Loading...
HomeMy WebLinkAboutAlaskan Coal-Resources & Development Constraints 1982Alaska Energy Authority LIBRARY COPY ANL/LRP-18 ALASKAN COAL: RESOURCES AND DEVELOPMENTAL CONSTRAINTS LAND RECLAMATION PROGRAM PV We Mk Operated for U. S. DEPARTMENT OF ENERGY Pv cola MN GLAS under Contract W-31-109-ENG-38 aie acs ee, rN fered The facilities of Argonne National Laboratory are owned by the United States Government. Under the terms of a contract (W-31-109-Eng-38) among the U. S. Department of Energy, Argonne Universities Association and The University of Chicago, the University employs the staff and operates the Laboratory in accordance with policies and programs formulated, approved and reviewed by the Association. MEMBERS OF ARGONNE UNIVERSITIES ASSOCIATION The University of Arizona The University of Kansas The Ohio State University Carnegie-Mellon University Kansas State University Ohio University Case Western Reserve University Loyola University of Chicago The Pennsylvania State University The University of Chicago Marquette University Purdue University University of Cincinnati The University of Michigan Saint Louis University Illinois Institute of Technology Michigan State University Southern Illinois University University of Illinois University of Minnesota The University of Texas at Austin Indiana University University of Missouri Washington University The University of Iowa Northwestern University Wayne State University Iowa State University University of Notre Dame The University of Wisconsin-Madison |OTICE This rep ort was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, com- pleteness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific com- mercial product, process, or service by trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof. Printed in the United States of America Available from National Technical Information Service U. S. Department of Commerce 5285 Port Royal Road Springfield, VA 22161 NTIS price codes Printed copy: A07 Microfiche copy: AOI LOA O10 Distribution Category: Coal Mining (UC-88) ANL/LRP-18 ARGONNE NATIONAL LABORATORY 9700 South Cass Avenue Argonne, Illinois 60439 ALASKAN COAL: RESOURCES AND DEVELOPMENTAL CONSTRAINTS by Dorland E. Edgar, Lawrence J. Onesti,* and Gary M. Kaszynskit Land Reclamation Program March 1982 work sponsored by U.S. DEPARTMENT OF ENERGY Office of Energy Research *associate Professor of Geography, Indiana University *ESCOR, Inc. PREF ACE This study was performed as part of the Argonne National Laboratory Land Reclamation Program, which is sponsored by the Department of Energy's Office of Energy Research. The program is a joint effort conducted by Ar- gonne's Energy and Environmental Systems Division and the Division of Envi- ronmental Impact Studies. The Land Reclamation Program, as the lead DOE program for reclamation research, conducts basic and applied research into the physical, ecological, and economic problems of land reclamation related to the mining of coal. The work is aimed at developing energy-efficient and cost-effective techniques for reclaiming and rehabilitating mined land to productive end uses. To achieve this goal, the Program has established integrated research and de- velopment projects focused on near- and long-term reclamation problems in all major U.S. coal resource regions. These research sites have been established to address both regional and site-specific problems. The activities of the Land Reclamation Program involve close cooperation with industry and the academic community and focus on establishing a comprehensive field and labor- atory effort. At five of its research sites, the Program has developed co- operative working arrangements with the operating coal companies. A primary goal of the Program has been close cooperation with related research projects at academic institutions and other agencies, in order to transfer pertinent information and avoid duplication of effort. Alaska has a tremendous quantity of coal situated in a variety of en- vironmental conditions. The only commercial coal mine today is a surface operation producing approximately 700,000 t/yr, but recent events indicate that production from this and new mines should be significantly greater dur- ing the next several years in response to the developing export market in the Far East. Mined-land reclamation experience in Alaska, however, is limited essentially to the one active mine, and the near- and long-term effects of surface mining under arctic and subarctic conditions are largely unknown. This report presents an overview of the state's coal resources and the envi- ronmental and socioeconomic factors that are related to past, present, and future use of these resources. Several areas are identified where research is needed to ensure that anticipated mining and reclamation endeavors will be planned and conducted in an environmentally sound yet cost-effective manner. Ralph P. Carter, Director Land Reclamation Program iii CONTENTS ACKNOWLEDGMENTS .... ee ee eee ABSTRACT . . 1 6 1 ee we ee we ww 1 2 INTRODUCTION . . 2. 2. 2 2 ee ee ee ENVIRONMENTAL SETTING. ....... 2.1 Ze Z 2.3 2.4 2.3 2.6 General Physical Features .. . Physiography and Geology . 2.2.1 Arctic Coastal Plain .. 2.2.2 Rocky Mountains ..... 2.2.3 Intermontane Plateaus . . 2.2.4 Pacific Mountains .... Climate «4 6 sw ee ee ew x 2.3.1 Arctic Zone. ...... 2.3.2 Continental Zone .... 2.3.3 Transition Zone ..... 2.3.4 Maritime Zone. ..... Permafrost} s «sts « @ ss * « 2.4.1 2.4.2 Occurrence and Formation Effects of Permafrost . . Water Resources and Hydrology . Surface Water ...... Ground Water ...... Water Quality ...... Water Use... . 2. e « « Ecoregions . . 1... sess es Arctic Tundra. ..... Bering Tundra. ..... Brooks Range ......- Yukon Forest ...... Alaska Range ...... Pacific Forest ..... POPULATION AND ECONOMY ....... 3.1 3.2 Population. ....... Economy .... +++ +e eee LAND OWNERSHIP, JURISDICTION, AND USE FRPP UPWNHe Early Federal Influence in Alaska The Alaska Statehood Act .... The Alaska Native Claims Settlement Act The Alaska National Interest Lands Conservation Act Summary . 2. 2 se ee eee ees ix au wu Lt i 12 14 14 15 21 21 21 21 25 26 26 29 33 33 34 35 a3 37 37 37 38 39 ao 39 49 49 50 52 55 56 CONTENTS (Contd.) 5 COAL RESOURCES AND DEVELOPMENT «2. «© © sss es vv ess so vee 63 Ss) Coal) Resources’ ©) a/c jo| 1 | i |i (2) je) ie el ee) [ol @) lie fa) ol) | ve | sie te) xe 63 Sele | (Rank, /and)iGriade) | 3) fir fs) |e) ee) et le) |e fe) ©) el) eis ie) 64 5.1.2 Eetimated Goal Besources sists cts ve seawenass 64 5.2. Goal Basins ati Geology - - eee etree t eee REE SD 68 Arctic Region je) 3 =) |) a |e) is! te) tol) @) le) |e fet) oe) tor) et) he) 70 faterion MWegiom ws is cea eee hr 70 South-Central Region’) < |. i) |e) +) 5 6 | (6) © Ws) so) mi bo) 71 Other Coal Areas . 2. 1 2 es 2 oe ee ee we ee wo eo 72 Unun NNNN FwWNnre 5.3 RBistory of Alaskan Coml Mining «1.12 esses menue eww wo 73 Dede ds) Pre-WOrLa=War GEE) ie) ea |e) lie! ve rod) © le) te |e! foi) is ell tle ite els 73 Dese2) \Post-World-War, Ec) a) le) ie) ele 2) ee) @ sis) © fel le lie lel 76 5.4 Prospects for Alaskan Coal Development . ........+ eee 84 5.4.1 Constraints on Coal Development . ........-s-.e-e 84 5.4.2 Coal Mining Forecasts ....56s6sse ees eevee ses 93 5.4.3 Hecent Activities 46. sc st tee hee 94 S544) (Summary. fa) io lier io) 1s) tot (ele |e) fo et) oe) is) is) ot) I) eo) |e) ts) 99 6 CONCLUSIONS AND RECOMMENDATIONS . . . 2... + + «© © © © © © we ee ©6101 6.1) | Conciliesions! |") fo) 6) |e) = fet fe) fe) er fe) ie St) fre) lo toca) ey it fe) EO 6.2) | Recommend aleions |e) 6) |) 1) fe) te |'s) |e) lel) of el) io) ret =) es) er |e) ei) fe te) LOG REFERENCES . . 2.2.2 es eee eet eeveee eevee eevee eee 109 APPENDIX: COAL RESOURCE CLASSIFICATION SYSTEM OF THE U.S. BUREAU OF MINES AND U.S. GEOLOGICAL SURVEY ..... +++ e+e e+e 121 FIGURES Alaska and the Conterminous 48 States Shown at the Same Scale ... Major Physical and Geographic Features of Alaska... ... + ees Physiographic Map of Alaska .... esse eeee esse vveee wo on wv Physiographic Provinces of Alaska . .. 1. ee ee eee ee eee Geologic Map of Alaska . . 2... 2 eee eee eee ese vevses 10 Glimatic Zones of Allaska)|< [5/2 ie) a\e)@ fe) ls) a) 2) 1\8 & le (9) @ | «1 |e 15 Mean Annual Precipitation Distribution . ..... 6 2 eee ee ee 16 Average Annual Snowfall Distribution ...... +. +e 22 ee eee 7 wo monn DU F WH Mean Daily Maximum and Minimum Temperature Distributions for Jandanry | js) | = i) |e fe) |e i | le) wis) 4) 2 [el @ li fe] sl ie) © |e ie 9/9 |f 18 vi 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 w N DW FIGURES (Contd.) Mean Daily Maximum and Minimum Temperature Distributions for July 2 6 ew ew wwe we we ee we we ee ww ww we we we ewe Hours of Sunlight and Twilight for Different Latitudes and Months of the Year . 2. 1 1 ee ee ee ee eee ee we eee Distributionof Permafrost . . «ss « ms 6 © wow ee ww ® Occurrence of Taliks in Relation to the Active Layer, Supra- permafrost Zone, Permafrost Table, and Permafrost ........ Effect of Surface Features on the Distribution of Permafrost Zone . 2 6 © © © © we ew ww te we th ww wh hw eh we wh wht th ww Estimated Mean Annual Runoff... .. 2... ee ee ee ee eee Estimated Mean Annual Peak Runoff ......4..-+655+ 22 eee Generalized Availability of Ground Water .........42.046- Ecoregions of Alaska . . 1. 6 6 ee ee wee ee ee we ee ee Major Components of Alaska's Transportation System........ Native Villages and Regional Corporations ........24.2e-6 Areas Withdrawn by the Alaska National Interest Lands Conservation Mee OE 1 wwe he ee he ee Comparison of Moist, Mineral-Matter-Free-Basis Heat Values and Proximate Analysis of Coal of Different Rank. .......4-24-. Location of Principal Coal Deposits in Alaska .......-2+.e-. Coal Fields of the Cook Inlet Basin ...... 2. eee ee eee Healy Creek and Lignite Creek Coal Fields and Locations of Current Mining Operations ..... 2. ee eee ee eee eee TABLES Major Alaskan Rivers, Runoff Rates, and Drainage Areas ...... Estimated Water Use by Subregion... .. 2... 2.2 ee ee eee Alaskan Communities with 1976 Estimated Population Greater than 1000 . 2... 2 we ee we we ee we te we ew te we ew ws Historical Production of Minerals in Alaska ......2 2-2 eee Value of Alaskan Mineral Production from 1959-1979 ........ Total Alaskan Wages and Employment by Industry, 1975-79 ..... Comparative Costs of Food at Home for a Week and Energy Costs in December 1979 for Anchorage and Other Alaskan Communities ... . Comparison of Alaskan and U.S. Per-Capita Income, with Adjustments for Cost-of-Living Differences . . .. 