Davis, C.J. and T.W. Reisinger. 1990. Evaluating terrain for harvesting equipment selection. Journal of Forest Engineering 2(1): 9-16. CONCLUSIONS Dietrich, W.E., C.J. Wilson and S.L. Reneau. 1986. Hollows, colluvium, and landslides in soil-mantled landscapes. P. 361-388 in Hillslope Processes. A. D. Abrahams (ed.). Allen and Unwin, Boston, MA. Information technologies such as Geographic Information Systems (GIS) have long been used to assist natural resources planning and similar models to the one presented herein have been developed (Bobbe 1987, Davis and Reisinger 1990). Existing models, however, do not specifically address the hazards associated with steep terrain, and their use is often limited by the need for specialized data. Acquiring the necessary spatial data is one of the biggest limitations in the modeling of complex natural phenomena. Database development typically constitutes a major expenditure with respect to both time and financial resources, often consuming up to 80% of a project’s budget (Antenucci, et al. 1991, Green 1999). GIS models designed to utilize publicly available spatial data, such as the steep terrain harvesting risk assessment model presented in this research, free up resources that would otherwise be needed for data acquisition and are accessible to a wide audience of users. Gibson, H.E. and C.J. Biller. 1975. A second look at cable logging in the Appalachians. Journal of Forestry 73(10): 649653. Green, K. 1999. Development of the spatial domain in resource management. P. 5-15 in GIS solutions in natural resource management: Balancing the technical-political equation. S. Morain (ed.). OnWord Press, Santa Fe, NM. Krag, R., K. Higginbotham and R. Rothwell. 1986. Logging and soil disturbance in southeast British Columbia. Canadian Journal of Forest Research 16(6): 1345-1354. Manwaring, J.C. and G.A. Conway. 2001. Helicopter logging in Alaska – surveillance and prevention of crashes. P. 9-20 in Proc. of the International Mountain Logging and 11th Pacific Northwest Skyline Symposium. P. Schiess and F. Krogstad (eds.). Seattle, WA. LITERATURE CITED Adams, P.W. 1998. Soil Compaction on Woodland Properties. Oregon State University Extension Service. 8p. Martin, C.W. 1988. Soil disturbance by logging in New England—review and management recommendations. Northern Journal of Applied Forestry 5(1): 30-34. Antenucci, J.C., K. Brown, P. Croswell, M. Kevany and H. Archer. 1991. Geographic Information Systems: A guide to the technology. Van Nostrand Reinhold, New York, NY. 301 p. Miller, J.H. and D.L. Sirois. 1986. Soil disturbance by skyline yarding vs. skidding in a loamy hill forest. Soil Science Society of America Journal 50(6): 1579-1583. Bobbe, T.J. 1987. An application of a geographic information system to the timber sale planning process on the Tongass National Forest - Ketchikan area. P. 554-562 in Proc. of the GIS ’87 - San Francisco: Second International Conference, Exhibits and Workshops on Geographic Information Systems. American Society for Photogrammetry and Remote Natural Resources Conservation Service. 1995. State Soil Geographic (STATSGO) Data Base Data Use Information. Natural Resources Conservation Service. 1998. National Forestry Manual. 107 Rice, R.M., J.S. Rothacher and W.F. Megahan. 1972. Erosional consequences of timber harvesting: an appraisal. P. 321-329 in Proc. of the Watersheds in Transition Symposium. American Water Resources Association, Urbana, IL. Toy, T.J., G.R. Foster and K.G. Renard. 2002. Soil erosion: Processes, prediction, measurement and control. John Wiley and Sons, Inc., New York, NY. 338 p. U.S. Geological Survey. 1987. Digital Elevation Models Data User’s Guide 5. U.S. Department of the Interior, USGS. 38 p Schnepf, C. 2002. Prevent forest soil compaction - designate skid trails. UI Extension Forestry Information Series, Forest Management No. 8. 1 p. Virginia Department of Forestry. 2002. Virginia’s Forestry Best Management Practices for Water Quality. 216 p. Shaw, S.C. and D.H. Johnson. 1995. Slope morphology model derived from digital elevation data. in Proc. of the Northwest ARC/INFO Users Conference. Coeur d’ Alene, ID. Washington State Forest Practices Board. 2000. Washington Forest Practices Board Manual (Section 16) - Guidelines for evaluating potentially unstable slopes and landforms. Washington State Department of Natural Resources, Forest Practices Division, Olympia, WA. Sloan, H. 2001. Appalachian Hardwood Logging Systems; Managing Change for Effective BMP Implementation. in Proc. of the 24th Annual Meeting of the Council on Forest Engineering. J. Wang, M. Wolford and J. McNeel (eds). Snowshoe, WV. Wu, W. and R.C. Sidle. 1995. A distributed slope stability model for steep forested basins. Water Resources Research 31(8): 2097-2110. 108 Use of the Analytic Hierarchy Process to Compare Disparate Data and Set Priorities ELIZABETH COULTER AND DR. JOHN SESSIONS Abstract: Given the promise of more and better data, both physical and biological, the question of how to use it for decision making still remains. The Analytic Hierarchy Process (AHP) may be useful. AHP is a technique that is used to compare alternatives based upon a number of criteria that may not be directly comparable. The AHP involves structuring problems as a hierarchy, completing pairwise comparisons between attributes to determine user preferences, and using these comparisons to calculate weightings for each of the individual attributes. The major strength of the AHP is that it allows attributes measured on different scales (such as length, area, and categorical variables) to be compared. The utility of AHP will be demonstrated using one or more examples. INTRODUCTION ANALYTIC HIERARCHY PROCESS As the ability to gather more and better data is increased, the challenge becomes one of determining how to use this information to make better, more informed decisions. Often this data is physical and biological, quantitative and qualitative, and measured on many different scales. Additionally, in many cases science has not determined quantifiable relationships between cause and effect, leaving the decisions up to professional judgment. Multi-Criteria Decision Analysis (MCDM) is a field of theory that deals with analyzing problems based on a number of criteria or on a number of attributes (also called Multi-Attribute Utility Theory, or MAUT). Many MDCM techniques exist, such as goal programming and combinatorial optimization. However, these techniques have several drawbacks. For example, the weights placed on individual attributes being compared, such as acres harvested, tons of sediment, and dollars of net present value, are required to serve two purposes: first, to make the variables measured on different scales comparables, and second to adjust the relative importance to the problem of each variable. An alternative MCDM method called the Analytic Hierarchy Process, or AHP, is presented here. AHP is not a new technique, but it is a model that has not been widely applied in natural resource situations and deserves a broader audience as it is well suited to many problems faced in forestry and natural resource management. This paper will discuss AHP methodology in general and give examples of its use in natural resource management situations. The Analytic Hierarchy Process (AHP) was originally developed in the mid-1970’s by Thomas L. Saaty (Saaty 1977) and has been used widely in many fields such as business and operations research. The AHP involves the following three basic steps: • Structuring problems as a hierarchy; • Completion of pairwise comparisons between attributes to determine the user’s preferences; and • Weighting of attributes and calculation of priority. Structuring Problems as a Hierarchy AHP requires that problems be structured hierarchically so that the overall goal is represented at the top and the individual alternatives to be compared form the base of the hierarchy. At the center of the hierarchy is one or more layers containing the attributes alternatives will be compared on. For example, consider a problem where traffic is to be routed through a network based on minimizing total transportation costs, represented by monetary costs (distance), and environmental costs related to unstable roads (various slope stability factors). This problem could be represented with the hierarchy in Figure 1. 109 Minimize total transportation cost (including environmental costs) Path Length (m) Path 1 Upslope Contributing Area (m2) Path 2 Mean Hillslope Angle (deg.) Path 3 Path 4 Surface Type (Gravel, Paved, Dirt) Path 5 Path 6 Figure 1: The example problem presented as a hierarchy. Pairwise Comparisons Weighting of Attributes and Calculation of Priority Pairwise comparisons are made between each of the attributes to be compared based on the contribution of each attribute to the overall goal (the highest level of the hierarchy). Comparisons use the one to nine scale shown in Figure 2, termed the fundamental scale, where one signifies equal importance between the attributes and nine is used when one attribute is strongly more important than the other attribute. Reciprocals are used to express the strength of the weaker of the two attributes. For example, if A is 7 times more important than B, then B is 1/7 as important as A. The results from pairwise comparisons create a positive reciprocal matrix as shown in Figure 3. AHP does not require that the user be rational or consistent in completing these pairwise comparisons. The original version of AHP required that the user also complete pairwise comparisons for each attribute of each alternative being compared (Saaty 1980), termed relative scaling. Relative scaling is an acceptable method if fewer than seven alternatives are being compared. If the problem becomes larger than this the comparisons between all possible alternatives become unwieldy. Another approach is to use an absolute scaling method where each alternative is scaled against an “ideal” alternative, often chosen as the largest alternative available. Depending on the specific nature of the problem, this relative value can be assigned linearly as a proportion of the largest value present or based on some other non-linear function (Weich 1995). Various methods for calculating attribute weights from the pairwise comparison matrix have been proposed. Saaty (1977, 1980) calculates the principle right eigenvector of this positive reciprocal matrix while others (Lootsma 1996) have used the normalized geometric mean of the rows of the priority matrix. The method involving geometric means is a simpler method and has not been conclusively shown to be inferior to the eigenvector method. For our example, the priority vector would contain the following weights: Distance 0.5876, Upslope Contributing Area 0.2230, Mean Hillslope Angle 0.1591, and Surface Type 0.0402. Distance received the highest priority value, meaning the user in this case feels that distance is the attribute that contributes most to minimizing overall transportation costs. The calculation of priority of one route as compared to another route (Pn) would be the product of the attribute weight and the relative attribute value summed across all attributes for each route. This can be written for the example here as: Pn = 0.5876dn + 0.2230an + 0.1591sn + 0.0402tn Where: Pn = Relative priority of route n dn = relative distance for route n an = relative mean hillslope angle for route n tn = relative surface type value for route n 110 Intensity of Importance Definition Explanation 1 Equal importance Two activities contribute equally to the objective 2 Weak 3 Moderate importance 4 Moderate plus 5 Strong importance 6 Strong plus 7 Very strong or demonstrated importance 8 Very, very strong 9 Extreme Importance Experience and judgment slightly favor one activity over another Experience and judgment strongly favor one activity over another An activity Is favored very strongly over another; its dominance demonstrated in practice The evidence favoring one activity over another is of the highest possible order of affirmation Figure 2: The Fundamental Scale used for pairwise comparisons in AHP. Distance Distance Upslope Area Slope Angle Surface Type 1 1/5 1/3 1/9 Upslope Area 5 1 1/3 1/5 Slope Angle 3 3 1 1/7 Surface Type 9 5 7 1 Figure 3: Matrix of pairwise comparisons used in the example problem. Route Distance (m) 1 2 3 4 5 6 600 1200 800 1500 2000 900 Upslope Contributing Area (m2) 1,000,000 3,000,000 2,500,000 50,000 100,000 150,000 Mean Hillslope Angle (degrees) Surface Type 45 20 15 50 10 75 Paved Gravel Gravel Dirt Dirt Gravel Figure 4: Example route data. 111 Dirt Gravel Paved Dirt Gravel Paved 1 1/3 1/9 3 1 1/7 9 7 1 Normalized Geometric Mean 0.66 0.29 0.05 Attribute Value 1.00 0.44 0.08 Figure 5: Using AHP to determine Surface Type relative attribute values. Route Distance (m) Upslope Contributing Area (m2) 1 2 3 4 5 6 0.30 0.60 0.40 0.75 1.00 0.45 0.33 1.00 0.83 0.02 0.03 0.05 Mean Hillslope Angle (degrees) 0.60 0.27 0.20 0.67 0.13 1.00 Surface Type Preference Value Rank 0.08 0.44 0.44 1.00 1.00 0.44 0.35 0.63 0.47 0.58 0.65 0.45 1 5 3 4 6 2 Figure 6: Example results using AHP to prioritize transportation routes based on minimizing total transportation costs, both economic and environmental. Example attribute values and the attribute weights, the larger the contribution to the overall cost of transportation. Therefore, lower values, both of attribute values and later, overall priority values, indicate the least costly, or more preferred options. Problems can be worked in either “direction”, but care must be taken to be consistent throughout the problem formulation, implementation, and interpretation. Figure 6 shows relative attribute values for the example problem, the total priority values, and the relative ranked preference for each route. Let us assume we have the six routes shown in Figure 4 to compare. Because of the widely varying scales used, it would be difficult to use these values as they are now. Instead, each of these values needs to be reduced to a relative value. For this example, we will assign attribute values for each route as a percentage of the largest value for each attribute overall. This will produce values between zero and one, with larger values being the more “expensive” values, or those that lead to a greater increase in the total transportation cost. For the Surface Type attribute a different approach must be taken. Here, we can either assign values between 0 and 1 for each surface type or we can use pairwise comparisons between the surface types to determine weighting values. Figure 5 shows a matrix of pairwise comparisons for Surface Type. The last two columns of Figure 5 give the normalized geometric mean of the rows as well as the value that will be used for attribute values. Either of these values could be used, however, to be consistent, all other attribute values are decimal percentages of the largest attribute value present and therefore this is the value that should be used. It is important to remember the “direction” of the problem. For this example, the higher the value, both for the relative CONCLUSION The major strength of the AHP is that it allows attributes measured on different scales to be compared (Saaty 1980). This is especially important to this problem where the comparison of values such as meters of distance, square meters of upslope contributing area, degrees of hillslope, and a categorical surface type must be undertaken in order to arrive at an overall priority for each proposed route. AHP also forces the user to make explicit values used in decision making (Keeney 1988) and is useful in situations where the quantification of cause and effect relationships is left up to professional judgment. 112 This paper has presented a brief overview of AHP methodology and an example demonstrating the technique’s usefulness in comparing alternatives with multiple criteria measured on different scales in a case where it is left up to professional judgment to determine the relative contribution of each attribute towards the objective. criteria in the Multiplicative AHP and SMART. European Journal of Operational Research. 94:467-476. Saaty, T.L. 1977. A scaling method for priorities in hierarchical structures. Journal of Mathematical Psychology. 15:234281. Saaty, T.L. 1980. The Analytic Hierarchy Process: Planning, Priority Setting, Resource Allocation. McGraw-Hill, New York. 287 p. LITERATURE CITED Keeney, R.L. 1988. Value-driven expert systems for decision support. Decision Support Systems. 