2. 2. 2 se ee eee ee eee Federal Withdrawals of Land before 1969 ........4+24+4+2468-2 vii 19 20 22 23 24 28 30 32 36 47 54 60 65 66 72 80 27 34 40 42 43 44 45 46 51 10 ll 12 13 14 15 16 17 18 19 20 TABLES (Contd.) Summary of Land Distribution Specified by ANCSA ....... Land Conservation Units and Acreages Established by the Alaska National Interest Lands Conservation Act of 1980 ........ Selected Publications on Geology and Coal Resources of Alaskan Comi-Bearing Arwas . se ese ee thee eee ee hh Quality Characteristics of Alaskan Coals ...... eee eee Comparison of Average Characteristics of Coal from Major Coal- Bearing Regions of the Conterminous United States ....... Geologic Characteristics of Alaskan Coal Deposits ....... Significant Events in Coal Development and Production in Alaska Revegetation Data for the Usibelli Coal Mine, 1977-1981 .... Socioeconomic Conditions in Alaskan Coal-Bearing Regions .... Environmental Conditions in Alaskan Coal-Bearing Regions .... Coal Lease Areas on State Land as of November 1980, Tabulated by Location and Lessee «ces ee ewe eee he viii 55 57 68 69 69 74 77 82 87 90 96 ACKNOWLEDGMENTS Numerous individuals provided valuable assistance during the course of this study and preparation of this report. Without their help, the content of the report would have been less accurate and comprehensive. Catherine A. Boyd and Lynell K. Daniels collected much of the literature during the ini- tial phase of the project. Margaret J. Mattson, Librarian at the Alaska Field Operation Center of the U.S. Bureau of Mines in Juneau, made special efforts to provide copies of reports, many of which were out of print. We thank the many individuals in Alaska's Department of Commerce and Economic Development, Department of Environmental Conservation, Department of Natural Resources, House Research Agency, and Legislative Affairs Agency; as well as those in the Arctic Environmental Information and Data Center, Institute of Water Resources, and School of Mineral Industry of the University of Alaska. The U.S. Army Corps of Engineers Cold Regions Research and Engineering Labor- atory, U.S. Bureau of Mines, and U.S. Geological Survey were consistently co- operative and always willing to discuss their current activities and to pro- vide their reports and documents. We acknowledge the special assistance pro- vided by Susan E. Brody, Issues Analyst, Alaska House Research Agency; Steve W. Denton, Chief Engineer, and Charles Boddy, Reclamation Manager, Usibelli Coal Mine, Inc.; Richard H. Eakins, Director, Alaska Division of Economic Enterprise; Gene Kennedy, Director, Alaska Legislative Information Office, Washington, D.C.; Noel W. Kirshenbaum, Manager of Mineral Projects, Placer Amex Inc.; Jay D. McKendrick, Associate Professor, Agricultural Experiment Station, University of Alaska; and P. Dharma Rao, Professor, Mineral Industry Research Laboratory, University of Alaska. These individuals gave freely of their time for discussions with the senior author and provided information that significantly improved the value of this report. Portions of this manuscript were reviewed by Ralph P. Carter and Ray R. Hinchman of Argonne's Land Reclamation Program and by Noel W. Kirschenbaum of Placer Amex Inc. The entire manuscript was reviewed by Steve W. Denton, Usibelli Coal Mine., Inc.; Jay D. McKendrick, University of Alaska; James R. Rastorfer, Associate Professor, Department of Biological Sciences, Chicago State University; John Sims, Director, Alaska Office of Minerals Development; and Stanley D. Zellmer, Land Reclamation Program. Their comments and sugges-— tions significantly improved the quality of this report. The text was typed by Barbara A. Rogowski, illustrations were drafted by Linda M. Samek, and the readability of the report was significantly improved by the editing of Charles A. Malefyt. ix ALASKAN COAL: RESOURCES AND DEVELOPMENTAL CONSTRAINTS D. E. Edgar, L. J. Onesti, and G. M. Kaszynski ABSTRACT Alaska has great physical diversity, unparalleled scenic beauty, abundant natural resources, and environmental conditions unlike those of any other state. The state's sparse population is concentrated principally in Anchorage, Fairbanks, and Juneau, and there is a significant native populace living in isolated small villages. The economy is based primarily on fishing, timber, and petroleum indus- tries; wages and cost of goods and services are above the national average. Environmental conditions exhibit great variability from one region of the state to another. Use of natural resources in the past has been hampered by unsettled land ownership; future development will be constrained by the severe climate, widespread permafrost, and inadequate transportation facilities. Alaska has perhaps as much coal as that found in all of the 48 conterminous states. Most of the coal deposits are located in the North Slope Basin and associated strata extending beneath the Chukchi Sea, the Nenana Basin of the interior, and the Cook Inlet Basin in the south. Most Alaskan coal is subbituminous to bituminous with relatively high moisture and ash content, moderate heat value, and very low sulfur content; many deposits are well suited for surface mining. At present, only one commercial coal mine is operating in Alaska; it is situated in the Nenana Basin about 160 km south of Fairbanks. Mining should increase dramatically in the next few years, however, in re- sponse to a developing Far East market. Near-term increases in production will probably come from the existing mine and from new mines in the interior and south-central regions of the state. Because there is limited experience in Alaskan surface coal mining and reclamation, research should be con- ducted in order to develop cost-effective mining and recla- mation technologies that minimize or prevent adverse impacts on Alaska's unique ecosystems. 1 INTRODUCTION Alaska is unique among the fifty states. It is at once the largest, least populated, most remote, and least developed of all the states. No other state contains such a diversity of physical environments and natural features, and there is an abundance of wildlife, wilderness areas, and spec- tacular scenery. In addition, this huge state, occupying one-sixth of the total U.S. area, ranges in elevation from sea level to 6194 m (20,320 ft) atop Mt. McKinley, the highest point in North America. Average annual pre- cipitation ranges from less than 13 cm (5 in.) on the North Slope to more than 760 cm (300 in.) in portions of the panhandle in the southeast. Varia- tions in other environmental characteristics are equally great. Alaska has vast areas underlain by permafrost, 4 x 10® ha (11 x 10© acres) of glacial ice, a coastline longer than that of the conterminous 48 states, the nation's sixth longest river, and more than 40 active volcanoes (Ritter, 1979). In addition to this abundance of land, water, and wildlife resources, Alaska has a vast wealth of mineral and energy resources. Metallic minerals of historical economic significance include gold, silver, copper, tin, mer- cury, and platinum. Mineral exploration has increased in recent years, but the state's full mineral potential is still unknown due to incomplete geo- logic information. The current production of crude oil and natural gas is centered at Cook Inlet and Prudhoe Bay, while several other areas appear to have good potential for future developments. Total recoverable oil and gas resources are reported to range from 4.33 to 8.59 x 1019 bbi and from 3.4 to 11.6 x 1012 m3 (1.21 to 4.10 x 1014 £3), respectively (McConkey et al., 1977a). There are also huge quantities of coal distributed throughout the state. At present the one active coal mine of significance produces approxi- mately 700,000 metric tons (775,000 tons) per year. However, total coal re- sources have been estimated at 1.69 to 4.53 x 1012 metric tons. Although these estimates must be considered only approximate, they represent 39 to 63% of the total estimated coal resources of the United States and from 10 to 23% of those of the world (McConkey et al., 1977a). In addition to its huge fos- sil fuel resources, Alaska has a tremendous hydroelectric potential, almost 200,000 ha classified as known geothermal resource areas, and numerous sedi- mentary basins that may have favorable uranium potential (Bottge, 1978). The full energy and mineral resource potential of the state will not be known un- til prospecting activities are complete. An important factor that has significantly influenced resource devel- opment within the state is that of land ownership and management jurisdic— tion. Before Alaska became the 49th state in 1959, approximately 99% of Alaskan land was federally owned. The Alaska Statehood Act of 1958 provided for the state to select 4.15 x 107 ha of land for entitlement by 1985. After considerable debate over land claims made by Alaskan natives, Congress passed the Alaska Native Claims Settlement Act (ANCSA) in 1971. This Act provided for monetary settlement of nearly $1 billion ($109) to be paid over a period of years and entitled native groups and organizations to select a total of nearly 1.78 x 107 ha of land for fee-simple title transfer (Federal-State Land Use Planning Commission for Alaska, 1977). Sections 17(d)(1) and (2) of ANCSA provided a means of withdrawing, classifying, and preserving millions of hectares of national-interest lands under federal ownership. The majority of these lands were to be ultimately designated for conservation through the national system of parks, forests, wild and scenic rivers, and wildlife refuges. Additional areas were withdrawn into federal ownership in 1978 when President Carter implemented the Antiquities Act of 1906 and designated 2.3 x 107 ha as national monuments. After all land selections and transfers are complete, it is anticipated that nearly 60% of the state will remain under federal control and almost 30% will be owned by the state government. Ap- proximately 10% of the state will be held by native corporations and associ- ated organizations and, depending in part upon quantities subsequently trans- ferred from state ownership, only about 1% will be held privately by other groups or individuals (McConkey et al., 1977a). Following a lengthy con- gressional and public debate over which lands would be retained under federal ownership as specified by Section 17(d)(2) of ANCSA, Congress in December 1980 passed the Alaska National Interest Lands Conservation Act (P.L. 96- 487), which incorporated approximately 4.3 x 107 ha of public lands into conservation units. Although passage of this Act is a major step toward set- tlement of the land ownership question, several complexities remain to be resolved. This report presents a brief review of the nature and quantity of Alaskan coal resources and the environmental characteristics and other fac- tors that influence existing and future coal development within the state. Research needs related to coal mining and reclamation are identified, and recommendations for future research are made. Based only upon the magnitude of the resource base, potential future coal utilization in Alaska is almost unlimited. Although Alaska has a history of coal mining, levels of extrac- tion have been extremely low in comparison to potential production. Future mining will be determined by a combination of environmental considerations, land ownership and usage classification, and numerous economic, techno- logical, social, and political factors; therefore, it is impossible to pre- dict precisely what developments will occur. In recent years, Alaska has become a focal point for opposing philos- ophies of natural resource management; i.e., strict conservation and preser- vation of this "last frontier" and wilderness area vs. full economic exploi- tation of the state's mineral and energy wealth. This dispute has centered on debate over the aforementioned "d-2 issue" arising from section 17(d)(2) of ANCSA. Undoubtedly, the ultimate disposition of this controversy will be a compromise between these two extremes, with resource development being permitted and even encouraged in certain geographic areas. Because such de- velopment would improve and expand the economic base of the state and its residents, the expansion of mineral and energy industries generally tends to be viewed favorably by the state government, provided such activities are conducted in an environmentally sound manner. This position is indicated by the fact that much of the land selected by the state and native corporations has a high potential for natural resource development. Furthermore, the Power Plant and Industrial Fuel Use Act (P.L. 95-620) of 1978, as well as recent federal energy policies, stresses increased utili- zation of abundant coal resources through both direct combustion and syn- thetic fuels production to replace some of the finite supplies of petroleum and natural gas. Considering that Alaska will undoubtedly experience indus- trial and economic growth in the future, and because the state is thought to contain almost two-thirds of the nation's total coal resources, all indica- tions are that future coal production is almost certain to be significantly higher than in the past. Scientific investigations have been conducted in Alaska since the 1800s, but the tempo of research and geological exploration has increased since the mid-1970s. However, due to the large land area and the diversity of conditions within the state, additional basic scientific information is badly needed to aid in understanding the specific effects of perturbations upon existing ecologic systems. Only a limited amount of research has been conducted on surface coal mining and reclamation in Alaska, and much of the information gained from the rest of the country is not directly applicable because of the large differences in natural settings. If surface coal mining in Alaska does indeed increase significantly in the near future, basic data and scientific results will be required to plan mining and reclamation activities so as to minimize adverse environmental consequences. Because environmental investigations must be conducted over a period of time adequate to observe natural temporal variations, it would be prudent to begin research activities at an early date to ensure proper plan- ning and the collection of representative data rather than beginning a con- centrated research effort only after escalated mining activity is imminent or under way. It is for these reasons that research needs and recommendations are presented in the concluding section of this report. 2 ENVIRONMENTAL SETTING Because environmental considerations will strongly influence future economic growth and natural resource development, a brief description of the environmental conditions is presented as background information prior to the detailed discussion of the state's coal resources and as a framework for in- terpreting the identified research needs related to coal extraction and land reclamation. The following discussion is a basic introduction to the general environmental setting of Alaska. 2.1 GENERAL PHYSICAL FEATURES Alaska occupies about 1.52 x 108 ha (3.75 x 108 acres), or approxi- mately 1.52 x 10© km? (586,000 mi2) of land, inland waters, and glacial ice. With one-sixth the total area of the U.S., Alaska is larger than Texas, Cali- fornia, and Montana combined. When a map of Alaska is superimposed upon a map of the conterminous 48 states of the same scale (Figure 1), the state stretches from the Atlantic to the Pacific coasts and from the Canadian to Mexican borders. Most of this vast area is in its natural state and displays few signs of human influence. Fig. 1 Alaska and the Conterminous 48 States Shown at the Same Scale As shown in Figure 2, Alaska is bounded on the north by the Arctic Ocean; the Chukchi and Bering seas lie to the west, and the Gulf of Alaska (a portion of the Pacific Ocean) borders the state on the south. The Yukon, one of the longest rivers of the world, flows almost 3060 km (1900 mi) from its headwaters in Canada to the Bering Sea; the Kuskokwim River is more than 1250 km long. The state has more than three million lakes with surface areas greater than 8 ha (20 acres) each, 11 of the 20 highest mountains in North America, and numerous glaciers, one of which (Malaspina) is as large as the state of Rhode Island. 2.2 PHYSIOGRAPHY AND GEOLOGY The topographic and physiographic characteristics of Alaska are highly diverse, ranging from rugged mountain chains to broad, low wetlands and the flat coastal plain adjacent to the Arctic Ocean (Figures 3 and 4). The land- scape can be divided into four broad physiographic divisions based upon large-scale landform characteristics and geologic origins: (1) Interior (Arctic) Plains, (2) Rocky Mountain System, (3) Intermontane Plateaus, and (4) Pacific Mountain System (Wahrhaftig, 1965). Each of these divisions is a northern extension of the same physiographic division in the conterminous 48 states and western Canada. Thus the Arctic Plains division is physiographi- cally analogous to the Great Plains of the western states, and the Intermon- tane Plateaus division correlates roughly with the landform region separating the Rocky Mountains from the Pacific Coast Mountain chain in the conterminous 48 states and western Canada. Although major differences do exist, the geo- logic history of each of these divisions in Alaska is broadly similar to that of its southern counterpart. Based upon a specific, characteristic topography and geologic struc- ture, each of the four physiographic divisions can be subdivided into a num- ber of provinces and each province can be further divided into smaller sec- tions with homogeneous landform characteristics. Wahrhaftig (1965) classifed the state into 12 distinct provinces and 60 sections (Figure 4). When the physiographic boundaries are compared with the areal distribution of rock type and structural features shown on a generalized geologic map of the state (Figure 5), many similarities in pattern become apparent. This is to be ex- pected because topographic form is an expression of geologic conditions and history as modified by geomorphic processes. 2.2.1 Arctic Coastal Plain The Arctic Plain is a smooth coastal plain of very gentle slope rising from sea level to a maximum altitude of 180 m (600 ft) at the southern mar- gin. Locally scattered groups of low hills, some sand dunes, and pingos (ice-cored hills) provide some relief above the level of the surface. Due to the presence of continuous permafrost, the area is very poorly drained and becomes quite marshy during summer thawing, with numerous shallow seasonal lakes dotting the surface. Also present are patterned ground features such as stone polygons, which are indicative of frost action and cold-climate pro- cesses. The coastal plain is crossed by several rivers, of which the Col- ville is the largest, originating in the foothills and mountains to the WW Approximate Location of Trans-Alaska Pipeline NPR-A_ National Petroleum Reserve ~Alaska Bettles oe y yuxon Y u eo Fairbanks \ i ®, \ We Kilometers © 100 200 Se ° 100 Miles Fig. 2 Major Physical and Geographic Features of Alaska Oceg a Point Borrow Gulf of Alaska ° Zam se Pribilot i * bn > Islonds FS Se Ocean Sutian Isiands Fig. 3 Physiographic Map of Alaska (from Natural Regions of the United States and Canada, by Charles B. Hunt; W.H. Freeman and Company, Copyright 1974) (mae “ Tae TT Tm de “| \ screen \ ror Hors. EXPLANATION High rugged mountains MAJOR NORTH AMERICAN PHYSIOGRAPHIC Summits more than 5,000 feet DIVISIONS IN ALASKA , Zw ow Low mountains, plateaus, and highlands we of generally rolling topography Summits 1000-5000 feet Generally leas than Lum) feet 4 | Boundaries between physiographic units © $0 100 180-200 ues cs sr martuew + [eee eee Aa © 50 100 150 200 KILOMETERS se xe oo | ~ 4 ° manor rans ~ 2 we Fig. 4 Physiographic Provinces of Alaska (from Wahrhaftig, 1965) TTI TTT wl TET ETT LTT TTT rT IL hr er ry rrr yt ohannow EXPLANATION SEDIMENTARY AND METAMORPHIC ROCKS © $0 100 180 200miLES. bp © 50 100 150 200 KILOMETERS Quaternary surficial deposits, allu- vium, glacial debris, eolian sand and silt Mesozoic sandstone and shale, marine and nonmarine;includes ‘some metamorphic rocks \\\ Paleozoic and Precambrian sand- stone, shale, limestone; mostly marine; includes some lower Mesozoic rocks Paleozoic and Precambrian meta- morphic rocks, schist, gneiss, and 80 forth; mainly Paleozoic IGNEOUS ROCKS Quaternary and Tertiary voleanic Tertiary and Upper Cretaceous in. trusive rocks: mainly granitic; ultramafic and rocks; mainly” granitic Unmapped o Contact Dnshed where inferred Ls Fig. 5 Geologic Map of Alaska (from Cobb, 1974c) OT fi south. The region is underlain by 15 to 45 m of unconsolidated Quaternary sediments, some lower Tertiary deposits, and nearly horizontal Cretaceous rocks containing several coal units (Wahrhaftig, 1965). 