4:405-412. Weich, B.G. 1995. Analytic hierarchy process using Microsoft Excel. In Engineering. California State University, Northridge, p. 24. Lootsma, F. A. 1996. A model for the relative importance of the 113 114 Use of Spatially Explicit Inventory Data for Forest Level Decisions BRUCE C. LARSON AND ALEXANDER EVANS Abstract: Society is demanding that forest managers produce more spatially complex forests even at the scale of within forest stand. Harvest techniques for complex even-age management have taken on a variety of names such as partial cutting, green tree retention, and partial overstory removal. Traditional growth models relying on stand averaging techniques are often imprecise estimators of timber growth in these situations because many growth processes are non-linear and would require a uniform pattern of leave trees. Likewise forest and landscape descriptions are less reliable predictors of non-timber values if the forest is viewed as a pattern of discrete polygons (stands) instead of a smaller grain-sized mosaic of different sized trees, especially in mixed species stands. Most inventory systems now include a GPS location for each plot. These data can be used in a raster-based GIS system to give a finer grain analysis of the forest. Information from each plot can be interpolated to give a smooth interpretation of variable values across the forest. Almost any variable or combination of variables can be used. Examples are basal area or volume either in total or for different species. Crown cover and downed wood volumes are example of other, non-timber values, that can be depicted. Most of our existing forest management quantitative tools were designed when desktop computational power was much more limiting. New tools will have to be written such that forests and even stands can be depicted in a much more precise manner. High precision data management and analysis will be the result of shifting computational paradigms. Much less averaging and use of representative stands will result. It is doubtful that new tools will replace the need for existing models; several models will be used in concert to make decisions. Early indications are, as to be expected, if stand age is the driving variable for all others and the primary disturbance in the forest is clearcutting, then traditional stand polygons are a more accurate representation of the forest. However, in many other situations, stand averaged polygons will obscure the variation that forest managers are trying to create. 115 116 Elements of Hierarchical Planning in Forestry: A Focus on the Mathematical Model S. D. PITTMAN Abstract: The hierarchical approach to forest management has been advanced as an integrated method for constructing large-scale forest plans. While the planning process functions within a hierarchical construct, the mathematical models describing the plan also have multi-level structure. Two models which consistently appear in multi-level planning, are mathematical programs with block angular structure, also referred to as a hierarchical production planning problems, and the hierarchical optimization problem. Depending on the conceptual model of the planning venture, each of these mathematical models is a possible realization. The implication of these modeling formulations is discussed within the context of the hierarchical approach to forest planning. 117 118 Update Strategies for Stand-Based Forest Inventories STEPHEN E. FAIRWEATHER Abstract: Stand-based forest inventories are typically kept current with a combination of cruising, growth modeling, and adjustments to represent harvest activity. At any point in time the inventory will have some stands with recent cruise data, some stands which have never been cruised but carry estimates for the stratum they belong to, and some stands which were cruised some time ago and have been grown each year using a growth model. There are many strategies for keeping the inventory up to date. For example, the entire ownership may be cruised at one point in time, and then grown and depleted annually until the ownership is cruised again. Or, cruising may be an ongoing annual activity, such that a different portion of the ownership is cruised each year. Each strategy has advantages and disadvantages in terms of costs, how accurately it will portray the true inventory at any point in time, how accurately individual stand volumes will be portrayed, and the degree to which the current inventory estimate will change from one year to the next simply as an artifact of the updating system. This paper defines the problem and presents a simulation model for evaluating different update strategies. The model allows the user to study the impact of update strategy and several sources of estimation error on the accuracy of the inventory estimates. DEFINITION OF THE PROBLEM ten years later? Or, perhaps it would be better to cruise some of the stands every year, such that each stand gets cruised every ten years, but not all stands are cruised at the same time. Each of these update strategies has advantages and disadvantages, and the selection of the proper strategy is not always clear. In a stand-based forest inventory system, the stand is the basic unit of inventory. As such, the sum of all the individual stand inventories at one particular point in time constitutes the inventory for the entire ownership. At any point in time the inventory estimate for any particular stand may be established in any of three ways: • • • GOALS FOR A STAND-BASED INVENTORY The stand may have an estimate based on a cruise of that stand in the current year; There are three goals for a stand-based forest inventory that will help to define criteria for evaluating alternative update strategies. The goals are: The stand may have an estimate based on a past cruise that has been grown, with a growth model, to the current year; 1. Provide an accurate estimate of the total forest inventory at any point in time. This is necessary to facilitate valuations and appraisals. The stand may have an estimate which is essentially the average for the stratum that the stand belongs to, where the average is based on the stands in the stratum which have been cruised either in the current year or in the past. 2. Provide accurate volume estimates at the stand level to support on-the-ground operations. It is particularly important for the system to provide inventory estimates that are close to removal volumes when a stand is actually harvested; the “cutout”, or the ratio of the inventory estimate to the harvest volume, should be close to 100%. If the cutout routinely runs much differently than that, the confidence of the field foresters in the inventory system will quickly erode, and their lack of support for the system will place it in jeopardy. As the forest-wide inventory is maintained over time, the question of an appropriate “update strategy” will eventually have to be considered. For example, should the strategy be to cruise every stand, every year? Or, should the strategy be to cruise all of the stands at one time, grow them ahead each year with a growth model, and then recruise all of them 119 where an error is defined as the difference between the cruise estimate and the true value of the stand inventory at that point in time. 3. Minimize the frequency and magnitude of year to year changes in the inventory estimate, both for the ownership and for individual stands, where such changes are an artifact of the estimation system being used. For example, inventory foresters recognize the possibility of two well designed and well executed cruises in subsequent years in a single stand suggesting a decrease in stand volume, when in fact the stand has been growing steadily in volume, simply due to random chance. By the same token, two cruises may suggest an increase in volume from one year to the next that is beyond reasonable expectations of growth, again due to random chance. At the larger scale, a current inventory established by cruising every stand ten years ago and growing each stand to the current point in time may show an alarming decrease in total volume when the ownership is cruised again, perhaps because the growth model was biased high, or perhaps because either cruise was not conducted carefully. Inventory foresters may understand how this could happen, but the folks in timberland accounting, who are used to thinking in terms of changes in the inventory due to growth, depletions, and changes in the land base, will be very uncomfortable with changes due to “better information”. 5. Define a bias in the growth model being used to grow stands cruised in the past to the current point in time. 6. Define a cost per cruise plot. The model can then calculate the total cost of cruising in any year, and the total discounted cost of cruising for the 11-year period. The cost of using a growth model or applying a stratum average to uncruised stands is assumed to be inconsequential relative to cruising. 7. Conduct repeated applications of the update strategy, and collect the results over all replications. The simulation model lets the user examine the impact of several sources of error in the inventory update process: • Non-homogeneous stands within an inventory stratum. As the stand-to-stand variation in volume per acre in a stratum increases, the usefulness of applying a stratum average to individual uncruised stands decreases. • Sampling error in cruise estimates for individual stands. • Growth model prediction error (bias). THE SIMULATION MODEL We have developed an easy to use simulation model in MS Excel to let the user evaluate a range of inventory update strategies. The model lets the user do the following: 1. Define a forest of 20 stands in terms of the actual (true) inventory in each stand in each year of an 11-year period, and the acres for each stand. It is helpful to think of the 20 stands as constituting a single stratum (cover type, or photo-interpreted type) in the ownership. SIMULATION RESULTS We simulated five update strategies for purposes of illustration. Our forest consisted of 20 stands ranging in size from 10 to 56 acres, averaging 32 acres. The volume in the stands averaged 3,040 units per acre, and ranged from 2,500 to 3,700. The growth rates varied from 2.8 to 4.5%, and averaged 3.9% overall. For any stand selected to be cruised, we specified an allowable error of +/- 20% at the 90% confidence level. At an assumed coefficient of variation of 50%, the resulting sample size was 18 plots per stand. We also specified that in any stand there would be no more than 1 plot in 2 acres. Therefore, cruised stands with less than 36 acres ended up with 1 plot for every 2 acres, and stands with more than 36 acres were cruised with 18 plots. For this set of simulations we assumed the growth model was “perfect”, i.e., the true growth percent for each stand was predicted without error. We assumed a cost per plot of $30, and discounted total cruising costs each year at a rate of 8%. We assumed that the costs of growth modeling and/or using stratum averages (expanding) was negligible compared to the cost of cruising. We collected results for 100 replications of each of the following update strategies: 2. Define an actual (true) annual growth rate for each of the 20 stands. 3. Describe the inventory update strategy to be evaluated. In each year of the 11-year period, the inventory estimate for a stand will be based either on a cruise of that stand, a past cruise that has been grown with a growth model, or a stratum average for the other stands which have either been cruised in that year or grown to that year from a past cruise. Figure 1 illustrates the characterization of one particular update strategy. 4. Define the number of plots that will be used to cruise any particular stand. The number of plots is calculated based on the expected CV (coefficient of variation) for volume, the allowable error, and the confidence level. In small stands, the user can specify a minimum number of acres per plot which will override the number of plots from the sample size calculation. The model uses the number of plots and the CV to randomly generate errors in the cruise estimate, 120 A. Cruise every stand, every year. This strategy required no expanding, and no growth modeling, so the only source of error would be due to sampling error in the cruising. The results of the first update strategy are shown in Figure 2. The first graph shows the average inventory estimate, over all stands, was equal to the true inventory in each year of the simulation, as would be expected if our cruising (sampling methodology) was unbiased. The graph also shows the range in inventory estimates each year. In 1999, for example, there was an estimate of the total inventory that was approximately 17% less than the true value, which might be surprising given that all the stands were cruised in that year, and every year. The second graph in Figure 2 displays how often the inventory estimate, over all stands, decreased from one year to the next. For example, in 1999, in 30 replications out of 100, the estimate of inventory was less than it was in 1998. Figures 3 illustrates the results of update strategies B and C. Cruising the entire ownership on either a 5-year or 10year interval, and using the growth model to update stand inventories between cruising, resulted in unbiased estimates with a range of errors no greater than was experienced when every stand was cruised in every year. The rate of decreasing inventory estimates was still between 20 and 30%, but at least this would only be experienced only once every 5 or 10 years. Figure 4 illustrates the results of update strategies D and E. Both strategies cruised 10% of the stands each year, such that each stand was cruised every 10 years. In strategy D, the B. Cruise every stand at one time, on a 5-year interval; use growth modeling to update the stands until they are cruised again. C. Cruise every stand at one time, on a 10-year interval; use growth modeling to update the stands until they are cruised again. D. Cruise 10% of the stands each year, such that each stand is cruised on a 10-year interval; expand the stratum average to uncruised stands each year. No growth modeling is used. E. Cruise 10% of the stands each year, such that each stand is cruised on a 10-year interval; expand the stratum average to uncruised stands in the first year only, and then grow all stands with the growth model until they are cruised again. This update strategy is illustrated in Figure 1. Stand A B C D E F G H I J K L M N O P Q R S T 1991 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 1 1 3 3 1992 2 2 2 2 2 2 2 2 2 2 1 1 2 2 2 2 2 2 2 2 1993 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 1994 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1995 2 2 2 2 2 2 1 1 2 2 2 2 2 2 2 2 2 2 2 2 1996 2 2 2 2 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1997 2 2 2 2 2 2 2 2 1 1 2 2 2 2 2 2 2 2 2 2 1998 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 2 2 2 2 1999 2 2 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2000 2 2 2 2 2 2 2 2 2 2 2 2 1 1 2 2 2 2 2 2 2001 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 1 2 2 Figure 1. Illustration of how an update strategy is defined in the simulation model. Cells denoted with a “1” indicate cruises in that stand in the given year. Cells denoted with a “2” indicate the estimate of inventory for the stand in the given year will based on growing the previous year’s estimate with a growth model. Cells denoted with a “3” indicate the estimate for the stand in that year is the average (weighted by acres) of the estimates in that year for the other stands that have either been cruised (“1”) or grown (“2”). This particular update strategy features cruising 10% of the ownership on a 10-year cycle. In the first year of the inventory, only two stands are actually cruised, and the other stands are “expanded” to, i.e., they take on the average of the two cruised stands. After the first year, all stands are either cruised or grown. 121 Total Inventory (MBF) Cruise Every Stand, Each Year No Growth Modeling, No Expansion 3500 3000 TRUE 2500 Avg Estimate 2000 Low Estimate High Estimate 1500 1000 1990 1992 1994 1996 1998 2000 2002 Year Frequency of Decreases in Estim ates of Total Inventory; Cruise Every Stand Each Year Freq/100 40 30 20 10 0 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 Year Figure 2. Results of 100 replications of update strategy A. In this strategy every stand was cruised in every year. egy. “RMSE” is the root mean squared error, defined as the square root of the sum of squared differences between the estimate of volume and the true volume for each stand, over 100 replications. A low RMSE would be preferred over a high RMSE. The graph shows no clear advantage of any particular strategy, but does show the tendency for stands with small areas to have the least accurate inventory estimates; the spikes in the RMSE values for stands I, N, and P correspond to the three smallest stands in the model forest. The average discounted cruising costs for the five strategies are shown in Table 1. Given that all the strategies appeared to be unbiased, and there was no difference between strategies with regard to the accuracy of individual stand estimates (by the year 2000), the low cost for strategies D and E might make them more attractive than A, B, or C. Strategy E also offers a low rate of decreasing estimates from year to year. It might be selected as the preferred strategy depending on the analyst’s willingness to accept the possi- average of the cruised stands was expanded to the uncruised stands in each year, resulting in high rates of decreasing inventory estimates. In strategy E, the average of the cruised stands was expanded to the uncruised stands in only the first year of the sequence; after that, each stand was grown with the growth model until it was cruised. Once a stand was cruised, its inventory was restated, and then grown from there. Strategy E appear to be unbiased, and the variability in the inventory estimates stabilizes enough by the year 2000 (i.e., 10 years into the cycle) to be as precise as any of the other strategies. Strategy E, however, displays a large advantage over strategy D in terms of minimizing the frequency of decreases in the inventory estimate from one year to the next. Figure 5 compares the accuracy of the update strategies on an individual stand basis in the year 2000. The year 2000 was selected as the benchmark year because by that time every stand has been cruised at least once in each strat122 Frequency of Decreases in Estimates of Total Inventory; Cruise Every Stand, 5-Year Interval 3500 40 3000 TRUE 2500 Avg Estimate 2000 Low Estimate 1500 High Estimate 1000 1990 Freq/100 Total Inventory (MBF) Cruise Every Stand, 5-Year Interval Use Growth Modeling 30 20 10 0 1992 1994 1996 1998 2000 2002 1992 1993 1994 1995 Year 1997 1998 1999 2000 2001 Year Cruise Every Stand, 10-Year Interval Use Growth Modeling Frequency of Decreases in Estimates of Total Inventory; Cruise Every Stand, 10-Year Interval 3500 40 3000 TRUE 2500 Avg Estimate 2000 Low Estimate High Estimate 1500 1000 1990 Freq/100 Total Inventory (MBF) 1996 30 20 10 0 1992 1994 1996 1998 2000 2002 1992 Year 1993 1994 1995 1996 1997 Year Figure 3. Results of 100 replications of update strategies B and C. Figure 4. Results of 100 replications of update strategies D and E. 123 1998 1999 2000 2001 RMSE in 2000 by Stand and Update Strategy 1100.0 1000.0 RMSE (Volume) 900.0 A 800.0 B 700.0 C D 600.0 E 500.0 400.0 300.0 A B C D E F G H I J K L M N O P Q R S T Stand Figure 5. Comparison of update strategies with regard to accuracy of individual stand inventory estimates in the year 2000. Table 1. Average discounted cruising costs by update strategy. Strategy Description Cost A Cruise every stand, every year $64,765 B Cruise every stand on 5-year cycle $18,008 C Cruise every stand on 10-year cycle $12,291 D Cruise 10% of the stands each year; expand each year $6,545 E Cruise 10% of the stands each year; expand in the first year only $6,545 124 bility of larger errors in the overall inventory estimate while the system is being established, i.e., in the early years of the process. A simple variation on strategy E that might mitigate the range in inventory estimates while the system is being established would be to cruise the larger stands in the beginning years of the program. That is, while 10% of the stands may be selected for cruising, they may account for 20% or 30% of the area. This idea, and the results of the simulation, are shown in Figure 6. In this particular case the results of the simulation in Figure 6 indicate that the largest stands tended to have more volume per acre, on average, than the rest of the stands in the stratum, resulting in a biased inventory update strategy until most of the stands have been cruised. Stand A B C D E F G H I J K L M N O P Q R S T 1991 3 3 1 3 1 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 1992 2 2 2 2 2 2 2 2 2 2 1 2 2 2 1 2 2 2 2 2 1993 2 2 2 2 2 2 2 2 2 1 2 2 2 2 2 2 2 1 2 2 1994 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 The lower bound on the inventory estimate was improved somewhat over update strategy E, and the average discounted cruising cost for the new strategy only increased to $6,852. But, the slight bias in the strategy underscores the importance of avoiding any relationship between stand size and volume per acre when stands are being assigned to strata. CONCLUSION This simple simulation model will be quite helpful in exploring different update strategies and the impacts of errors attributable to cruise design, stratification, variability within and between stands, and growth modeling. Future applications will use the growth modeling error control to understand the importance of calibrating the growth model to be unbiased, i.e., to have 0% error. 1995 2 2 2 2 2 2 1 1 2 2 2 2 2 2 2 2 2 2 2 2 1996 2 2 2 1 2 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1997 2 2 2 2 2 2 2 2 1 2 2 2 2 2 2 2 2 2 1 2 Cruise 10% on 10-Year Interval, Use Growth Modeling, Expand in First Year Only, Cruise Largest Stands First 1998 2 2 2 2 2 2 2 2 2 2 2 1 2 2 2 1 2 2 2 2 1999 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 1 2 2 1 2000 2 2 2 2 2 2 2 2 2 2 2 2 1 1 2 2 2 2 2 2 2001 2 2 1 2 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 Frequency of Decreases in Estimates of Total Inventory; Cruise 10% on 10-Year Interval, Expand in First Year Only, Cruise Largest Stands First 40 3000 TRUE 2500 Freq/100 Total Inventory (MBF) 3500 Avg Estimate Low Estimate 2000 High Estimate 20 10 1500 1000 1990 30 0 1992 1994 1996 1998 2000 1992 2002 1993 1994 1995 1996 1997 1998 1999 2000 2001 Year Year Figure 6. Simulation results for an update strategy similar to strategy E, but concentrating the cruising on the largest stands in the stratum in the early years. Stands C, E, J, K, O, and R are the largest stands (by area) in the stratum. 125 126 A New Precision Forest Road Design and Visualization Tool: PEGGER LUKE ROGERS AND PETER SCHIESS Abstract: By evaluating alternative routes in the office using a pegging routine, days or even weeks can be saved of valuable field time and ultimately, a better design can emerge. Initial road design in forested landscapes often includes pegging roads on large-scale contour maps with dividers and an engineers scale. An automated GIS based road-pegging tool (PEGGER) was developed to assist in initial road planning by automating the road pegging process. PEGGER is an extension for the commonly available GIS software Arcview®. PEGGER imports topography as digital contours. The user identifies the origin of the new road, clicks in the direction they want to go and PEGGER automatically pegs in road at a specified grade. Through the use of PEGGER, many alternatives can be quickly analyzed for alignment, slope stability, grades and construction cost using standard GIS functionality. The resulting cuts and fills are then displayed in ROADVIEW, a road visualization package for Arcview®. This paper looks at the algorithm used, evaluates it’s usefulness in an operations planning environment and suggests additional methods which might be incorporated into PEGGER to further assist the forest engineer. INTRODUCTION tional road design techniques into the GIS. With the availability of free 10-meter digital elevation data for the United States and the continually decreasing cost of LIDAR data it is possible to extend the road pegging technique to include a more detailed analysis. A computer program is presented that automates initial forest road location through the use of a Geographic Information System and digital terrain data. Using PEGGER, forest planners can quickly analyze many road location alternatives and, by taking advantage of standard GIS functionality, evaluate environmental and economic opportunities. EXISTING MODELS While many road design packages exist (RoadEng, AutoCAD, F.L.R.D.S…) only one has given the user the ability to quickly look at alternative road locations at varying scales, ROUTES (Reutebuch, Stephen E. 1988). Traditional road design software relies on survey data collected in the field to generate terrain models and very detailed engineered road location and construction plans. Others have taken a more holistic approach and looked at optimization of road locations for a particular set of topographical, environmental or economical constraints (Xu 1996, Thompson 1988, Wijngaard and Reinders 1985, Cha, Nako and Watahiki 1991). All these programs have relied on a high degree of training on the part of the user and few of the non-commercial packages have matured into an easy to use software package. ROUTES was developed to automate the road pegging process. Using a large-scale contour map (1in = 400ft) and a digitizer, the user could digitize the contours and use the BACKGROUND Traditional methods for designing a forest road system consisted largely of aerial photo interpretation and field reconnaissance. More recently, forest engineers have used large-scale contour maps to select preliminary routes with dividers, a process known as route projection or “pegging”. According to Pearce (1960), “Route projection is the laying out of a route for a road on a topographic map of aerial photo. The route defines the narrow strip of land within which the field preliminary survey is made.” This trial and error method of initial paper based road location has proven itself as a cost effective method for preliminary design and analysis by avoiding intensive field investigations. With the overwhelming popularity of Geographic Information Systems (GIS) in natural resource management it is appropriate to explore opportunities to integrate tradi- 127 digitizer puck to locate the road. While the user interface was primitive consisting of high and low pitch beeps from the digitizer puck to signal that the user was “on-grade”, the program worked well and kept track of such things as grade, road length and stationing. ROUTES reliance on a digitizer, it’s HP 9000 code base and the general lack of a graphical user interface (GUI) left the program without many users. tured decision making tasks.” It is with the intention of providing an initial decision support system that PEGGER was developed. PEGGER is an Arcview® GIS extension that automates the route projection (“road pegging”) process for use by engineers and forest planners. PEGGER imports topography as digital contours much like using a paper contour map. Standard tools available within Arcview GIS allow the user to import the contours from Shapefiles, ESRI coverages, AutoCAD dwg and dxf, and Microstation dgn files. In addition to importing data as digital contours, users can use the Arcview Spatial Analyst extension or other publicly available tools to convert USGS digital elevation models to contours. One of the goals of the PEGGER project was to make the program as usable as possible for as many people as practical. One of the problems with technology is training users to use the software. Forestry professionals responsible for fieldwork have been slow to adopt new technology into their work largely due to the complexity of the software and the THE PEGGER PROGRAM With the growing availability of LIDAR and IFSAR data, locating roads in the office is becoming a more realistic and practical exercise. Within the GIS framework many tools exist to locate geographic features, examine spatial relationships among natural elements and act as a foundation for a decision support system. Watson and Hill (1983) define a decision support system as an “interactive system that provides the user with easy access to decision models and data in order to support semi structured and unstruc- Figure 1 - The simple PEGGER interface in Arcview GIS. 128 can attach the grade attributes to the route segments, merge the segments into one long road or spline to smooth sharp corners (much like a finalized design). time commitment of training. The PEGGER program was designed to avoid these common pitfalls, requiring no training, minimal setup time and a simplified user interface. Included with the software are a detailed help file and complete tutorial. Once digital contours have been imported into Arcview the user must supply a few parameters, the road theme they would like to edit, the contour theme they would like to use as well as confirm the detected contour interval. In addition to the contour and road themes the user can have any number of other layers available in the GIS such as soils, slope classes, streams, wetlands, unstable slopes and property lines. The next step is to locate the desired beginning and/or endpoints of the new road given operational parameters. Using standard tools available in the GIS (ruler and identify) the user can estimate the necessary grade for the road. To start a road the user shift-clicks on the location where they wish to begin and enters the desired grade. To “peg” the road the user only has to click in the general direction they wish to go in order to project the route into the GIS. Successive clicks peg in additional segments of road from contour to contour as fast as the user can press the mouse buttons. Grade changes can be accomplished by using the Roads pull-down menu or by right clicking the mouse and selecting Increase or Decrease Grade. If the road fails to reach the desired end point, the previously pegged segments can be quickly deleted and a new grade can be tried. This method of trial and error that used to mean changing the divider spacing and erasing undesirable segments from the map can now be accomplished in the GIS in a fraction of the time. LIMITATIONS The PEGGER program relies on digital topographic information to identify potential road locations. To be of value to the forest professional, the topographic information must accurately represent the actual ground conditions. Steve Reutebuch noted about ROUTES that “the accuracy of the 30-meter (USGS) DEM’s available at the time were insufficient for accurate route projection.” With the availability of 10-meter digital elevation data and the current popularity of LIDAR data, route projection has become more feasible but discrepancy between the data and actual field conditions should be expected. The PEGGER program is a tool for quickly identifying possible route location alternatives given grades specified by the user. The tool does not evaluate additional environmental and economic constraints that must be considered by the forest professional such as soil types, hydrology, property lines and slope classes. The GIS provides a framework where these analyses can be implemented but it is outside the scope of the PEGGER program. NEXT STEPS In addition to providing quick alternative location analysis, PEGGER should be extended to include some additional functionality. With greater availability of high resolution digital elevation data it will be possible to identify a route location or P-Line (preliminary location line) using PEGGER and then “survey” the surrounding area for export into a road design package like ROADENG or AutoCAD. This digital survey within the GIS can be used ANALYTICAL DESCRIPTION PEGGER works by identifying contour lines that meet a specific set of criteria. Every projected route segment must begin and end on a contour line. To project a segment the user enters a desired grade and PEGGER looks for a point on an adjacent contour line at a distance computed by: d = ci / (g / 100) where d = the distance, ci = the contour interval, and g = the desired grade. NOTE: For pegging on paper maps, the distance would need to be multiplied by the map scale (ie: 1/4800) to get the appropriate divider width. If a point is found, a new route segment is created in the GIS. If a point is not found, the user is notified that the desired grade is not feasible and potential solutions are proposed. Unlike ROUTES, which allowed for a grade tolerance (+/- some tol), PEGGER gives an exact solution in the GIS. After a desirable route location has been found the user Figure 2 - ROADVIEW visualization of a route located with PEGGER. 