2.2.2 Rocky Mountains The Rocky Mountain System consists of the Arctic Foothills and Arctic Mountains provinces. The foothills province has a rolling upland form with altitudes ranging from 180 m in the north to 1067 m near the mountains. East-west trending ridges, knobs, mesas, irregular buttes, and undulating tundra plains are present. Most streams flow swiftly in braided channels across broad gravel flats. Cretaceous sedimentary rocks are deformed into long, linear folds. Toward the mountains, the structure becomes more complex with Devonian to Cretaceous clastics and carbonates being tightly folded, in- truded, and overthrust northward (Cobb, 1974c; Wahrhaftig, 1965). The Arctic Mountains province comprises the Brooks Range (altitudes ranging from 1200 to over 2440 m) the DeLong Mountains, the Baird Mountains, and some intervening low-land areas. This mountain chain extends the entire width of the state and forms the drainage divide between the Arctic Ocean and Bering Sea. Topographically, this land mass is generally very rugged with narrow, steep ridges, high relief, and features reflecting Pleistocene alpine glaciation. Cirque glaciers are present in the Brooks Range, and there is continuous permafrost throughout the entire region. Surface drainage occurs in steep-sided valleys and ravines near headwaters; downstream segments often flow in broad, U-shaped valleys that were enlarged during Pleistocene glaci- ation. Geologically, the Arctic Mountains province is complex, with igneous, sedimentary, and metamorphic rocks all present. Devonian to Cretaceous sedi- mentary rocks, intruded with massive diabase sills, are found in the DeLong Mountains, and folded Paleozoic limestones and volcanics are present in the Bairds. The Brooks Range is predominantly Paleozoic limestone, shale, quartzite, slate, and schist that exhibit significant structural deformation. Overturned and faulted folds and giant nappes thrust to the north are indica- tive of the mountain-building processes. Granitic intrusions also underlie portions of the region (Cobb, 1974c; Wahrhaftig, 1965). 2.2.3 Intermontane Plateaus The Intermontane Plateaus division includes essentially all of inte- rior Alaska between the Brooks and Alaska ranges. This vast area consists of a number of low mountain ranges, rolling uplands, and lowlands covered with alluvium. The area generally slopes downward to the west, with altitudes of upland areas rarely exceeding 1830 m in the east and generally below 915 m in the west. Most of this region is within the drainage basin of the Yukon River. Most of the remaining area not drained by the Yukon is within the watershed of the Kuskokwim River to the south. Both the Yukon and Kuskokwim cross a large, low deltaic area containing numerous swamps before entering the Bering Sea. This area includes the lower 240 km of each river; tide- waters extend approximately 160 km upstream from the sea. 12 Upstream, the gradient of the Yukon remains low, and Fairbanks, lo- cated several hundred kilometers distant on the Tanana, lies only 137 m above sea level (Hunt, 1967). Along this middle reach, the Yukon and its larger tributaries have incised into a dissected plateau surface that is 300 to 600 m above river level. This surface generally slopes upward away from the river to both the north and south toward bordering mountain ranges. Several large structural basins filled with alluvium are also scattered throughout the region. One of these, the Yukon Flats (see Figure 4), is about 320 km long and covers approximately 25,900 km2 (Hunt, 1967; Thornbury, 1965). The geologic features of the intermediate plateaus are also complex and variable (see Figure 5). Most of the lowland areas of the interior are blanketed with 60 m or more of Quaternary clays, silts, sands, and gravels, much of which is glacial or eolian in origin (Péwe, 1975). Large areas of unconsolidated alluvial deposits flank the Yukon and its principal tribu- taries; much of the covered bedrock consists of Paleozoic sedimentary units and volcanics of varied lithologies. Locally, highly metamorphosed schists and gneisses, altered volcanics and limestones extend vertically through the younger deposits. The central highlands contain a mixture of folded and faulted Paleozoic and Mesozoic sediments (mostly shales, sandstones, and con- glomerates) and numerous isolated hills and ridges of igneous rock. Granitic intrusives and contact metamorphic facies are also present. Geologic varia- bility is indicated by the presence of volcanic extrusives of several geo- logic ages, hot springs, large areas of loess deposits, and beds of Tertiary lignites (Stearns, 1965). To the west, the Seward Peninsula is underlain by Paleozoic and Precambrian metamorphics (schist, slate, and marble) and by various sedimentary deposits of the same era. Smaller areas contain Quater- nary basalts, volcanic ash deposits, and unconsolidated alluvial and eolian sediments (see Cobb, 1974c). 2.2.4 Pacific Mountains The Pacific Mountain System is an arcuate zone of high mountains rim- ming the Pacific Ocean and consisting generally of two mountain chains sepa- rated by a belt of lowlands. In Alaska, the northern mountains include the Aleutian Range, the Alaska Range, and the Coast mountains in the southeast; the southern range consists of the Kodiak, Kenai-Chugach, Baranof, and Prince of Wales mountains. The two mountainous zones are separated by the Cook Inlet-Susitna, Copper River, and Kupreanof lowlands (see Figure 4). The two mountain chains merge to form the St. Elias Mountains, possibly the most spectacular mountains in North America. Relief here is extreme, with peaks that rise to nearly 6100 m and lowlands that are near sea level (Wahrhaftig, 1965). The Aleutian Range is composed primarily of east-west ridges 300 to 1220 m high and includes volcanoes 1370 to 2600 m in altitude, many of which are classified as active. To the north and east, the Alaska Range forms an arc more than 960 km long and up to 190 km wide. Parallel, rugged ridges and peaks generally range between 1830 and 2890 m; Mt. McKinley, however, exceeds 6100 m and is the highest peak in North America. The Coast Mountains, also called the Boundary Ranges, form an upland 1520 to 2130 m high in the pan- handle of southeastern Alaska. Rounded summits on the ridges have been modi- fied by glacial action, and many arétes and glacial horns rise an additional 13) 300 to 1520 m above the ridge elevations. The remaining mountains that make up the southern range of the Pacific Mountain System are equally spectacular in their scenery. All were extensively glaciated during the Pleistocene to add to the angularity of ridges and peaks. Many mountain groups rise di- rectly from sea level to elevations exceeding 1000 m, and there are spectacu- lar fjords along the coast. Furthermore, most of these highlands support active cirque, valley, and piedmont glaciers that add to their scenic value (see Wahrhaftig, 1965). The geology of the Pacific Mountain System in Alaska is similar in many ways to that of the lower Pacific Coast States. The Aleutian Range con- sists of folded and faulted Mesozoic and Cenozoic sedimentary units with lo- cal granitic intrusions. Tertiary to Recent volcanic deposits and structures are common. Most of the southern portion of the Alaska Range is underlain by large granitic batholiths intruded into deformed Paleozoic and Mesozoic vol- canic and sedimentary units. The central and eastern parts consist of a syn- clinorium with Paleozoic and possibly Precambrian units on the flanks and Cretaceous rocks in the center. This structure is broken by numerous longi- tudinal faults parallel to the mountain axis; many intrusive structures are also present. The Boundary Ranges are underlain by the Coast Range batholith with associated metamorphic facies. The Talkeetna, Wrangell, Kenai-Chugach, St. Elias, and other mountains of the Pacific coast system are all extremely complex geologically. Most are complex assemblages of Paleozoic and Mesozoic sedimentary and volcanic units that have been structurally deformed and lith- ologically altered by igneous intrusions. Mesozoic and Cenozoic continental sediments have accumulated where downfaulted and downwarped basins were formed within and flanking the mountain structure proper (Cobb, 1974c). Alaska's position on the Pacific Rim places it within a tectonically very active zone of the Earth's crust. This is shown by the fact that almost one-half of the more than 75 volcanoes on the Alaska Peninsula and Aleutian Islands are known to have erupted during the last 200 years (Hunt, 1967) and that the southern half of the state has a history of earthquake activity. The generalized geologic map of the state (Figure 5) shows the presence of a series of subparallel, arcuate fault zones and swarms that essentially tra- verse the entire state from east to west and follow the general orientation of the southern coastline. Three principal fault complexes have been identi- fied: (1) a northern zone extending from the vicinity of Nunivak Island in the Bering Sea to approximately 80 to 120 km north of Fairbanks, then curving to the southeast along the southern border of the Yukon Flats area (Figure 4); (2) a central zone including the Holitna, Farewell, McKinley, Denali, and Totschunda faults and extending from approximately 160 km west of Dillingham northeastward to follow the northern flank of the Alaska Range; and (3) a southern zone that includes the Bruin Bay, Eagle River, Border Ranges, and Castle Mountain faults and rims the state from the Alaska Peninsula through the Anchorage area to the southern tip of the panhandle (Beikman, 1980). A fourth major zone can also be identified within the structural band of the Brooks Range. Thus, most areas of the state are subject to earthquake activity due to the regional structural setting; however, based upon historical records, the southern half of the state is particularly susceptible to severe quakes and associated tsunami, or sea waves. Hunt (1967) reported that in Anchor- age, a household may feel six earthquakes per year. Based upon the work of 14 Thenhaus and others, the National Research Council (1980) reported that earthquakes of magnitude 6.4 to 7.0 on the Richter Scale can be expected to occur in the Cook Inlet region once every three years. A magnitude of 7.0 is the accepted lower limit for a major destructive earthquake, but under cer- tain geologic conditions and poor design practices, considerable damage can occur at lesser magnitudes. The Good Friday earthquake of 1964 was centered in an area about 120 km east of Anchorage in Prince William Sound near Valdez. This quake, which measured 8.4-8.6 on the Richter Scale, left 114 dead or missing, was felt over more than a million square kilometers, and caused property damage totaling hundreds of millions of dollars (Grantz et al., 1964). Much of the damage was in coastal areas as a result of the tsunami generated by the crustal movement and submarine landslides. 