129 to generate the topographic information and field notes necessary to do a complete design in the road design package. The final L-Line (location line) and slope staking notes can be generated using the GIS and the road design package for use in the field by the forest professional. Complementing PEGGER is a companion program ROADVIEW that takes the preliminary route location generated by PEGGER and creates a 3-dimensional model of the road’s cuts, fills and running surface. Using the 3-D model and a visualization program such as EnVision, professionals can look at the road as it might be constructed and effectively communicate with non-forest professionals regarding scenic and environmental impacts. tutorial, a typical user can be locating roads in a few minutes on their own PC taking full advantage of forest technology. LITERATURE CITED Pearce, J. Kenneth. 1960. Forest engineering handbook. Portland, OR: U.S. Department of the Interior, Bureau of Land Management. 220p. Cha DS, Nako H, Watahiki K. 1991. A computerized arrangement of forest roads using a digital terrain model. Journal of the Faculty of Agriculture Kyushu University. 36(12):131-142. CONCLUSION Reutebuch, SE. 1988. ROUTES: A Computer Program for Preliminary Route Location. Pacific Northwest Research Station: U.S. Department of Agriculture, Forest Service. General Technical Report PNW-GTR-216. 18p. While route location has been used by forest professionals for many years and computerized in the 1980’s with the introduction of ROUTES, it has never become a widely used technology to evaluate initial road locations. With PEGGER, the forest planner can quickly evaluate route locations within a GIS framework, giving the planner access to additional GIS functionality. PEGGER was designed with simplicity and minimal investment cost as primary objectives. Through the use of a carefully designed user interface and extensive Watson, H. J. and M. M. Hill. 1983. Decision support systems or what didn’t happen with MIS. Interface. 13(5):81-88. Xu, Shenglin. 1996. Preliminary planning of forest roads using ARC GRID. Corvallis, OR: Oregon State University, Department of Forest Engineering. 112p. 130 Harvest Scheduling with Aggregation Adjacent Constraints: A Threshold Acceptance Approach. HAMISH MARSHALL, KEVIN BOSTON, JOHN SESSIONS Abstract: Three different forest management planning unit sizes were used to compare the results from a tactical planning model that included a maximum opening size constraint with aggregation and even-flow goals. The smallest size had a 22acre average size, the second had an average size of 41-acres, while the largest unit had an average size of 59-acres. There was a direct correlation between discounted net revenue and unit size as the smallest unit definition produced $45 per acre more than the second set of units and $225 more than the largest unit size. These results suggest that planners use the latest technology when defining individual settings and managing their unit sizes as one method to improve the financial performance of their assets. INTRODUCTION form of the adjacency constraint limits the maximum opening size that can be created. This constraint is found in many forest practices rules including those of Sweden, Canada, and various western US states (Boston and Bettinger 2001, Dahlin and Sallnas 1993). The Sustainable Forestry Initiative (SFI), a voluntary certification scheme that has been adopted by much of US forest products industry, restricts the average opening size to less than 120 acres (AFAPA 2002) Digital surveying equipment, global positioning systems (GPS), and geographical information systems (GIS) technologies allows for smaller settings to be accurately defined and used in forest planning; however, many forest planners restrict the solution space of the tactical scheduling model by aggregating settings into larger units prior to solving the tactical problems. This paper explores the impact of planning unit size on the quality of the solutions produced by the tactical plan. Three degrees of pre aggregation were used (Table 1) to describe the same 4450-acre planning area with the identical yields. The three data sets were used in a tactical planning model that had the goal to maximize the discounted net revenue subject to an area-restriction green-up constraint. The maximum opening allowed is 120 acres as specified in the Oregon Forest Practices Rules (Oregon Department of Forestry 2003). One potential risk of reducing the planning unit size is that the model will widely disperse each period’s harvest over the entire landscape, hence increasing harvesting and transportation costs. To reduce this possibility, an additional goal was added to the model to aggregate original planning settings into larger planning units. The objective of the aggregation was to group settings on opposite sides of valleys into planning units that will be logged in the same year to improve the logging efficiency. Stumps The strategic planner’s goal is to develop a plan that will allow the firm to compete effectively (Porter 1986). Tactical planning aligns the operations to implement that strategy. Forestry planning problems, especially considering the long rotations, are some of the more difficult business planning problems because of economic, biological and operational uncertainty in the data used. As the forest products industry is one of the most capital-intensive industries in the world, a detailed planning and scheduling system is required for competitive share–holder returns (Propper De Callejon et al. 1998). Strategic plans have traditionally assumed the data is continuous and linear; thus allowing these problems to be solved using linear programming algorithms. Commercially available products such as FOLPI (Manley et al. 1991), Magis (MAGIS 2003) and WOODSTOCK (REMSOFT 2003) have been available for approximately 20 years to solve the long-term strategic planning problems. Initially, the solutions from these strategic planning solutions have been implemented using various ad hoc procedures. To improve both the financial and environmental performances, optimization routines have been developed to solve tactical scheduling planning problems that incorporate various spatially explicit constraints. These constraints have included various forms of green-up restrictions and unit-fixed cost tactical planning problems. Unfortunately, these new tactical planning models quickly exceeded the capacity of commercial solvers and have resulted in the employment of a variety of heuristics to solve these problems. A common spatial component found in tactical forest planning involves the harvest of adjacent planning units. One 131 Table 1. Description of the planning data sets. Aggregation Level 1 2 3 Mean Planning Unit Size (acres) 21.7 40.6 58.9 can be used as tailhold anchors for cable logging systems with less fear of failure than if they were allowed to begin to decompose. A third even volume flow goal was added to the model to regulate volume flow. Maximum Planning Unit Size (acres) 39.1 76.8 92.1 w = penalty weight, tv = target volume (Mbf), vit = total unit volume (Mbf) Aggregation Incentive The aggregation adjacent constraints have been formulated to encourage the harvesting of adjacent planning units in the same period. This constraint has also been formulated as a goal or “soft constraint”, modeled by increasing the value of the objective function when aggregation is included in the model. The amount of the incentive was based on the sum of the proportion of the perimeter of a planning unit that it shared with other planning units harvested in the same period, divided by the number of planning units. MODEL FORMULATION Objective Function The objective function goal was to maximize the net present value of the forest over a period of 20 years plus the sum of incentives and penalty values. A discount rate of 5% was applied to revenues at the end of each one-year period. The details of this incentive and penalty values are discussed below. The objective function is formulated as: n ( ∑ ( pi X it)) (w * n Rit 20 max( ∑ ∑ i =1 t =1 (1.05 ) n t 20 −∑∑ i =1 t =1 Cit (1.05 ) t ) Maximum Opening Constraint (Oregon State Rules) The maximum opening constraint or area-restriction problem (Murray 1998) was formulated as a hard constraint that cannot be violated. The area-restriction constraint was formulated so that a neighborhood of adjacent settings harvested within 4 years has to have a combined area of less than 120 acres to not violate the Oregon State Forest Practices Rules. s (4) Ai X it + ∑ Ai X it <= 120 Even Volume Flow Penalty Maintaining continuity of supply to customers is a key component in successfully operating a forestry business. operations. We have assumed the goal is to maintain an evenflow of volume throughout the 20 year planning horizon, although we recognize that it can be misapplied to small areas where the result of an even-flow constraint can significantly reduce the harvest from an area that cannot support yearly. In this study it was formulated as a goal or “soft constraint” where the objective function was penalized for any deviation of the discounted volume from a target volume in each period. The formulation is given below; the squared deviation was multiplied by a weighting factor and subtracted from the objective function. i =1 (1.05) t ( w * ∑ (tv − ) 2 (3) ) w = penalty weight, pi = proportion of perimeter shared with units harvested the same period that are either on the other side of the valley or above or below, Xit = a 0,1 binary variable to identify where a unit has been cut. Rit = Gross Revenue, Cit = Costs, i = harvest units, t = planning period, n = number of harvest units, the Even_Flow_Penalty and Aggregation_Incentive formulations are described below. vit n (1) −( Even _ Flow _ Penalty ) + ( Aggregation _ Incentive ) n i =1 i =1 Ai = area, s = a subset of adjacency units all of which are harvested within 4 years of each other. HEURISTIC ALGORITHM: THRESHOLD ACCEPTANCE ALGORITHM Threshold acceptance (TA) was first developed by Dueck and Scheuer (1990) when they claimed that it appeared to be superior to simulated annealing (SA). The idea behind (2) 132 Randomly select candidate stand and harvest period No Does the new candidate solution form an opening less than 120 acres Yes Calculate the difference between volume from the target volume for each period raised to the power of two. a Calculate the average proportion of the shared perimeter of adjacent units harvested in the same period c Calculate the net present value for each period b Calculate Objective Function = b - wa + zc Is proposed objective function less than best ever objection function Yes Retain old solution Make current solution equal best ever solution No Is it within current threshold levels No Is proposed objective function less than current objective function No Yes Yes Accept solution repetition = repetition + 1 Current solution = proposed solution Does repetition = maximum number of repetitions ? No Yes Select next threshold level Stop and report the best solution found during search and the iterations at which it first occurred Any threshold levels left? No Yes Figure 1. Flow Diagram of the Threshold Acceptance Algorithm. 5000 4500 $10.40 $10.20 Net Present Value (million) 4000 Volume (mbf) 3500 3000 No Aggregation Aggregation 1 Aggregation 2 Aggregation 3 2500 2000 $10.00 $9.80 $9.60 $9.40 $9.20 $9.00 $8.80 $8.60 1500 $8.40 No Aggregation 1000 Aggregation 1 Aggregation 2 Aggregation 3 Model 500 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 Years Figure 2. Net Present Value for the Different Aggregation Models. Figure 3. Projected Volumes (Mbf) over the 20 Year Planning Horizon. 133 the TA algorithm (Figure 1) is similar to SA but much easier to understand and implement. As in SA, the new candidate is selected randomly from the neighborhood of the existing solution and the objective function for the new solution is calculated. Bettinger et al. (2002) looked at the performance of eight heuristic planning techniques to solve three increasingly difficult wildlife-planning problems. The results showed that despite the simplistic nature of threshold acceptance (TA), it performed well compared to some of the more complex heuristic techniques. The threshold acceptance algorithm was implemented using Microsoft C# programming language. All spatial and yield data was stored in an ESRI geodatabase. Due to the stochastic components in the threshold acceptance algorithm, 20 final solutions were generated for each scenario with each scenario using a new random starting solution. Each solution considered 50,000 iterations at each of the eight threshold levels. The run with the highest objective function for each scenario has been presented in the following section of this paper. Figure 5. Aggregation 1 RESULTS As expected, the smaller planning unit size in aggregation 1 produced a higher net present value than the larger units in aggregation 2 and 3. The result was an increase in net present value of approximately $45 and $225 per acre (Figure 2). These net present values do not include the penalties and incentives. An additional cost of increasing planning unit size is a greater variation in the scheduled volume flow between periods. The smaller planning units allows for them to be selected in a manner that will minimize the volume penalty values (Figure 3). By incorporating the aggregation goal, the final harvest unit size has nearly doubled the average original planning unit size. This suggests that a well designed model can still produce the larger harvest units, without sacrificing the net present value (Figures 47 and Table 2). Figure 6. Aggregation 2 Figure 4. No Aggregation Figure 7: Aggregation 3 134 Table 2. Harvest Unit Size Summary (acres). Harvest Unit Size No Aggregation Aggregation 1 Aggregation 2 Aggregation 3 Maximum 105.0 102.6 119.2 119.7 Minimum 0.4 0.4 20.3 14.0 Average 26.4 40.6 65.9 75.3 Boston, K., and P. Bettinger. 2001. The Economic impact of green-up constraints in the SE USA. Forest Ecology and Management. 145: 191-202. CONCLUSIONS Although in the past planners have pre-aggregated into large units, this is no longer necessary given the current planning tools available. The results of this paper show that there are potential gains to be made in financial performance when smaller units are the primary data incorporated into the model. This paper also demonstrates that goals or constraints can be incorporated to the model formulation that encourages or requires the aggregation of model planning units into larger harvest units with minimal impact on the revenues. By allowing the model to aggregate settings into planning units during the modeling, as apposed to aggregating prior to modeling, allows a larger number of solutions to be explored leading to better results. These results should encourage organizations to utilize the current technology such as GPS and GIS that allows for the creation and management of smaller unit size and to maintain their identity throughout the planning process. Dahlin, B., and O. Sallnas. 1993. Harvest scheduling under adjacency constraints-A case study from the Swedish sub-alpine region. Scand. J. For. Res. 8:281-290. Dueck, G., and Scheuer, T. 1990. Threshold Accepting: A General Purpose Optimization Algorithm Appearing Superior to Simulated Annealing. Journal of Computational Physics 90, 161-175. MAGIS. 2003. A Multi-Resource Analysis and Geographic Information System. www.forestry.umt.edu/magis/ (accessed on 4/27/2003) Manley, B., Papps, S., Threadgill, J., Wakelin, S. 1991. Application of FOLPI. A linear programming estate modeling system for forest management planning. FRI Bulletin. No. 164, 14 pp. LITERATURE CITED Murray, A. 1998. Spatial Restrictions in Harvest Scheduling. For. Sci. 45(1): 45-52 American Forest and Paper Association. 2002. The 2002-2004 Edition, Sustainable Forestry Initiative Program (SFI) SM Oregon Department of Forestry. 2003. Oregon forest practices rules. Salem Oregon. Porter M. 1986. Competition in global industries: A conceptual Framework. In Competition in Global Industries. ed M. Porter. Harvard Business School Press Boston Mass. P 15-60. h t t p : / / w w w. a f a n d p a . o r g / C o n t e n t / N a v i g a t i o n M e n u / Environment_and_Recycling/SFI/Publications1/ C u r r e n t _ P u b l i c a t i o n s / 2 0 0 2 2004_SFI_Standard_and_Verification_Procedures/20022004_SFI_Standard_and_Verification_Procedures.pdf. Accessed Nov 4. 2003. Propper De Callejon, D., T. Lent, M. Skelly. C. A. Webster. 1998. Sustainable Forestry within an Industry Context. In M. B. Jenkins. The John D. and Catherine T. MacArthur Foundation Press. P 2-1 to 2-39. Bettinger, P., D. Graetz, K. Boston, J. Sessions, and W. Chung. 2002. Eight heuristic planning techniques applied to three increasingly difficult wildlife-planning problems. Silva Fennica. 36(2) 561-584. REMSOFT. 2003. Intelligent software for the environment. www.remosft.com (accessed on 04/27/03). 135 136 Preliminary Investigation of Digital Elevation Model Resolution for Transportation Routing in Forested Landscapes MICHAEL G. WING, JOHN SESSIONS AND ELIZABETH D. COULTER Abstract: Several transportation planning decision support systems utilize geographic information systems (GIS) technology to plan forest operations. Current decision support systems do not address upslope terrain conditions that may influence the stability of road networks. This paper describes an on-going research project that uses a GIS and digital elevation model (DEM) to identify transportation route alternatives. Our goal was to develop an algorithm that identified transportation routes guided by an objective function that weighted road grade and potential drainage area. We used a 9 x 9 meter resolution DEM. We found that the resolution of the DEM (9 x 9 meter) was unable to provide reliable road grade and landscape slope estimates. In both road grade and landscape slope results, gradient estimations based on the DEM data appeared to overestimate expected values. These results encourage further investigations, including the use of finer resolution DEMs to model topographic surfaces for transportation routing purposes. INTRODUCTION els for slope and other topographic landscape representations for input into a decision support system. With terrain information, planners may be able to minimize traffic on forested routes that are potentially less stable than others, and reduce road failures, maintenance needs, sediment delivery to streams, and other factors related to transportation costs. Alternately, the identification of problem sites along a transportation network that is in use may help direct monitoring and maintenance efforts. Although a few studies have reported progress in this area (Wing et al. 2001), there remains a need for further work. We present results in this paper of efforts to use a DEM to model road grade, slope conditions, and the amount of upslope contributing area for the use of transportation route planning. We investigated the usefulness of a 9-meter DEM to provide reliable road grade and landscape slope estimates for transportation purposes. An important part of forest operations is the development of an efficient transportation system that incorporates economic, environmental, and safety considerations. Several decision support systems have been developed to assist forest planners with scheduling transportation routes in forested terrain (Reutebuch 1988, Liu and Sessions 1993). Previously, forest transportation planners relied on hard copy maps and other manual techniques for transportation scheduling and were subject to the time and efficiency limits imposed by these techniques. Often, this meant that a full range of options may not have been developed and considered. The development of decision support systems has helped planners with the identification and prioritization of potential transportation routes, given a set of parameters and accompanying constraints. Decision support systems have also allowed planners to quickly create a range of transportation options with indices or other benchmarks through which to evaluate and choose among alternatives. While others have used GIS technology as a decision support system for forested route siting and analysis (O’Neill 1991, MacNaughton et al. 1997, Epstein et al. 1999, Chung 2002, Akay 2002), accounting for potentially unstable or landslide prone sites in the terrain surrounding the road is not considered. With the increasing availability of digital elevation models (DEMs), it is now possible to find elevation data for most parts of the U.S. at 10 meter, or finer, resolution. DEMs can be used to create terrain mod- METHODS Our study area is the Elliott State Forest, located in the Oregon coast range. The Elliott State Forest is an actively managed forest of 145 square miles (376 km2) and has relatively steep terrain with an approximate average ground slope of 53%. The Elliott has a well-developed transportation network (3.8 miles per square mile) with approximately 550 miles (885 km) of roads, both paved and unpaved. We obtained base GIS data from the Elliott staff for our project with layers representing ownership boundaries, roads, and 137 a digital elevation model (DEM) that was derived from aerial photography. All GIS operations were done using either ArcInfo workstation or ArcGIS 8.3 software with the Spatial Analyst extension. Our analyses were completed primarily with raster (grid) data. The Spatial Analyst extension allows users to create and manipulate raster data, and offers a number of tools for calculating preferred routes in a transportation network. Our goal was to guide the search for preferred routes by use of a weighted objective function of environmental variables that influence environmental performance of roads. The weighted objective function was intended to identify the optimal route based on variable weightings. We chose two variables: road grade and upslope contributing area. Road grade was chosen because water power increases with road gradient and steep road grades have been linked with sediment delivery potential from forest roads (Boise Cascade 1999). The second variable was upslope contributing area. Upslope contributing area is intended to provide a potential indicator of saturated terrain conditions: the contributing drainage area flowing into each grid cell. This measure provides a relative index of wetness and is a primary variable for many popular hydrologic models (Beven and Kirkby 1979). To derive road grade, we initially created a raster-based GIS layer of the Elliott road system by converting the existing vector roads layer to this data structure. We then calculated the grade of the Elliott roads by using the existing 9meter resolution DEM from the Elliott forest, overlaying the DEM on the roads layer, and calculating a grade using the SLOPE function within ArcInfo. This resulting grade value was used to approximate road grade. We considered only those DEM cells that were coincident with the roads and did not use other adjacent raster cells. Contributing areas were calculated for each grid cell by summing the area of all cells that drained into that cell. These processes resulted in separate raster layers for road grade and contributing area. We partitioned our analysis into two parts. First we used the existing roads in the forest as the transportation network and constrained our route selection process to consider only these existing roads. Second, we relaxed the route selection process and did not constrain the route location search to existing roads. We determined the shortest route between the landing and exit sites as a function of distance, and then compared this route to others generated through different weightings of road grade and contributing area importance. Our first application used a potential timber landing located in the northeastern portion of the Elliott and an exit site on the forest’s western perimeter (Figure 1). These example sites were chosen as the area between them encompasses a major portion of the Elliott. We found unexpected results in the topographic variable summaries for shortest paths created through our first application. Since we had constrained the search to existing roads, we anticipated that all raster derived grades would be F !( W X Legend !( W X 0 1.5 3 6 9 12 Kilometers Landing Exit Roads Forest Boundary Figure 1. Elliott State Forest, road network, landing location, and exit point. within normal truck operating road gradients. Although the average grades were reasonable, there were a number of road segments with grades exceeding 20 percent. This raised doubts as to the correct location of the existing road network relative to the DEM we used for analysis. To investigate possible road network georeferencing problems, we relaxed the search to examine the entire terrain. If the only problem was georeferencing, we anticipated that the relaxed search would identify routes that avoided the excessive gradients. We used the same landing and exit site but we altered our approach by constructing slope and contributing area models from the DEM for the entire forest. Using these layers, the route selection algorithm was not constrained to the existing road network and was free to consider the entire Elliott forest. For the second application, we also determined the shortest path between the landing and exit sites as well as other routes through the same combination of grade and contributing area weightings that we used in the first application. To facilitate weighting of the slope and contributing area values, both the road grade and contributing area layers were reclassified from continuous data into a 10-category equal area distribution. The equal area distribution creates continuous categories that contain an approximately equal number of observations. We manipulated variable weightings to calculate several different routes from our landing to the forest outlet. RESULTS For the single landing and exit application, the first route we identified was the shortest linear path along the road network from the landing to the exit. This shortest path created a base route for comparative purposes; no weights 138 Table 1. Variable weights and route distance, mean and maximum route grade, and mean and maximum contributing area for existing road network. Grade Weight (%) 0 0 10 25 50 75 90 100 Contributing Area Weight (%) 0 100 90 75 50 25 10 0 Route Distance (miles) 36.5 38.2 38.1 38.0 37.8 36.8 36.8 36.8 Mean Route Grade (%) 9.3 9.5 9.4 9.3 9.2 8.9 8.9 8.9 Max Route Grade (%) 67.2 42.8 42.8 42.8 42.8 67.2 67.2 67.2 Mean Contributing Area (acres) 26.3 12.6 12.7 12.8 12.9 17.0 17.3 27.4 Max Contributing Area (acres) 12916.9 11391.5 11391.5 11391.5 11391.5 12916.9 12916.9 12916.9 Table 2. Variable weights and route distance, mean and maximum route grade, and mean and maximum contributing area for unconstrained network routing. Grade Weight (%) 0 0 10 25 50 75 0 100 Contributing Area Weight (%) 0 100 90 75 50 25 10 0 Route Distance (miles) 18.8 22.5 23.0 23.0 24.7 24.5 22.7 24.5 Mean Route Grade (%) 50.7 35.0 22.7 21.9 18.8 18.6 14.5 11.2 Max Route Grade (%) 137.5 109.8 88.5 171.1 84.0 82.6 88.6 145.1 Mean Contributing Area (acres) 14.6 3.9 7.3 7.3 0.5 9.9 354.9 1026.5 Max Contributing Area (acres) 17337.3 18032.9 18032.9 18032.9 1870.2 43470.7 16000.1 15996.3 mean route grade decreased and contributing area increased, although the changes in mean route grade were not pronounced. The high grade values that resulted from our initial analyses indicated that the DEM we used was not able to provide sufficient information for determining reliable road grade estimates. Possible sources of error are DEM resolution, incorrect location of the existing road network relative to the DEM, a processing error, or an inherent bias in the methodology that creates the raster elevation values. To test for a road location error, we applied a different approach to route creation with the shortest path algorithm and different weightings of topographic variables; we did not constrain routes to the existing road network. By allowing the shortest path algorithm to navigate freely throughout the Elliott’s topography, we believed that the identification of routes with less than a 20% grade maximum would confirm a road location error. Results varied more dramatically for the unconstrained routing approach. The shortest path had a distance of 19 miles, mean grade of 50%, and a mean contributing area of 14.63 acres (Table 2). Route distances were dramatically shorter for the all of the unconstrained routes when compared to the network constrained routes. Distances ranged from approximately 23 to 25 miles. Route mean and maximum grades were also very different than the network con- were applied in this initial route for road grade or contributing area. We then selected routes based on varying weights of road grade and contributing area in order to determine optimal routes given a range of variable importance (Table 1). Regardless of the route parameters, the range of resulting route distances was consistent. The shortest distance between these points identified by the base route was about 37 miles and, with modifications of the grade and contributing area variables, the range of distances was between 37 and 38 miles. The mean grade of all routes was also consistent and ranged from 8 to 9%. The maximum grade differed from 43 to 67%. These high gradient sections clearly exceed the gradients of the existing road network. Contributing areas differed markedly with the shortest path having a mean contributing area of 26 acres and the routes resulting from considering grade and contributing area having a range from 13 to 27 acres. The shortest path and routes that had a grade weight of at least 75% all had the same maximum contributing area (12,917 acres) whereas all other routes had a maximum of 11,392 acres. Closer inspection of the location of these routes revealed that the maximum contributing areas all occurred along the same small section of road. This section of road crossed a major stream twice and gained the large contributing area values associated with the stream. In general, as the weight of the road grade increased (and contributing area decreased), the 139 strained routes; grades were consistently and, in many cases, considerably higher. Contributing area results for the unconstrained routes varied considerably in terms of the mean. With the exception of the 90 and 100% grade weights, mean contributing areas were roughly half, or less, than those of the constrained routes. Maximum contributing areas were larger for all of the unconstrained route results and, with two exceptions, ranged from 16,000 to 18,000 acres. These large maximum contributing areas were again the result of routes crossing major streams. statistics that were not constrained to the existing network. Whereas the average grade (47%) was slightly less than our previous results (Table 2), the maximum grade was slightly larger (143%). These similar results led us to believe that it was not a processing error, or necessarily inaccuracy, in our original DEM that contributed to the large grade values. Rather, a more likely explanation is that a finer resolution DEM is needed to provide a more reliable approximation of road grade and terrain. The DEMs we used were unable to accurately capture the lower gradients that should exist along the existing road network. In addition, our inability to create any route throughout the forest that avoided grades above 20% suggests that slopes were systematically over estimated throughout the forest. Wilson et al. (2000) detected differences in slope as a function of DEM resolution, and considered resolutions between 30 and 200 meters. One approach to verifying systematic slope exaggeration estimates would be to create or obtain finer resolution (1-5 meter) DEM data for the Elliott, calculate slope values, and examine values to compare differences with our reported findings. These could be compared to measured road grades and cross section data through precision instrumentation, such as total station or digital clinometer, in order to better understand what is being represented in the DEM. DISCUSSION All of the routes we developed from the existing road network had average road grades that were well within normal acceptable grade tolerances (16-20%). In addition, all routes also had maximum grades of more than 42%. While travel distances were significantly lower for the routes we created that were not constrained to existing roads, all had maximum grades greater than 80% and all had average grades that exceeded those that were created using the existing Elliott road network. These results indicated that the DEM values were not providing reliable grade and slope data. We wanted to determine whether viable transportation routes could be determined through the 9 meter DEM. In order to create transportation routes that could be used by typical log hauling vehicles, we adjusted the constraints of our routing algorithms so that grades in excess of 20% would not be considered in final route creation. We then attempted to create routes that avoided 20% grades through the confines of the existing network and also through the unconstrained approach, where the entire landscape would be potentially available for transportation routes. We found that this was not possible; every route possibility included multiple grade values that exceeded 20%. Given that many parts of the existing route system have been used for log hauling, these results shed doubt on the reliability of the DEM that served as the basis for our topography representations. We suspected that perhaps a processing error could have contributed to these results and contacted the Elliott staff to verify the DEM’s history. The base DEM was created from elevation points derived from aerial photography taken in 1996. The points were converted into a triangular irregular network (TIN) data structure and then converted into a raster file. We used the resulting raster file for our analysis. Potential errors could have occurred during operations performed on the data prior to our receiving the data, or could have resulted from our manipulations during this project. For comparative purposes, we obtained USGS 10 meter DEMs for the Elliott and used these data to create slope and contributing area models of the Elliott. The average slope of the USGS DEM was slightly less (50%) than the photogrametrically derived DEM (53%). We then used the baseline USGS data to calculate a shortest path and grade LITERATURE CITED Akay, A. 2002. Minimizing total cost of construction, maintenance, and transportation costs with computer-aided forest road design. PhD dissertation, Oregon State University, Corvallis. 