2.3. CLIMATE The climatic conditions of Alaska are strongly influenced by the com- bined effects of its latitude (mostly between 60° and 70° north), its great size and geographic extent, the diversity of topography and relief that in- cludes large mountain masses traversing the width of the state, and the fact that it is bordered on three sides by ocean waters. As expected, Alaskan climate varies widely with location and season; however, four major climatic zones (Figure 6) can be delineated on the basis of temperature and precipita- tion. 2.3.1 Arctic Zone This zone extends from the Brooks Range north to the Arctic Ocean and is characterized generally by extended periods of low temperatures and strong winds, particularly near the coast. Average annual precipitation ranges from about 50 cm (20 in.) at higher elevations in the Brooks Range to 10 cm (4 in.) or less along the coast (Figure 7). The area receives comparatively little snowfall (Figure 8), and average accumulation is only about 76 cm; once snow has fallen, however, it remains until summer. Temperatures in the Arctic Zone remain relatively low throughout the year and seasonal variations are less than those experienced in the interior part of the state (Figures 9 and 10). Mean annual temperature is about -8°C (17°F) and the average monthly temperature remains below freezing for about eight months of the year. Although temperatures may approach -50° to -57°C (-60° to -70°F) (Hartman and Johnson, 1978), winter temperatures commonly range between -21° and -32°C (National Research Council, 1980). Adding to the intensity of winter conditions is the fact that for nearly two months of this season the sun does not rise and for more than three months there are less than eight hours of twilight conditions (Figure 11). Summer temperatures are usually within the -1° to 10°C range (see Figure 10), with rare extremes reaching 21° to 27°C. During the summer months daily periods of light average 20 hours; length of the growing season is only 9 to 11 weeks (National Research Coun- cil, 1980). 15 A- ARCTIC C~— CONTINENTAL T ~ TRANSITION M_—MaARITIMe Fig.6 Climatic Zones of Alaska (from Selkregg, 1974-76) 2.3.2 Continental Zone The Continental climatic zone includes the vast interior portion of the state and is characterized by extreme daily and seasonal temperature variations, low precipitation and humidity, and clear skies. Annual precipi- tation generally ranges from less than 20 to about 76 cm, with most of the area receiving about 25 to 50 cm (Figure 7). Snowfall tends to vary with elevation and ranges from approximately 130 cm in the lowlands to 250 cm or more at higher elevations. Monthly average temperatures are below freezing approximately seven months of the year and the normal range of winter temper- atures is -29° to -4°C. Temperatures occasionally fall to -57°C or below. Summer temperatures are commonly in the range of 3° to 21°C, but temperatures as high as 38°C have been recorded. There is an 82°C range between the re- corded maximum and minimum temperatures in this climatic zone (Selkregg, 1974-76). Daylight hours vary from less than 8 in the winter to more than 18 in the summer. The growing season lasts 10 to 15 weeks (National Research Council, 1980). % ; 7 . aN A Freee HEE SERA : Ss j \ He aa Nie } OZ a ae \B&n ia ® 0 Adapted from National Weather Service and U. S. Geological Survey ad, Fig. 7 Mean Annual Precipitation Distribution, in Inches (from Selkregg, 1974-76) 9T al Weather Service and U. S. Geological Survey 18 MAXIMUM TEMPERATURE JANUARY MINIMUM TEMPERATURE JANUARY Fig. 9 Mean Daily Maximum and Minimum Temperature Distribu- tions for January, in Degrees Fahrenheit (from Selkregg, 1974-76) 19 MAXIMUM TEMPERATURE JULY MINIMUM TEMPERATURE JULY Fig. 10 Mean Daily Maximum and Minimum Temperature Distribu- tions for July, in Degrees Fahrenheit (from Selkregg, 1974-76) a —— + NUOUS SUNLIGHT Fig. it Hours of Sunlight and Twilight for Different Latitudes and Months of the Year (from Selkregg, 1974-76) 0c 21 2.3.3 Transition Zone Lying south and west of the Continental zone (Figure 6), the Transi- tion zone has a less extreme climate than that of the interior. Climatic conditions in this zone resemble those of both adjacent zones. Mean annual temperature is about -2°C in the northwestern part of this zone (Figure 6) and about 0°C in the south. Precipitation usually ranges from less than 30 cm in the north to 150 cm in the south (Federal-State Land Use Planning Commission for Alaska, 1974). 2.3.4 Maritime Zone This zone borders the Gulf of Alaska and is characterized by small temperature variations, high humidity, heavy precipitation, and high fre- quencies of clouds and fog (Hartman and Johnson, 1978). Average annual tem- peratures for the southeastern, south-central, and southwestern portions of this zone are 6°C, 4°C, and 3°C respectively, and mean annual precipitation equals 230 cm, 100 cm, and 56 cm for the same areas (Federal-State Land Use Planning Commission for Alaska, 1974). Some portions of southeastern Alaska receive more than 500 cm of precipitation annually, including snowfall depths that exceed 1000 cm. In the central portion of this zone and in the south-central part of the state, winter temperatures generally range from -18° to 4°C, with average annual snowfall being 180 to 250 cm. Several hundred centimeters of snow may accumulate in mountainous areas. Average monthly temperature is below freez- ing about five months of the year. Summer temperatures range from 7° to 18°C but may rise to 32°C on rare occasions; length of the growing season is 10 to 16 weeks (National Research Council, 1980). 2.4 PERMAFROST Permafrost is probably the most unusual environmental attribute of Alaska, as well as one of the more important considerations in regard to land use. Permafrost is defined as soil, unconsolidated deposits, and bedrock in which temperatures have been less than 0°C for two years or longer; it is de- fined solely on the basis of temperature and does not consider texture, de- gree of induration, or water or ice content (Ferrians et al., 1969; Péwée, 1975; Williams, 1970). Although the term is usually interpreted as "perma- nently frozen ground," liquid water can be present at temperatures below 0°C due to depression of the freezing point by the presence of dissolved mate- rials or elevated capillary forces between clay particles; furthermore, dry permafrost, which contains no pore waters or ice, has also been found. 2.4.1 Occurrence and Formation Ferrians et al. (1969) estimated that approximately 20% of the world's land area is underlain by permafrost and that 50% of Canada and 85% of Alaska are within permafrost zones. As shown in Figure 12, permafrost conditions in Alaska range from continuous permafrost, generally north of the Arctic Circle, to a zone along the southern boundary of the state where permafrost Generally underlain by continuous permafrost Underlain by discontinuous permafrost oe. Underlain by isolated ma ~ masses of permafrost StL “sland Generally free from permafrost kilometers oO 100 200 300 a 0 100 200 — miles Fig. 12 Distribution of Permafrost (Used with permission from C. W. Hartman and P. R. Johnson, Copyright 1978, Environmental Atlas of Alaska, Institute of Water Resources, University of Alaska, Fairbanks 99701) eZ aa is generally lacking. The vast interior of the state is underlain by discon- tinuous and isolated (sporadic) masses of permafrost. Thickness of perma-— frost in Alaska generally decreases in a southerly direction, ranging from more than 396 m near Point Barrow to less than 0.3 m in the south. Within the interior, zones that range from 30 to 60 m in thickness are common, with local occurrences of 120 m or more (National Research Council, 1980). The permafrost table is the upper boundary of permanently frozen ground, and the zone between this and the ground surface is called the supra- permafrost layer (Figure 13). The active layer is that portion of the supra- permafrost zone containing seasonally frozen ground that thaws during summer. The active zone may include all or a part of the suprapermafrost zone, and zones of unfrozen ground called taliks may be found above, within, or below the permafrost zone (Ferrians et al., 1969). Water trapped in taliks may exist under considerable pressure; under some conditions this water may rup- ture the overlying material, flow out onto the surface as springs, and freeze in a mound or sheet called aufeis. Ice wedges (vertical orientation) and lenses (horizontal orientation) may make up a large portion of the subsurface volume. Pingos, patterned ground (stone rings, mounds, polygons, stripes), solifluction lobes, thermokarst features (depressions and lakes resulting from the melting of subsurface ice), tundra and muskeg areas, muck deposits (frozen, organic-rich silts), beaded or button drainage, and other surface manifestations of freeze-thaw processes can be observed in various permafrost areas within Alaska. Surface of ground tween base of active layer and permafrost table) Suprapermafrost layer (ground above permafrost table in \ cluding taliks, if present) Permafrost table (upper sur- Talik (unfroz ‘ound withi: face of permafrost) ( en ground within permafrost) Permafrost (perennially frozen ground) Talik (unfrozen ground below permafrost) Fig. 13 Occurrence of Taliks in Relation to the Active Layer, Suprapermafrost Zone, Permafrost Table, and Perma- frost (from Ferrians et al., 1969) 24 The formation and maintenance of permafrost depends upon the thermal regime and thus the ground-surface temperature, as well as the rate of heat exchange between the soil body and the atmosphere. Consequently, the thick- ness and three-dimensional properties of permafrost are affected by the size, shape, and distribution of landscape features (e.g., lakes, rivers, vegeta- tion, and topographic form) that