229 p. Boise Cascade Corporation. 1999. SEDMODL-Boise Cascade road erosion delivery model. Technical documentation. Boise Cascade Corporation, Boise, ID. 19 p. Beven, K. J. and M.J. Kirkby. 1979. A physically based variable contributing area model of basin hydrology. Hydrological Sciences Bulletin 24(1):43-69. Chung, W. 2002. Optimization of cable logging layout using a heuristic algorithm for network programming. Phd dissertation, Oregon State University, Corvallis. 206 p. Epstein, R., A. Weintraub, J. Sessions, J. B. Sessions, P. Sapunar, E. Nieto, F. Bustamante, and H. Musante. 1999. PLANEX: an equipment and road location system. In Proceedings of the International Mountain Logging and 10th Pacific Northwest Skyline Symposium, March 28-April 1, 1999, Dept. of Forest Engineering, Oregon State University, Corvallis. pp. 365-368. Kramer, B. W. 2001. Forest road contracting, construction, and maintenance for small forest woodland owners. Research Contribution 35, Forest Research Laboratory, Oregon State University, Corvallis. 140 Liu, K. and J. Sessions. 1993. Preliminary planning of road systems using digital terrain models. Journal of Forest Engineering 4:27-32. Wilson, J. P., P. L. Repetto, and R. D. Snyder. 2000. Effect of data source, grid resolution, and flow routing method on computed topographic attributes. In: Wilson J P and J C Gallant (editors), Terrain Analysis: Principles and Applications. New York, John Wiley and Sons, pp 133-161. MacNaughton, J., J. Sessions, and S. Xu. 1997. Preliminary planning of forest roads using ARC GRID. In: GIS ‘97 Conference Proceedings. Fort Collins: GIS World, Inc. 67-71. Wing, M. G., E. D. Coulter, and J. Sessions. 2001. Developing a decision support system to improve transportation planning in landslide prone terrain. In Proceedings of the International Mountain Logging and 11th Pacific Northwest Skyline Symposium, December 10-11, 2001, College of Forest Resources, University of Washington and International Union of Forestry Research Organizations, Seattle, WA. pp. 56-60. O’Neill, W. A. 1991. Developing optimal traffic analysis zones using GIS. ITE Journal 61: 33-36. Reutebuch, S. 1988. ROUTES: A computer program for preliminary route location. USDA General Technical Report. PNW-GTR-216, Portland, OR. 18 p. 141 142 Comparison of Techniques for Measuring Forested Areas DEREK SOLMIE, LOREN KELLOGG , MICHAEL G. WING ANDJIM KISER Abstract: Operational planning and layout are important steps in determining the feasibility of harvesting operations. Higher-precision technologies may increase measurement accuracy and efficiency while decreasing total planning costs. Although a number of trials have been completed on the potential implementation of some of these new technologies, few have quantified the benefits of such devices in an operational setting. Sixteen (~1 ac) units were identified for an evaluation of different spatial data-collection instruments as well as techniques for measuring area. Unit boundaries were measured by three surveying techniques, comprising 1) a string box, manual compass, and clinometer; 2) a laser, digital compass, and digital data collector; and 3) a global positioning system. The collected data were compared with a series of benchmarks established with a total station. Techniques were statistically analyzed and error distributions were developed at either a unit or an individual data-point scale. Time studies were conducted to determine the overall efficiencies of each technique. Our results should assist forest resource managers in their decisions when selecting alternate measurement tools for collecting spatial data. INTRODUCTION methods are needed to fully understand the benefits of these newer tools. Global positioning systems (GPS) have also been used to collect spatial data in forested environments (Forgues 1998). Studies have identified variables that affect its usefulness, including the amount of canopy closure (Stjernberg 1997, Mancebo and Chamberlain 2001), receiver type and grade (Darche 1998), weather conditions (Forgues 2001), and topography (Liu and Brantigan 1996). Historically, one of the challenges when using GPS has been the effect of multi-path signals caused by the forested canopy (Stjernberg 1997, Forgues 2001). However, this effect has largely been mitigated by manufacturers incorporating ‘multipath’ recognition into their firmware. Signal availability is another problem (Karsky et al. 2000), primarily of the limited visibility of satellites due to forest cover and topography. Differential GPS (DGPS) has been to be a cost-effective technique for measuring land areas (Liu and Brantigan 1996). Both forest canopy and undulating terrain exert a definite effect on traverse surveys completed by DGPS, with accuracy being reduced as variations in canopy closure and topography increase. Nevertheless, kinematic DGPS traverses have proved more capable of achieving a closer forest stand-area approximation than that obtained from a traditional compass-and-chain traverse. The objectives of this study were to: 1) gather time and costing information to determine the relative efficiencies of each measurement technique; 2) to compare information on precision and accuracy of each method; and 3) to analyze the patch-orientation due to discrepancies in angular measurements. Studies of conventional methods employed in the Pacific Northwest have analyzed the use of nylon tapes, handheld compasses, and clinometers for operational measurements. Researchers have reported that costs can vary according to the type of harvesting system (Edwards 1993, Kellogg et al. 1998), unit size and shape (Dunham 2001a), silvicultural treatments (Kellogg et al. 1991, Edwards 1993, Kellogg 1996a, Dunham 2001a, b), and level of crew experience (Kellogg et al. 1996b). However, no studies have been published concerning more recent data-capturing technologies available to the forest industry. Higher-precision technologies may increase measurement accuracy and efficiency while decreasing total planning costs. Although a number of trials have been completed on the potential implementation of some of these new technologies, few have quantified the benefits of such devices in an operational setting. Mixed results have been reported for the usefulness of electronic distance- and azimuth-measuring (EDM) devices to traverse forest stand boundaries (Liu 1995) and low volume road surveys (Moll 1993). The distance- and vertical angle-measuring capabilities of the lasers generally met the survey requirements, but the azimuth measurements with the compass did not due to offsets in the magnetic field. Several studies have illustrated the potential for digital data collectors, compasses, and laser rangefinders in operational settings including woodpile volumes (Turcotte 1999) and skyline corridor traversing (Wing and Kellogg 2001). In operational settings measurements are often difficult to obtain due to understory brush. Further comparisons between the laser rangefinder and more conventional 143 METHODS distance, inclination, and azimuth were recorded in the data collector. Two data recorders, one operating on a Windows CE and DOS platform (Juniper Allegro), the other on a Windows CE platform (Tripod Data Systems (TDS) Ranger), were used in tandem with the laser to determine the most efficient data recording technique. One advantage in using a DOS-based application was that the data could be directly downloaded into the mapping software. The office work for the Juniper data collector consisted of downloading the information to a desktop computer via an ActiveSync program. Data Plus software allowed the user to structure the database to match the required input for the mapping program. The data were then imported into RoadEng, using the Terrain Module, and subsequently analyzed. Office work for the TDS data collector involved a computer spreadsheet program that adjusted the coordinates to a format that RoadEng could recognize. The third survey technique incorporated a Trimble Pro XR GPS. A one-person crew traversed the perimeter of the patches, simultaneously logging points and using the area function within the TSC1 data collector while moving between stations. This traverse was completed in a kinematic mode, so that no differentiation existed among the stations but, rather, the entire boundary was traversed as a single segment. Therefore, the GPS portion of this study did not include between-station measurements, and comparisons could be made only at the patch level. Data were downloaded to Trimble Pathfinder Office version 2.01 and base station data were used to differentially correct the data and determine patch areas. The previously described techniques were also compared with a benchmark method that could produce the most accurate measurements. A Nikon DT-310 total station was used along with 2 prisms and a four-person crew. A side shot method was used that minimized the number of instrument set-ups required to traverse the patch (Fig. 1), while collecting measurements at each station. Two crew members maneuvered prisms between the stations, while another cleared sight-paths between the total station and the survey points. Study Site This study was located on the McDonald-Dunn College Forest, managed by Oregon State University. The site is a 55-year-old mixed stand comprised primarily of Douglasfir (Pseudotsuga menziesii), big leaf maple (Acer macrophyllum), and red alder (Alnus rubrus). The stand also had minor shrub vegetation consisting of vine maple (Acer circinatum), salal (Gaultheria shallon), and salmonberry (Rubus spectabilis). Slopes ranged from 0 to 76% (average of ~25%). Canopy closure was 60 to 95%, with an average stand density of 280 trees per acre. The average tree was ~97 ft tall, with a dbh of 17 in.. Approximate volume per acre was 15 to 24 mbf. Data Collection Sixteen study patches (~1 ac each) were selected based on stand descriptions, topographies, and their location relative to other patches. Boundaries were delineated in the field with surveyor’s flagging and paper tags. Benchmark stations, established along the vertices of each patch, were flagged and locations measured with a Nikon DT-310 total station. Measurement accuracies were reported to within 0.02 in. of the horizontal and vertical distances. Three techniques for determining land area were compared against the benchmark measurements. These included the use of: 1) a string box with a distance counter, handheld compass and a Suunto clinometer; 2) electronic distance- and electronic bearing-measurement devices; and 3) a global positioning system. Time required to complete the operational layout and planning was separated into three components to determine the relative efficiencies of each method as time spent surveying each patch, recording the data, and either downloading or entering the information into a database. Crew sizes depended on the surveying method being employed. All members had at least one year of experience with the survey equipment and were proficient in its operation. The first method consisted of a single person measuring slope distance and slope percent. Data were collected station to station, recorded in a field book and manually entered into a software program, RoadEng (Softree; Vancouver, BC), in the office. Traverse adjustments were done using the compass rule (Mikhail and Gracie 1981, Buckner 1983). The second method employed an electronic distance- and electronic bearing-measurement device manufactured by Laser Technology Incorporated (LTI). The Impulse 200 EDM was linked with a MapStar digital compass, which provided data on slope distance, slope percent, and horizontal angles. This system required a two-person crew, and the collected data were logged into a handheld digital data recorder. The lead traverser maneuvered between stations and held the reflective prism at eye level, directly above the pin flag. The rear traverser aimed the laser at the reflective prism and the LEGEND Total Station Measurements Stations Figure 1. Side shot method for traversing with a total station. 144 140 Field Survey and Office Data Entry Time(Min) 120 100 80 60 40 20 0 10 11 12 14 15 16 17 18 19 20 21 22 23 24 25 26 Patch Number String Laser GPS Total Station Figure 2. Time required to complete various forest-area measurement techniques. The string-box method required 19 minutes more per patch (195% increase) compared with the laser/Juniper data collector. Contributing factors included the time required to manually record the data in the field book and the need to manually transcribe the field notes, whereas the laser method included a digital download. Likewise, the GPS method took 23 minutes longer (210% increase) than the Juniper data collector, mainly because of intermittent satellite reception due to topography, canopy closure, and satellite orbits. The GPS data includes time spent on those patches that were abandoned after one hour because of poor satellite configuration. Average time difference for the total station was 54 minutes per patch (370% increase) compared with the laser/ Juniper data collector. This was due primarily to required instrument set-up time. The time, type of equipment, and crew size required to complete a traverse were used to calculate the variable cost of each survey method (Table 1). Equipment was depreciated over an two-year period. Hourly wages, which included benefits were obtained from the 2001 Associated Oregon Loggers Annual Wage Survey (Salem, OR, USA). Survey data were transformed to determine the x, y, and z coordinates, which were then downloaded as an ASCII file into the RoadEng. RESULTS AND DISCUSSION Time and Cost Information Time required to survey a patch and complete the office work varied substantially, depending on the technique (Fig. 2). The task was considered complete when all information was processed and entered into a common mapping program. The method involving the laser, digital compass, and Juniper data collector required the least amount of time per patch (17 minutes). The second most time-efficient technique was that using the laser, digital compass, and the TDS data collector. The latter method required approximately two extra minutes per patch because of the additional step taken by the TDS data collector to arrange the data in an acceptable format for the mapping program. Table 1. Costs involved in completing land-area survey of 16 forested patches. Method String Box Laser (Ranger) Laser (Allegro) GPS Total Station Crew size Labor cost ($/hr) 1 2 2 1 4 18.90 37.80 37.80 18.90 75.60 Equipment cost ($/hr) 0.05 1.18 1.31 1.88 1.13 145 Total time (hr) 10.3 5.9 5.3 11.5 19.7 Total cost ($) 195.19 229.98 207.28 238.97 1511.58 Cost per acre ($) 11.67 13.22 11.92 38.86 86.52 Cost per mbf ($) 0.62 0.70 0.63 1.85 4.59 Table 2. Precision of measurements for patch areas, by survey method. Method String Box Laser GPS Total Station Mean patch area (ac) 1.03 1.06 1.03 1.07 Mean difference in patch area (ac) -0.04 -0.01 -0.04 0 Percent difference (%) 3.7 0.93 3.7 0 Mean patch precision (%) 1.15 2.65 N/A 0.014 Table 3. Average distance errors produced by each survey method. Method String Box Laser GPS Total Station Mean slope distance error (ft) 3.02 1.33 N/A 0 Mean horizontal distance error (ft) 2.78 1.14 N/A 0 Mean vertical distance error (ft) 2.82 1.81 N/A 0 Table 4. Average mean effectiveness values for each survey method. Method String Box Laser (Ranger) Laser (Allegro) Total Station Total Cost ($) 195.19 229.98 207.28 1511.58 Closing Error (%) 1.15 2.65 2.65 0 n 16 16 16 16 Mean Effectiveness (M.E.) 1236 3247 2902 94 mounted on a staff. Although the manufacturer’s accuracies had been achieved in trials with the equipment mounted, this positioning was found to limit the user’s mobility in the forested environment. The mean precision of the string box (1.15%) was good given the perceived minimal precision of the equipment. Accuracy is defined as the degree of conformity or closeness of a measurement to the true value (Buckner 1983). Survey methods were analyzed for significant differences at the station level (Table 3). Average accuracy was calculated from the difference in measurements between the total station and each of the laser and string box methods. GPS data are station independent and thus were not involved. String box error can be attributed to several factors. For example, use of the string box was affected by the amount of brush and branches between stations. The string may have gotten caught on the branches, preventing the traverser from following a straight path. Likewise, the string may have become taut when maneuvering around obstacles, thereby contributing to the error. Errors for the total station and laser methods were primarily brought due to the operator. Both the laser and the target had to be positioned vertically above the station. A common problem involved the laser operator needing to bend and shift away from the station in order to gain a clear sight path toward the target. Hourly labor and initial equipment costs played the most significant role in overall operating costs. The difference between the two digital-data methods could be attributed to the additional office time the TDS data collector required for formatting the field data. PRECISION AND ACCURACY Precision is the degree of closeness or conformity among repeated measurements of the same quantity (Mikhail and Gracie 1981). The average patch precision gained from each of our methods is shown in Table 2. The mean difference in area compared to the total-station technique appears to be relatively small. Although this was a fairly small land area (1742 ft2), one may assume a random error effect, for which the percent error would hold fairly constant. Therefore, this effect could dramatically impact area calculations, timber volume estimates, and other operational considerations on larger unit areas. Because the survey of each patch started and ended at the same point, the precision or repeatability could be calculated from the difference in coordinates. This difference was then divided by the total perimeter distance for each patch, resulting in a percent error term that was averaged for the16 traversed patches. The laser and compass method produced the least amount of precision because the instrument was not 146 Figure 3. Differences in patch orientation generated by survey methods. A multiple range test was used to confirm that all data points were from the same population. In addition, t-tests were conducted at the patch level to determine significant differences in accuracy. Values from both the laser and the string-box methods were significantly different (p<0.05) from those obtained with the total-station technique. Likewise, the t-test used to compare the string box and laser data also indicated a significant difference between these two methods (p<0.05). volved a combination of techniques for each method, a total-cost variable was calculated. Mean effectiveness (M.E.) for each method (Table 4) was calculated by multiplying the total cost by the closing error and dividing this by the number of patches (16) to determine a mean effectiveness. Here, the smaller the value, the more effective the surveying method. ((Total Cost)( M.E. = ORIENTATION Closing Error % Closing Error Total Station % )) n The total-station technique was the most effective, although it was the most time-consuming and expensive of all the methods, it had a significantly smaller closure error. The large difference between the laser method and the string-box technique was a result of the higher accuracy and initial costs associated with the former. Effectiveness with the GPS method was not included because no level of accuracy had been calculated. Although all traverses closed with adequate precision and approximately equal areas regardless of the survey technique employed, orientations varied substantially (Fig. 3). This effect on alignment might have major consequences for a number of tasks completed during operational planning and layout. For example, such errors could be costly to both parties when working with legal boundaries between property owners. This difference was most evident when the digitalcompass method was implemented because the position at which the user held the equipment influenced the reading. Although very good closing precision could be attained, large deviations from patch alignment occurred. This effect could have been minimized by mounting the laser and digital compass on a staff. SUMMARY Different methods for measuring forest areas may be used to meet specific land-management objectives. This study compared four techniques for completing a traverse of partial harvests within an uneven-aged management plan. The method entailing the string box, manual compass, and clinometer was approximately 6% less expensive than the laser method. However, although the initial purchase price and labor rates with the string-box technique were lower, 48% more time was spent conducting the traverse of all the patches. The total-station technique was the most expensive because of the larger crew and time required to clear the sight lines. EFFECTIVENESS It is difficult to account for practicality when comparing survey techniques. Liu (1995) assessed individual methods that used different equipment by multiplying the time needed to complete a task by the resulting accuracy, thereby basing effectiveness on time instead of cost. Because our study in147 The effectiveness of each survey method also varied substantially. Low (better) values were the result of a combination of small costs and/or high accuracies. The total-station method rated well (value = 94) because of the high amount of precision gained with its use. Although it was the most expensive to operate, its resulting precision was magnitudes higher than that gained by the other methods. Relative to their specific measurement activities, each method has its strengths (time, cost, and accuracy) and weaknesses (alignment, repeatability, and cost). Therefore, the potential benefits must be weighed when allocating resources to specific duties for operational planning. In conclusion, this study illustrated that, although time was saved by using the digital instruments, their performances were not always as effective as those achieved via traditional methods. Kellogg, L.D., Pilkerton, S., and R. Edwards, 1991. Logging Requirements to Meet New Forestry Prescriptions. P. 4349 In Proceedings of Council of Forest Engineering Annual Meeting, Nanaimo, BC, Canada. Kellogg, L.D., Bettinger, P., and R.M. Edwards, 1996a. A Comparison of Logging Planning, Felling, and Skyline Costs between Clearcutting and Five Group-Selection Harvesting Methods. Western Journal of Applied Forestry 11(3): 9096. Kellogg, L.D., Milota, G.V., and M. Millar Jr., 1996b. A Comparison of Skyline Harvesting Costs for Alternative Commercial Thinning Prescriptions. Journal of Forest Engineering 1: 7-23. Kellogg, L.D., Milota, G.V., and B. Stringham, 1998. Logging Planning and Layout Costs for Thinning: Experience from the Willamette Young Stand Project. Forestry Publications Office, Oregon State University, Corvallis, OR, USA. 20 p. LITERATURE CITED Buckner, R.B., 1983. Surveying Measurements and their Analysis. Landmark Enterprises, Rancho Cordova, CA, USA. 275 p. Liu, C.J., 1995. Using Portable Laser EDM for Forest Traverse Surveys. Canadian Journal of Forestry Research 25: 753766. Darche, M.-H., 1998. A Comparison of Four New GPS Systems under Forestry Conditions. Forest Engineering Institute of Canada Special Report 128, Pointe Claire, Quebec, Canada. 16 p. Liu, C.J., and R. Brantigan, 1996. Using Differential GPS for Forest Traverse Surveys. Canadian Journal of Forestry Research 25: 1795-1805. Dunham, M.T., 2001a. Planning and Layout Costs I: Group Selection and Clear-cut Prescriptions. Forest Engineering Research Institute of Canada, Vancouver, BC, Canada. 2(22): 6. Mancebo, S., and K. Chamberlain, 2001. Performance Testing of the Trimble Pathfinder Pro XR Global Positioning System Receiver. USDA Forest Service Technical Note 10 p. Dunham, M.T., 2001b. Planning and Layout Costs II: Tree Marking Costs for Uniform Shelterwood Prescriptions. Forest Engineering Research Institute of Canada, Vancouver, BC, Canada. 2(34): 4. Mikhail E.M., and G. Gracie, 1981. Analysis and Adjustment of Survey Measurements. van Nostrand Reinhold Company, New York, USA. 340 p. Edwards, R.M., 1993. Logging Planning, Felling, and Yarding Costs in Five Alternative Skyline Group Selection Harvests. Master of Forestry paper, Department of Forest Engineering, Oregon State University, Corvallis, OR, USA. 213 p. Moll, J.E., 1993. Development of an Engineering Survey Method for Use with the Laser Technology, Inc. Tree Laser Device. Master of Science thesis, Department of Civil Engineering, Oregon State University, Corvallis, OR, USA. 74 p. Forgues, I., 1998. The Current State of Utilization of GPS and GIS Technologies in Forestry. Forest Engineering Research Institute of Canada Field Note, Pointe Claire, Quebec, Canada. 2 p. Stjernberg, E., 1997. A Test of GPS Receivers in Old-growth Forest Stands on the Queen Charlotte Islands. Forest Engineering Institute of Canada Special Report 125, Vancouver, BC, Canada. 26 p. Forgues, I., 2001. Trials of the GeoExplorer 3 GPS Receiver under Forestry Conditions. Forest Engineering Research Institute of Canada, Pointe Claire, Quebec, Canada. 2(8): 4. Turcotte, P., 1999. The Use of a Laser Rangefinder for Measuring Wood Piles. Forest Engineering Institute of Canada Field Note 76, Pointe Claire, Quebec, Canada. 2 p. Karsky, D., Chamberlain, K., Mancebo, S., Patterson, D., and T. Jasumback, 2000. Comparison of GPS Receivers under a Forest Canopy with Selective Availability Off. USDA Forest Service Project Report 7100. 21 p. Wing, M., and L.D. Kellogg, 2001. Using a Laser Range Finder to Assist Harvest Planning. P. 147-150 In Proceedings of the First International Precision Forestry Cooperative Symposium, Seattle, WA, USA. 148 149 Poster Abstracts Can Tracer Help Design Forest Roads? ABDULLAH E. AKAY, GRADUATE RESEARCH ASSISTANT, DEPARTMENT OF FOREST ENGINEERING, COLLEGE OF FORESTRY, OREGON STATE UNIVERSITY, CORVALLIS, OR 97331 JOHN SESSIONS, PROFESSOR, DEPARTMENT OF FOREST ENGINEERING, COLLEGE OF FORESTRY, OREGON STATE UNIVERSITY, CORVALLIS, OR 97331 CPLAN: A Computer Program for Cable Logging Layout Design WOODAM CHUNG, ASSISTANT PROFESSOR, SCHOOL OF FORESTRY, UNIVERSITY OF MONTANA, MISSOULA, MT 59812 JOHN SESSIONS, PROFESSOR, DEPARTMENT OF FOREST ENGINEERING, OREGON STATE UNIVERSITY, CORVALLIS, OR 97331 Abstract: A computerized method for optimizing cable logging layouts using a heuristic network algorithm has been developed. A timber harvest unit layout is formulated as a network problem. Each grid cell containing timber volume to be harvested is identified as an individual entry node of the network. Mill locations or proposed timber exit locations are recognized as destinations. Each origin will then be connected to one of the destinations through alternative links representing alternative cable corridors, harvesting equipment, landing locations, and truck road segments. A heuristic algorithm for network programming is used to solve the cost minimization network problem. A computerized model has been developed to implement the method. Logging feasibility and cost analysis modules are included in the model in order to evaluate the logging feasibility of alternative cable corridors and estimate yarding and transportation costs. 150 List of Contributors Jeffrey Adams Virginia Tech 775A Sterling Drive Charleston, SC 29412 USA jeadams@vt.edu Arnab Bhowmick University of Washington College of Forest Resources Box 352100 Seattle, WA 98195-2100 USA arnabqis@hotmail.com Kamal Ahmed University of Washington 121C More Hall, Box 352700 Seattle, WA 98195-2700 USA kamal@u.washington.edu Tom Bobbe USDA Forest Service Remote Sensing Applications Center 2222 W. 2300 S. Salt Lake City, UT 84119 USA tbobbe@fs.fed.us Abdullah E. Akay Oregon State University Corvallis, OR 97331 USA akaya@ucs.orst.edu Jeremy Allan Intermap Technologies Corp. Calgary, Alberta, CANADA T2P 1H4 Kevin Boston Oregon State University Department of Forest Engineering 213 Peavy Hall Corvallis, OR 97331-5706 USA kevin.boston.cof.orst.edu Hans-Erik Andersen University of Washington College of Forest Resources Box 352100 Seattle, WA 98195-2100 USA hanserik@u.washington.edu David Briggs University of Washington College of Forest Resources Box 352100 Seattle, WA 98195-2100 USA dbriggs@u.washington.edu Kazuhiro Aruga Oregon State University Peavy Hall Department of Forest Engineering Corvallis, OR 97331-5706 USA aruga@fr.a.u-tokyo.ac.jp Ward Carson University of Washington Pacific Northwest Research Station Box 352100 Seattle, WA 98195-2100 USA carsonw@u.washington.edu R. James Barbour USDA Forest Service Pacific Northwest Region PO Box 3623, Portland, Oregon 97208-3623 USA jbarbour@fs.fed.us Woodam Chung School of Forestry University of Montana Missoula, MT 59812 USA wchung@forestry.umt.edu B. Bruce Bare University of Washington College of Forest Resources Box 352100 Seattle, WA 98195-2100 USA bare@u.washington.edu Jennie L. Cornell Oregon State University Forest Engineering Operations Corvallis, OR 97331 USA bryancornell7014@msn.com 151 Sean Hoyt University of Washington Box 352500 Seattle, WA 98195 USA naestyoh@u.washington.edu Elizabeth Coulter Oregon State University Department of Forest Engineering 215 Peavy Hall Corvallis, OR 97333 USA Elizabeth.Coulter@orst.edu Loren Kellogg Oregon State University Department of Forest Engineering 213 Peavy Hall Corvallis, OR 97331-5706 USA loren.kellogg.cof.orst.edu Bill Dyck Ltd. PO Box 11236 Palm Beach, Papamoa 3003 NEW ZEALAND billdyck@xtra.co.nz John R. Erickson USDA Forest Service Forest Products Laboratory One Gifford Pinchot Drive Madison, WI 53726-2398 USA Andrei Kirilenko Purdue University Department of Forestry and Natural Resources Forestry Building 195 Marsteller St. West Lafayette, IN 47907-2033 USA kirilenk@fnr.purdue.edu Alexander Evans Yale School of Forestry & Environmental Studies 205 Prospect Street New Haven, CT 06511 USA Jim Kiser Oregon State University Department of Forest Engineering 213 Peavy Hall Corvallis, OR 97331-5706 USA Jim.Kiser.cof.orst.edu Stephen E. Fairweather Mason, Bruce, & Girard, Inc. 707 SW Washington St., Suite 