act as heat sources, sinks, or insulation. Shallow lakes act like the active layer of permafrost; however, spring warm- ing is more efficient in the circulating water and results in higher mean an- nual temperatures on these shallow lake bottoms than in adjacent surface areas. Thus, permafrost is thinned beneath these water bodies (Figure 14). Deep lakes that do not freeze solidly will maintain elevated temperatures on the bottom and are underlain by a thaw basin within the permafrost. If the lake is drained, water in this thinned basin refreezes, expands upward, and a pingo is formed. Under certain conditions (see Ferrians et al., 1969), an unfrozen "chimney" through the permafrost layer can form beneath large, deep lakes and provide a year-round source of ground water, which is otherwise scarce at shallow depths in continuous permafrost areas. In the absence of significant bodies of water, the top and bottom surfaces of the permafrost layer tend to parallel the ground surface (Figure 14). Although the upper portions of a permafrost zone may be quick to respond to surface changes, hundreds or thousands of years are required for surface changes to affect the permafrost bottom where thick units exist (Ferrians et al., 1969). Large deep lake Fig. 14 Effect of Surface Features on the Distribution of Permafrost Zone (from Ferrians et al., 1969, after Lachenbruch, 1968) 25 2.4.2 Effects of Permafrost Permafrost strongly influences soil properties, vegetation, and hydro- logic characteristics. The frozen state limits rooting depths of plants, prevents infiltration and vertical water percolation (thus inhibiting ground- water recharge and causing surface runoff), and is of paramount importance in the engineering design of any construction project. Under most ambient con- ditions, ice acts as a cementing agent to unconsolidated sediments or soil materials, and the permafrost unit is strong and stable. In fact, under cer- tain circumstances, frozen earth materials can be used as construction ele- ments themselves (Swinzow, 1969). However, when these materials are physi- cally or thermally disturbed, they may become extremely unstable and lose structural integrity. In unfrozen sediments, individual sediment grains are in contact with each other and, under most circumstances, water can only fill the interstices or intergranular pore spaces. Under permafrost conditions, water has frozen, expanded, and occupied a greater percentage of the total volume. Over long periods of time, large masses of ice (such as wedges and lenses) have formed in some areas and may occupy more than 90% of the total volume (National Re- search Council, 1980). Upon thawing, the grains of sediment or soil tend to settle and the water is squeezed out. This settling action tends to be vari- able laterally due to the heterogeneous nature of the permafrost materials; consequently, the resulting differential surface subsidence can destroy buildings, roads, and other facilities on the surface. While settling is in progress and water under pressure is present in the pores and voids, individual grains are not in frictional contact and the water acts as a lubricating agent. As a result, the soil-water-ice unit has limited bearing strength and is highly susceptible to down-slope movement (mass wasting) under the influence of gravity. The permeability of the geo- logic material is an important determinant in the severity of this latter problem. Coarse-grained materials of high permeability remain relatively stable during the thawing process because settling is minimal and water can escape easily. If fine-grained materials (silts and clays) are abundant, porosity is high but permeability is low. Water formed during melting re- mains trapped and causes instability. Furthermore, as these materials con- tinue to thaw, the entire unit evolves into a dense fluid that is almost impossible to contain or handle unless the water is removed or refrozen. When thawing permafrost is refrozen, expansion due to ice formation causes frost heaving and differential surface uplift. Thus, engineering problems in permafrost areas are related to two pri- mary characteristics of the frozen material: the volume of ice present and the physical characteristics (principally grain size) of the earth material. More extreme problems can be anticipated where ice volumes in the subsurface are greatest and the materials are predominantly fine-grained with low perme- abilities. Engineering practice in permafrost regions generally consists of three basic approaches: avoid it (in discontinuous or sporadic zones), de- stroy it (strip the surface to increase depth of thaw or excavate and replace with coarse materials), or preserve it (use piles, gravel insulating pads, refrigeration, etc.) (Hartman and Johnson, 1978). Special study, planning, and design are required for the construction of airfields, roads and high- 26 ways, railroads, bridges, dams, buildings, and pipelines; subsurface explora- tion techniques; water supply development; distribution of utility services; and sewage and solid waste disposal (Brown and Grave, 1979; Ferrians et al., 1969; Stearns, 1966; and Swinzow, 1969). However, these measures increase design and construction costs and, in many cases, maintenance costs after construction. Consequently, the presence and location of permafrost condi- tions in Alaska are important environmental factors to be considered during the development of energy and mineral resources in the state. 2.5 WATER RESOURCES AND HYDROLOGY The hydrologic characteristics and water resources of Alaska are de- termined principally by climatic, geologic, and physiographic conditions and are strongly influenced by the distribution of permafrost. Detailed hydro- logic information and data, however, are not available for much of the state. As an indication of the general paucity of data, the U.S. Geological Survey collected continuous discharge data during the 1977 water year at only 112 gaging stations in Alaska (U.S. Geological Survey, 1978a). By comparison, a total of 163 continuously recording stations were operated during the same water year in Illinois, which is less than one-tenth the size of Alaska (U.S. Geological Survey, 1978b). However, despite this shortage of detailed data and observations, the regional quantity and quality characteristics of Alaska's surface- and ground-water systems are fairly well understood. 2.5.1 Surface Water The average annual runoff in Alaska is estimated to be 25,500 to 31,150 m3/sec (900,000 to 1,100,000 ft3/sec), with an additional 4530 to 5660 m3/sec (160,000 to 200,000 ft3/sec) of flow originating in Canada (Balding, 1976; Feulner et al., 1971). This runoff represents approximately 40% of all of the freshwater runoff for the entire United States (Slaughter et al., 1974). All major Alaskan streams originate within the state except for the Yukon and Porcupine rivers of central Alaska and the Alsek, Stikine, and Taku rivers in the southeast (see Figure 2), which have their headwaters in Canada. The Yukon has the largest drainage basin in the state and collects runoff from more than 518,000 km2 of the Alaskan interior; its average annual flow rate exceeds 5660 m3/sec (Iseri and Langbein, 1974). Drainage areas and runoff quantities for the major rivers of Alaska are given in Table 1; the locations of these rivers are shown in Figure 2. Additional information on areal distribution and variation of runoff, and thus surface water availability, is obtained when runoff per unit of drainage area is examined. Figure 15 shows the contoured values of estimated mean annual runoff in cubic feet per second per square mile (cfsm or £t3/sec/ mi?) for the state. This map shows that mean annual runoff ranges from less than 1 cfsm in the north slope and interior regions to about 12 cfsm along the Gulf of Alaska and 30 cfsm in southeastern Alaska. A comparison of Fig- ures 7 and 15 clearly indicates that most of the variation in runoff can be attributed directly to variation in precipitation. 27 Table 1 Major Alaskan Rivers, Runoff Rates, and Drainage Areas Average Annual Drainage Area Runoff Runoff Natural Runoff River (mi*)4 (£t3/sec)> (£t3/sec/mi2)¢ (109 gal/day)4 Alsek 9, 500° 20,000 Ziel 12.9 Colville 24,000 12,000 0.5 7.8 Copper 24,400f 51,000 Del 34.0 Kobuk 12,000 13,000 Led 8.4 Kuskokwim 43,600 62,000 1.4 40.1 Kvichak 7,700 22,000 2.9 14.2 Noatak 12, 600 10,000 0.8 6.5 Nushagak 14,100 20,000 1.4 12.9 Stikine 19, 700¢ 62,000 Sel 40.1 Susitna 20,000 46,500 253) 2160 Taku 6, 700€ 11,000 136 Tek Yukon 330, 0008 216,000 0.7 139 Source: Feulner et al. (1971). a] mi2 = 2.59 km2, bi £t3/sec = 0.0283 m3/sec. C1 £t3/sec/mi2 = 0.0109 m3/sec/km2. 4109 gal/day = 0.438 m/sec. Most of drainage area is in Canada. fIncludes 1270 mi2 in Canada. SIncludes 110,000 mi2 in Canada. Seasonal variations in runoff and streamflow tend to be greatest in the central interior and northern parts of Alaska. Due to the low winter temperatures (see Figure 9), precipitation is in the form of snow and no run- off occurs; many smaller streams freeze solidly and flow ceases altogether. On the North Slope, 90 to 95% of the runoff occurs from June to mid-Sep- tember; within the interior 80 to 85% of runoff occurs from May through Sep- tember. South of the Alaska Range, seasonal variations are less, but almost 75% of the runoff occurs during the five months of May through September (National Research Council, 1980). Mean annual low monthly runoff (Feulner et al., 1971, Figure 11) is zero north of the southern margin of the Brooks Range, 0.1 to 0.3 cfsm throughout the interior, and about 0.5 to 10 cfsm south of the Alaska Range, with the higher values originating in the pan- handle. Variations in flow other than those due to freezing are largely a function of topography and seasonal fluctuations in temperature and precipi- tation. At lower elevations in the southern part of the state, rains in September and October can produce high flow conditions comparable to those resulting from snowmelt in May and June. At higher elevations throughout the state, minimum flows occur before spring breakup of ice, and high flows fol- low as a result of snow and ice melting. Similar conditions exist on the North Slope (National Research Council, 1980). wae St. Lawrence Istand Fig. Kilometers © 100 200 300 ° 100 200 Miles 15 Estimated Mean Annual Runoff,in Cubic Feet per Second per Square Mile (from Balding, 1976) 872 29 Peak flows and floods caused by rain usually occur in August and September. Some runoff peaks of higher magnitude result from the combined effect of precipitation and melting snow and ice. The pattern of variation of estimated mean annual peak runoff (Figure 16) reflects the combined ef- fects of precipitation distribution (see Figure 7) and topographic influence on snowfall accumulation (see Figure 8). Figure 16 indicates that peak run- off per unit of drainage area is greatest in lower areas along the southern margin of the state, is intermediate at higher elevations and other areas that receive increased snowfall throughout the state, and is lowest within the interior and the Cook Inlet area. Another type of flooding common to Alaska is caused by ice formation and subsequent flow blockage in stream channels. This ice often extends be- yond the channel limits and causes subsequent runoff to flow through the flood plain or adjacent valley-bottom areas (National Research Council, 1980). Additionally, ice in spring meltwater can cause jamming and thus in- creased flood depths, as well as significant structural damage to bridges. About 4% of Alaska is covered by glaciers, principally in the coastal mountains adjacent to the Gulf of Alaska and within the Alaska Range. Gla- ciers are important in the hydrologic regime because of the large quantities of water they store as ice and because of their effect on the quantity and quality of meltwater that enters streams. Streams receiving significant quantities of meltwater usually carry larger flows than do other streams; this is particularly important during late summer when flows have diminished to low levels following spring runoff. Glacier-fed streams usually have significantly elevated sediment concentrations, particularly silts and clays, due to the finely-ground rock material (glacial flour) contained in the melt- water. It is estimated that Alaska has more than 3,000,000 lakes that cover about 1% (more than 15,000 km2) of the state (National Research Council, 1980). Iliamna Lake, located about 160 km east of Dillingham, is the largest with a surface area of approximately 2,600 km?. The state's seven largest lakes each have an area in excess of 260 km?, and a total of 94 lakes are known to have areas greater than 26 km2 (Feulner et al., 1971). Most of the larger lakes are in south-central and southwestern Alaska. At least 20 lakes are known to be at least 76 m deep. The larger lakes were probably formed by glacial scour and many smaller ones had their origin in associated glacial processes. Other Alaskan lakes, particularly the many located (at least sea- sonally) on the North Slope are thermokarst features formed by ground sub- sidence as a result of permafrost thawing. Thaw lakes are also abundant in the more level parts of the Yukon-Kuskokwim Delta and Yukon Flats areas (National Research Council, 1980). 2.5.2 Ground Water Ground-water conditions in Alaska are highly diverse and are dependent upon local and regional geologic factors and permafrost conditions. The thickness and extent of permafrost generally decreases from north to south (see Figure 13) and the quantity of available ground water shows a corres- ponding increase to the south where geological conditions are favorable. ma St. Lawrence Island Fig. Kodiak laland Kilometers © 100 200 300 pe ey ° 100 200 Miles 16 Estimated Mean Annual Peak Runoff, in Cubic Feet per Second per Square Mile (from Balding, 1976) o€ 31 Four generalized geohydrologic environments are recognized in Alaska (Will- iams, 1970; Zenone and Anderson, 1978): (1) alluvial deposits of river valleys, including terraces, flood plains, and alluvial fans of major valleys and smaller upland and mountain valleys; (2) glacial and glaciolacustrine (mixed glacial and lake-bottom sediments) deposits of interior valleys; (3) coastal-lowland deposits; and (4) bedrock uplands. The occurrence and avail- ability of ground water in each of these environments are modified by perma- frost conditions. The greatest volume of ground water in storage is con- tained in alluvium. Such deposits have the greatest recharge potential be- cause they are usually connected hydraulically to the surface water system; furthermore, these deposits tend to have the greatest water yield potential due to their greater permeability (Figure 17). Within the zone of continuous permafrost, unfrozen alluvium -- and thus ground water -- is found only under major streams and beneath lakes greater than about 2 m deep (Balding, 1976; Williams, 1970). In the zone of discontinuous permafrost (see Figure 13), the presence of permafrost does not significantly affect alluvial aquifers associated with major rivers (e.g., Yukon, Tanana, and Kuskokwim) because the total alluvial thickness is much greater than that of the permafrost. In fact, the greatest ground-water resources of the state are found under these circumstances (Figure 17). However, in less extensive alluvial deposits associated with upland and smaller tributary streams, ground-water movement is restricted by the presence of permafrost. Glacial and glaciolacustrine deposits may contain interbedded gravels, sands, silts, beds or lenses of clay, and glaciofluvial sediments associated with glacial tills of variable characteristics. Most such deposits are pre- dominantly fine-grained and are rendered even more impermeable by the pres- ence of permafrost. Although glacial deposits are widespread in Alaska, they have not been adequately explored for ground water and probably do not con- tain significant aquifers except in the southern part of the state (Zenone and Anderson, 1978). Coastal lowland deposits include coastal plain sediments; deltas; bars, spits, and beach sediments; and unconsolidated deposits contained with- in coastal basins and valleys. Geologically, the majority of these deposits are favorable for ground-water availability, but actual or potential yields are restricted by permafrost conditions (see Williams, 1970). Unconsolidated deposits are either thin, of poor permeability, or ab- sent over approximately 75% of Alaska. Under these conditions, bedrock aqui- fers represent the only source of ground water. Although the flow rates of a number of springs at scattered locations are known, detailed ground-water in- formation on potential aquifer units is generally not available for most of the area (Williams, 1970; Zenone and Anderson, 1978). The generalized availability of ground water is shown in Figure 17. As discussed, unconsolidated alluvium and related fluvial and glaciofluvial sediments associated with the major streams show the most promise for signif- icant ground-water yields. More detailed discussion of geohydrologic condi- tions throughout the state can be found in Balding (1976), Hopkins et al. (1955), Williams (1970), Zenone and Anderson (1978), and other reports pub- lished by the U.S. Geological Survey. St. Lawrence Island Fig. 17 Generalized Key Availability of Water in Gallons per Minute (C_) Less than 10 10-100 (22 100-1000 MB More than 1000 Kilometers 0 100 200 300 ° 100 200 Miles Availability of Ground Water (after Feulner et al., 1971) ce 33 2.5.3 Water Quality Because Alaska has only minimal industrial development, water quality variations are due almost entirely to natural influences. Current standards for public water supplies suggest that dissolved solids not exceed 500 mg/L; water containing 250 mg/L or less dissolved solids concentrations is prefer- able for most municipal and industrial uses (Feulner et al., 1971). Most streams in Alaska above tidal influences contain water of the calcium bicar- bonate type with less than 200 mg/L of total dissolved solids (Balding, 1976) and a normal pH range of 6.0 to 8.2 (National Research Council, 1980). Streams that drain lowlands and intermontane basins and those in areas of lower precipitation tend to have greater concentrations of dissolved solids than do streams in mountains and upland areas and those in areas of higher precipitation. Some streams, particularly those draining bog and muskeg areas, contain water that is brown due to elevated levels of iron and organic matter. Generally, concentrations of total dissolved solids in surface waters fall below federal and state standards; some streams, however, exceed existing limits of turbidity and iron content (National Research Council, 1980). Chemical characteristics of ground water vary widely and are related to geologic and permafrost conditions. Concentrations of dissolved constit- uents range from 19 to 64,200 mg/L, but most samples contain less than 250 mg/L (Balding, 1976). Calcium bicarbonate or calcium magnesium bicarbonate | water is most common within the interior, whereas sodium bicarbonate or sodium chloride waters are common in coastal areas. Water obtained beneath permafrost often contains elevated concentrations of dissolved materials, commonly magnesium sulfate or sodium chloride; shallow wells often contain excessive amounts of iron and magnesium (National Research Council, 1980). The amount of suspended sediment in Alaskan streams varies greatly from stream to stream and with time. However, streamflow containing glacial meltwater usually has significantly greater sediment concentrations than com- parable flows containing only surface runoff. Nonglacial streams generally contain less than 200 mg/L of suspended sediment during the summer, whereas concentrations in glacial streams may reach 2,000 mg/L. Concentrations in nonglacial streams are usually highest during spring melt or periods of heavy rainfall; glacial streams carry the most sediment in middle or late summer. Both types of streams normally carry less sediment during fall and winter (Balding, 1976). 2.5.4 Water Use Only a fraction of Alaska's vast water resources is currently used. Water is used principally for homes, agriculture, mining and oil and gas pro- duction, petrochemical plants, hydroelectric and steam-electric plants, and the food and fiber industry, especially in seafood processing establishments and pulp mills (Balding, 1976). The estimated water use for Alaska, by cate- gory and region, is given in Table 2. These values reflect estimated usage during 1975 (Balding, 1976), and current use is undoubtedly greater. With the exception of the Arctic (North Slope) Region, abundant water resources exist for future development. In the Arctic Region, however, any significant 34 Table 2 Estimated Water Use by Subregion (108 gal/day)@ Subregion North- South- South- South- Use Water Use Arctic west Yukon west Central east Totals Domestic 0.551 0.458 9.403 0.864 38.604 17.400 67.307 Livestock a 0.045 0.011 0.112 0.078 = 0.246 Seafood Processing -- ard ac 0.515 3.299 1.260 5.074 Oil and Gas Devel opment 0.400 -- -- -- --b -- 0.400 Petrochemical == = =— == 2.578 -- 2.578 Pulp Mills = =< == = = 80.000 80.000 Coal Processing == = 0.270 = == => 0.270 Steam Electric ad == 26.625 = 7.000 = 33.625 Sand and Gravel == a 9.033 0.231 1.570 0.036 10.870 Fish Hatcheries - == == 0.480 67.730 17.380 85.590 Irrigation (in acre-ft/yr)° -- -- 0.158 -- 0.455 a 0.613 Source: Balding (1976). 