1300 Portland, OR 97205 USA sfairweather@masonbruce.com Bruce Larson University of British Columbia 2329 West Mall Vancouver, BC V6T 1Z4 CANADA blarson@interchange.ubc.ca John W. Forsman School of Forestry and Wood Products Michigan Technological University Houghton, MI 49931 USA Email:jwforsman@mtu.edu Hamish Marshall Oregon State University Forest Engineering Department 215 Peavy Hall Corvallis, OR 97331-5706 USA hamish.marshall@orst.edu Jeffrey R. Foster Forestry Branch, Fort Lewis Military Reservation Fort Lewis, WA USA Joel Gillet Applanix Corp 85 Leek Crescent Richmond Hill, ON L4B 3B3 CANADA JGillet@applanix.com John Mateski Western Helicopter Services, INC. PO Box 369 Newberg, OR 97132 USA westernhelicopter@earthlink.net Richard A. Grotefendt University of Washington College of Forest Resources Box 352100 Seattle, WA 98195 USA grotefen@u.washington.edu Brett Martin Prudue University 2226 Willowbrook Dr. Apt. #192 West LaFayette, Indiana 47906 USA brettm@fnr.purdue.edu 152 Douglas J. Martin Martin Environmenta; 2103 North 62nd Street Seattle, WA 98103 USA doug@martinenv.com Rober J. Ross USDA Forest Service Forest Products Laboratory One Gifford Pinchot Drive Madison, WI 53726-2398 USA rjross@fr.fed.us Glen Murphy Oregon State University Forest Engineering Department Peavy 271 Corvallis, OR 97331 USA glen.murphy@orst.edu Peter P. Siska Stephen F. Austin State University 1639 North Street Nacogdoches, TX 75962 USA siska@sfasu.edu Peter Schiess University of Washington College of Forest Resources Box 352100 Seattle, WA 98195 USA schiess@u.washington.edu Ross F. Nelson NASA Biospheric Sciences Branch Code 923 NASA Goddard Space Flight Center Greenbelt, MD 20771 USA ross@ltpmail.gsfc.nasa.gov Daniel L. Schmoldt USDA/CSREES/PAS Instrumentation & Sensors Mail Stop 2220 Washington, DC 20250-2220 USA dschmoldt@reeusda.gov Sam Pittman University of Washington Box 352100 Seattle, WA 98195 USA Sam.Pittman@weyerhaeuser.com Stephen P. Prisley Virginia Tech 229 Cheatham Hall Blacksburg, VA 24061 USA prisley@vt.edu Gerard Schreuder University of Washington College of Forest Resources Box 352100 Seattle, WA 98195 USA gsch@u.washington.edu John Punches Douglas Co Extension Oregon State University 1134 SE Douglas Roseburg, OR 97470-4344 USA John.Punches@oregonstate.edu John Sessions Oregon State University 213 Peavy Hall Corvallis, OR 97331-5706 USA John@Sessions.cof.orst.edu Steve Reutebuch University of Washington Pacific Northwest Research Station Box 352100 Seattle, WA 98195-2100 USA sreutebu@u.washington.edu Guofan Shao Purdue University Department of Forestry and Natural Resources Forestry Building 195 Marsteller St. West Lafayette, IN 47907-2033 USA gshao@fnr.purdue.edu Alex Sinclair Feric Western Division 2601 East Mall Vancouver, BC V6T 1Z4 CANADA alex-s@vcr.feric.ca Luke Rogers University of Washington Rural Technology Initiative Seattle, WA 98195 USA lwrogers@u.washington.edu 153 Derek Solmie Oregon State University Department of Forest Engineering 215 Peavy Hall Corvallis, OR 97330 USA derek.solmie@orst.edu Bernd-M. Straub University of Hannover Institute for Photogrammetry and GeoInformation Nienburger Strasse 1 Hannover 30167 GERMANY bernd-m.straub@ipi.uni-hannover.de Pierre Turcotte FERIC 580 Boul. St-Jean Pointe-Claire, QC H9R 3J9 CANADA pierre-t@mtl.feric.ca Rien Visser Virginia Tech 229 Cheatham Hall Blacksburg, VA 24061 USA visser@vt.edu Xiping Wang University of Minnesota Duluth C1 Gifford Pinchot Drive Madison, WI 53705-2398 USA xwang@fs.fed.us Graham West Forest Research Private Bag 3020 Rotorua, NEW ZEALAND Graham.West@ForestResearch.co.nz Denise Wilson University of Washington PO Box 352500 Seattle, WA 98195-2500 USA wilson@ee.washington.edu Michael G. Wing Oregon State University Forest Engineering Department 213 Peavy Hall Corvallis, OR 97333 USA michael.wing@orst.edu Jianyang Zheng University of Washington Department of Civil and Environmental Engineering Seattle, WA 98195-2700 USA David Yates Forest Technology Group 3950 Faber Place Dr. North Charleston, SC 29405 USA david.yates@ftgrp.com List of Attendees Jeffrey Adams Virginia Tech 775A Sterling Drive Charleston, SC 29412 USA jeadams@vt.edu Hans-Erik Andersen University of Washington College of Forest Resources Box 352100 Seattle, WA 98195-2100 USA hanserik@u.washington.edu Carolyn Anderson Weyerhaeuser Company 361 Schooner Cove, NW Calgary, Alberta T3L123 CANADA carolynn.anderson@weyerhaeuser.com Kazuhiro Aruga Oregon State University Peavy Hall Dept of Forest Engineering Corvallis, OR 97331-5706 USA aruga@fr.a.u-tokyo.ac.jp B. Bruce Bare University of Washington College of Forest Resources Box 352100 Seattle, WA 98195-2100 USA bare@u.washington.edu Arnab Bhowmick Stephen F. Austin State University College of Forestry 1639 North Street Nacogdoches, TX 75962 USA arnabqis@hotmail.com Earl T. Birdsall Weyerhaeuser Co. PO Box 9777, WWC-IF6 Federal Way, WA 98063-9777 USA earl.birdsall@weyerhaeuser.com Tom Bobbe USDA Forest Service Remote Sensing Applications Center 2222 W. 2300 S. Salt Lake City, UT 84119 USA tbobbe@fs.fed.us Andrew Bourque Potlatch Corporation - Hybrid Poplar Program PO Box 38 Boardman, OR 97818 USA andrew.bourque@potlatchcorp.com David Briggs University of Washington College of Forest Resources Box 352100 Seattle, WA 98195-2100 USA dbriggs@u.washington.edu Anne G. Briggs PO Box 663 Issaquah, WA 98027 USA Ward Carson University of Washington Pacific Northwest Research Station Box 352100 Seattle ,WA 98195-2100 USA carsonw@u.washington.edu Woodam Chung University of Montana School of Forestry 32 Campus Drive Missoula, MT 59812 USA wchung@forestry.umt.edu Jennie L. Cornell Oregon State University Forest Engineering Operations Corvallis, OR 97331 USA bryancornell7014@msn.com Elizabeth Coulter Oregon State University Department of Forest Engineering 215 Peavy Hall Corvallis, OR 97333 USA Elizabeth.Coulter@orst.edu Christopher Davidson International Paper 1201 West Lathrop Avenue Savannah, GA 31415 USA chris.davidson@ipaper.com Weihe Guan Weyerhaeuser Co. 33405 8th Avenue S. Federal Way, WA 98003 USA weihe.guan@weyerhaeuser.com Andrew Hill University of Washington Box 352100 Seattle, WA 98195 USA adh2@u.washington.edu Bill Dyck Bill Dyck Ltd. PO Box 11236 Palm Beach, Papamoa 3003 NEW ZEALAND billdyck@xtra.co.nz Olav Albert Høibø Agricultural University of Norway Department of Forest Science P.O. Box 5044 N-1432 AAS NORWAY olav.hoibo@isf.nlh.no Stephen E. Fairweather Mason, Bruce, & Girard, Inc. 707 SW Washington St., Suite 1300 Portland, OR 97205 USA sfairweather@masonbruce.com Sean Hoyt University of Washington Box 352100 Seattle, WA 98195 USA naestyoh@u.washington.edu Dave Furtwangler Cascade Timber Consulting Inc PO Box 446 Sweet Home, OR 97386 USA dfurtwangler@cascadetimber.com Andrew Hudak US Forest Service Rocky Mountain Research Station 1221 S. Main St. Moscow, ID 83143 USA ahudak@fs.fed.us Joel Gillet Applanix Corp 85 Leek Crescent Richmond Hill, ON L4B 3B3 CANADA JGillet@applanix.com David Gilluly Weyerhaeuser Co. 33405 8th Avenue S., WWC 2B2 Federal Way, WA 98003 USA Richard A. Grotefendt University of Washington College of Forest Resources Box 352100 Seattle, WA 98195 USA grotefen@u.washington.edu Yan Jiang University of Washington Box 352100 Seattle, WA 98195 USA Dick Karsky USDA Forest Service 5785 Highway 10 West Missoula, MT 59808 USA rkarsky@fs.fed.us Phil Lacy World Forestry Center 4033 SW Canyon Road Portland, OR 97221 USA placy@worldforestry.org Bruce Larson University of British Columbia Vancouver, BC CANADA blarson@interchange.ubc.ca Stephen Lewis Timberline Forest Inventory Consultants 315-10357-109 Street Edmonton, Alberta T5J IN3 CANDA sjl@timberline.ca Hamish Marshall Oregon State University Forest Engineering Department 215 Peavy Hall Corvallis, OR 97331-5706 USA hamish.marshall@orst.edu Brett Martin Prudue University 2226 Willowbrook Dr. Apt. #192 West LaFayette, Indiana 47906 USA brettm@fnr.purdue.edu Bob McGaughey University of Washington Pacific Northwest Research Station Box 352100 Seattle, WA 98195-2100 USA mcgoy@u.washington.edu Kurt Muller Forest Technology Group 16703 SE McGillivray Blvd. Suite 215 Vancouver, WA 98683 USA kurt.muller@ftgrp.com Ewald Pertlik University of Bodenkultur Vienna Peter-Jordan-Strasse 70 Vienna A-1190 AUSTRIA ewald.pertlik@boku.ac.at Charles Peterson USDA Forest Service PNW 620 SW Main Street Suite 400 Portland, OR 97205 USA cepetersen@fs.fed.us Lester Power Weyerhaeuser Co. PO Box 9777 Federal Way, WA 98063-9777 USA lester.power@weyerhaeuser.com Steve Reutebuch University of Washington Pacific Northwest Research Station Box 352100 Seattle, WA 98195-2100 USA sreutebu@u.washington.edu Luke Rogers University of Washington Rural Technology Initiative Seattle, WA 98195 USA lwrogers@u.washington.edu Peter Schiessr University of Washington Box 352100 Seattle, WA 98195 USA schiess@u.washington.edu Glen Murphy Oregon State University Forest Engineering Department Peavy Hall 271 Corvallis, OR 97331 USA glen.murphy@orst.edu Daniel L. Schmoldt USDA/CSREES/PAS Instrumentation & Sensors Mail Stop 2220 Washington, DC 20250-2220 USA dschmoldt@reeusda.gov Megan O’Shea University of Washington College of Forest Resources Box 352100 Seattle, WA 98195 USA moshea@u.washington.edu Gerard Schreuder University of Washington College of Forest Resources Box 352100 Seattle, WA 98195 USA gsch@u.washington.edu Guofan Shao Purdue University Department of Forestry and Natural Resources Forestry Building 195 Marsteller St. West Lafayette, IN 47907-2033 USA gshao@fnr.purdue.edu Pierre Turcotte FERIC 580 Boul. St-Jean Canada Pointe-Claire, QC H9R 3J9 CANADA pierre-t@mtl.feric.ca Alex Sinclair Feric Western Division 2601 East Mall Vancouver, BC V6T 1Z4 CANADA alex-s@vcr.feric.ca Eric Turnblom University of Washington College of Forest Resources Box 352100 Seattle, WA 98195 USA ect@u.washington.edu Jack A. Sjostrom DigitShare / Sentry Dynamics, Inc. 721 Lochsa St., Suite 16 Post Falls, ID 83854 USA jsjostrom@digitshare.org Rien Visser Virginia Tech 229 Cheatham Hall Blacksburg, VA 24061 USA visser@vt.edu Derek Solmie Oregon State University Department of Forest Engineering 215 Peavy Hall Corvallis, OR 97330 USA derek.solmie@orst.edu Matt Walsh University of Washington Box 352100 Seattle, WA 98195 USA Xiping Wang University of Minnesota Duluth USDA Forest Products Laboratory Gifford Pinchot Drive Madison, WI 53705-2398 USA xwang@fs.fed.us Brant Steigers Potlatch Corporation 807 Mill Road Lewiston, ID 83501 USA brant.steigers@potlatchcorp.com Bernd-M. Straub University of Hannover Institute for Photogrammetry and GeoInformation Nienburger Strasse 1 Hannover 30167 GERMANY bernd-m.straub@ipi.uni-hannover.de Jack Ward Temperate Forest Solutions PO Box 33 Asford, WA 98304 USA tfs@rainerconneet.com Welsey Wasson VAP Timberland 695 W Satsop Rd Montesano, WA 98563 USA wsw@olynet.com Cheryl Talbert Weyerhaeuser Co. PO Box 9777 Mail Stop: CH 2D25 Federal Way, WA 98063-9777 USA cheryl.talbert@weyerhaeuser.com 158 Denise Wilson University of Washington PO Box 352500 Seattle, WA 98195-2500 USA wilson@ee.washington.edu David Yates Forest Technology Group 3950 Faber Place Dr. North Charleston, SC 29405 USA david.yates@ftgrp.com Michael G. Wing Oregon State University Forest Engineering Department 213 Peavy Hall Corvallis, OR 97333 USA michael.wing@orst.edu 159 Second International Precision Forestry Symposium Agenda Sunday, June 15, 2003 5:00 PM to 7:00 PM Reception at the UW Waterfront Activities Center Monday, June 16, 2003 7:00 AM Registration Desk Opens at Kane Hall room 220 7:00 AM Continental Breakfast 8:30 AM Welcome & Introductory Remarks - Dean B. Bruce Bare 8:45 AM Keynote Speaker -Bill Dyck Plenary Session A: Precision Operations and Equipment - Moderator, Alex Sinclair 9:05 AM Multidat and Opti-Grade: Two Innovative Solutions to Better Manage Forestry Operations presented by Pierre Turcotte, FERIC, Canada 9:25 AM A Test of the Applanix POS LS Positioning System for the Collection of Terrestrial Coordinates Under a Closed Forest Canopy - presented by Stephen E. Reutebuch and Ward W. Carson, USDA Forest Service, Pacific Northwest Research Station 9:50 AM Break & Poster Session 10:20 AM 160 Ground Navigation through the use of Inertial Measurements, a UXO Survey - presented by Joel Gillet, Applanix Corp. 10:45 AM Precision Forestry Operations and Equipment in Japan - Kazuhiro Aruga, University of Tokyo 11:10 AM Precision Forestry Applications: Use of DGPS Data to Plan and Implement Aerial Forest Operations - presented by Jennie L. Cornell Plenary Session B: Remote Sensing and Measurement of Forest Lands and Vegetation - Moderator, Tom Bobbe 11:35 AM Estimating Forest Structure Parameters Within Fort Lewis Military Reservation Using Airborne Laser Scanner (LIDAR) Data - presented by Hans-Erik Andersen, University of Washington, College of Forest Resources 12:00 PM Lunch 1:00 PM Geo-Spatial Analysis in GIS and LIDAR Remote Sensing using Component Object Modeling of Visual Basic: Application to Forest Inventory Assessment - presented by Arnab Bhowmick and Dr. Peter Siska, College of Forestry, Stephen F. Austin State University 1:25 PM Large Scale Photography Meets Rigorous Statistical Design for Monitoring Riparian Buffers and LWD - presented by Richard A. Grotefendt, University of Washington 1:50 PM Forest Canopy Models Derived from LIDAR and INSAR Data in a Pacific Northwest Conifer Forest - presented by Hans-Erik Andersen, University of Washington, College of Forest Resources 2:15 PM Fine Tuning Forest Change Detections with a Combined Accuracy Index - presented by Guofan Shao, Department of Forestry and Natural Resources, Purdue University 161 2:40 PM Break & Poster Session 3:10 PM Automatic Extraction of Trees From Height Data Using Scale Space and Snakes - presented by Bernd-M. Straub, Institute for Photogrammetry and GeoInformation, Germany 3:35 PM RFID Research-presented by Sean Hoyt 4:05 PM Sean Hoyt Tree Tour 4:30 PM Adjourn Tuesday, June 17, 2003 7:00 AM Registration Desk Opens at Kane Hall room 220 7:00 AM Continental Breakfast 8:05 AM Keynote Speaker - Dan Schmoldt Plenary Session C: Terrestrial Sensing, Measurement and Monitoring Moderator, Steve Reutebuch 8:30 AM Value Maximization Software-Extracting the Most from the Forest Rersource - presented by Hamish Marshall and Graham West 8:55 AM Costs and Benefits of Four Procedures for Scanning on Mechanical Processors - presented by Glen E. Murphy and Hamish Marshall 162 9:20 AM Evaluation of Small-diameter Timber for Value-added Manufacturing: A Stress Wave Approach - presented by Xiping Wang and Robert J. Ross 9:45 AM Break & Poster Session 10:15 AM Aroma Tagging and Electronic Nose Technology for Tracking Log and Wood Products: Early Experience - presented by Glen Murphy Plenary Session D: Design Tools and Decision Support Systems - Moderator, Glen Murphy 10:40 AM Modeling Steep Terrain Harvesting Risks using GIS - presented by Jeffrey Adams, Rien Visser, and Steve Prisley, Department of Forestry, Virginia Tech 11:05 AM Use of the Analytic Hierarchy Process to Compare Disparate Data and Set Priorities presented by Elizabeth Coulter and John Sessions, Department of Forest Engineering, Oregon State University 11:30 AM Use of Spatially Explicit Inventory Data for Forest Level Decisions - presented by Bruce C. Larson, University of British Columbia, Faculty of Forestry 11:55 AM Lunch 1:00 PM Elements of Hierarchical Planning on Forestry: A Focus on the Mathematical Model presented by Sam Pittman, University of Washington, College of Forest Resources 1:25 PM Update Strategies for Stand-Based Forest Inventories - presented by Stephen E. Fairweather, 163 Mason, Bruce, & Girard, Inc. 1:50 PM A New Precision Forest Road Design and Visualization Tool: PEGGER - presented by Luke Rogers, Geographic Information Scientist; UW Rural Technology Initiative 2:15 PM Harvest Scheduling with Aggregation Adjacent Constraint: A Threshold Acceptance Approach presented by Hamish Marshall, Graduate Student, Oregon State University 2:40 PM Break (Poster Session Breakdown) 3:10 PM Optimizing Road Network Location in Forested Landscapes - presented by Michael G. Wing, John Sessions, and Elizabeth Coulter, Oregon State University 3:35 PM Comparing Forest Area Measurement Techniques - presented by Derek Solmie, Department of Forest Engineering, College of Forestry, Oregon State University 4:00 PM Closing Remarks 4:25 PM Adjourn Wednesday, June 18, 2003 - Field trip CANCELED 164 165