4106 gallons/day = 3,785 m3/day. b_- = Information not available. ©) acre-ft = 1,233.5 m. economic or natural resource development, with its attendant population in- creases and services expansions, will require careful planning in order to ensure an adequate water supply. 2.6 ECOREGIONS Environmental conditions, as well as the types and varieties of plants and animals, in Alaska are dramatically different from those in the contermi- nous 48 states. Alaskan wildlife in particular has received much attention and is of particular significance from not only a conservationist point of view but an economic and cultural one as well. Much of the populace relies upon wildlife for commercial, recreational, and subsistence uses. Further- more, many aspects of Alaska's plant and animal communities and environmental characteristics are unique in that the arctic, subarctic, and marine ecosys- tems of Alaska have no exact counterparts elsewhere in the United States (National Research Council, 1980). Ecology is concerned with all of the interrelationships between organ- isms and their environment. This includes a consideration of the types and numbers of plants and animals present, their relationships to each other, landforms and geologic conditions, climate, soils, and the various biological and physicochemical processes through which each of these factors influence others. As mentioned previously, many of the physical characteristics vary widely across Alaska; there is also considerable variation in ecological re- lationships and in the types of plants and animals adapted to these condi- tions. 35 To simplify the examination and consideration of ecological variations for land management needs, Bailey (1976, 1980) presented a hierarchical clas- sification scheme with which to regionalize both the biotic and abiotic char- acteristics of different geographic areas in the United States. Although Bailey presented (1976) classification units that ranged from domain (subcon- tinental area of broad climatic similarity) to individual site (single soil and habitat type or phase), he used the term ecoregion to designate biogeo- graphical units of any size or rank. He defined (1976) this latter term as "... a geographical area over which the environmental complex, produced by climate, topography, and soil, is sufficiently uniform to permit development of characteristic types of ecological associations." The ecoregions of Alaska are shown in Figure 18; the following descriptions of these units are derived from Bailey (1980). 2.6.1 Arctic Tundra This ecoregion occupies 178,450 km? and includes all of the Arctic Coastal Plain, or North Slope, physiographic region. Extensive marshes and lakes exist during the summer because permafrost inhibits internal drainage. Wet, cold soils with weakly differentiated horizons are common. Those soils underlain by coarse glacial outwash and till are well drained and loamy, and all have permafrost and ice features. Lowland soils tend to be deep, wet, and silty, whereas upland areas contain localized areas of poorly drained soils high in clay content. Cottongrass-tussock is the most widespread vege- tation system in the Arctic and is found in association with sedges, lichens, mosses, dwarf birch, Labrador-tea, and other dwarf shrubs. Ptarmigans, hawks, ravens, and owls are common birds, and the shore and lake areas are habitat for millions of migratory birds during the summer. Brown bear, wol- verine, wolf, caribou, arctic hare, weasel, mink, and lemming inhabit inland areas, and polar bear, arctic fox, and walrus are found on ice packs and coastal areas during winter. 2.6.2 Bering Tundra A western extension of the Arctic coastal plain and with similar to- pography, this 224, 600-km2 region includes the Seward Peninsula and the Bering Sea and Bristol Bay coastal plains. Numerous shallow lakes and marshes are found along the coast. The climate is less severe than that of the Arctic coastal plain, but variations between summer and winter are more extreme. Generally the poorly developed coastal soils formed from silt, sand, and marine sediments, with some alluvial soils having formed in the valley bottoms of the Yukon and Kuskokwim rivers. Vegetation along the wet coastal areas is mainly sedge and cottongrass with woody species in higher areas. Birch-willow-alder thickets exist among the alpine tundra vegetation on the Seward Peninsula. The bottom lands adjacent to the lower Yukon and Kuskokwim contain white spruce, cottonwood, and balsam poplar with a dense undergrowth of willow, alder, and blueberry bushes. These bottom lands pro- vide habitat for game birds, furbearers, and moose. Brown and black bear, wolf, coyote, wolverine, reindeer, caribou, fox, lynx, beaver, mink, marten, and weasel are found in upland and coastal areas; polar bear, walrus, and arctic fox can also be found in northern coastal areas. The coastal areas also provide extensive habitat for migrating waterfowl and shorebirds. a St. Lawrence Island Arctic Tundra Brooks Range Yukon Forest Kilometers © 100 200 300 Sao ° 100 Miles Fig. 18 Ecoregions of Alaska (after Bailey, 1976) 9€ 37 2.6.3 Brooks Range A 965-km-long band extending from Canada to the Chukchi Sea and occu- pying 138,000 km2 makes up this ecoregion. Soils in this mountainous area are generally thin, rocky, poorly developed, and have rapid internal drain- age. At higher elevations, plant cover is sparse and consists mainly of mats of herbaceous and shrubby plants such as dwarf arctic birch, crowberry, arctic willow, resin birch, and dwarf blueberry. At lower elevations, pro- ductive mats of cottongrass, bluejoint reedgrass, mosses, dwarf willow, dwarf birch, Labrador-tea, and bistort provide valuable forage for caribou. The Brooks Range is an important big-game area, with brown and black bear, wolf, wolverine, caribou, and Dall sheep found throughout. Other mammals include marmot, red and arctic fox, ground squirrel, lemming, and pika. The region is an important summer nesting area for migrating birds and is habitat for several raptors including golden eagles, marsh hawks, gyrfalcons, and owls. 2.6.4 Yukon Forest Approximately 480,450 km? of interior Alaska are included in this eco- region. This area of broad valleys, dissected uplands, and lowland basins is covered by extensive alluvial deposits and occupies the area between the Brooks and Alaska Ranges. A semiarid, continental climate with short hot summers and long severe winters is characteristic of the area. Upland soils are well-drained, shallow, and support spruce-hardwood forests. Deep, well- drained soils overlying poorly weathered, coarse-grained deposits are common on lower slopes and in valley bottoms. Major river bottoms are covered with dense white spruce-cotton-wood-poplar forest and undergrowth of alder, wil- low, and assorted shrubs and berry bushes. Valley edges support coniferous and evergreen forests with abundant black spruce; the undergrowth consists of willow, dwarf birch, fern, blueberry, lichens, and mosses. Dense forest of white spruce-birch and aspen-poplar cover upland areas as high as 900 m above sea level, with alpine tundra vegetation above. The extensive forests pro- vide good habitat for furbearing animals, woodland game birds, and mammals. Black and brown bear, wolf, wolverine, and moose are common, as are such smaller mammals as red fox, beaver, mink, muskrat, weasel, marten, squirrels, and mice. Sharptail, spruce and ruffed grouse, ptarmigan, hawks, owls, and raven are among the upland bird population; several raptor species, including osprey, gyrfalcon, hawks, and American peregrine falcon, inhabit cliffs along the Yukon and Porcupine rivers. 2.6.5 Alaska Range This ecoregion includes the Alaska Range, Alaska Peninsula, and Aleu- tian Islands, and occupies 264,700 km2. Topography is rugged with high re- lief and the coastlines are dissected, steep sloped, and rocky. Bottom land in the major river valleys generally has stratified, well-drained soils; shallow, well-drained, poorly developed soils are most common in the upland forest areas. Wet, organic-rich soils are found in moist tundra areas and in the Aleutians. Vegetation types are zoned vertically throughout the Alaska and Aleutian ranges. White and black spruce and cottonwood are found in bot- tom lands below 300 m. Between 300 m and timberline (760-1060 m) are spruce- hardwood forests of white spruce, birch, aspen, and poplar with a moss, fern, 38 grass, and berry undergrowth. Tundra ecosystems occur above timberline, with low shrubs, herbaceous plants, moss, lichens, grasses, and sedges dominating. Low heath shrubs, crowberry, bog blueberry, mountain cranberry, alpine- azalea, and dwarf willow are dominant in the Aleutians, and tall grass mead- ows and dense low shrubs predominate in moist, low areas. An abundance of big game is found in the Alaska Range and Peninsula in the form of moose, Dall sheep, bear, wolf, caribou, and wolverine. Beaver, red fox, lynx, otter, squirrel, marten, weasel, and various rodents are also present. Mi- grating waterfowl and shorebirds are abundant along coastal areas during sum- mer. Several bird species, including golden eagles, ptarmigan, and hawks, inhabit the uplands; bald eagles and osprey are found along the coast. No large mammals are found in the Aleutians, but foxes, bald eagles, and hawks prey on the large populations of seabirds inhabiting the islands. Seal, sea lion, and sea otter are abundant in coastal waters. 2.6.6 Pacific Forest About 172,750 km2 of southern and southeastern Alaska are contained in this ecoregion, which is characterized by rugged topography, high local relief, and abundant moisture. Soils are predominantly shallow and poorly developed. Most of the area is covered with dense conifer forests composed of various types of fir, hemlock, cedar, and spruce trees; several species of shrubs provide a dense undergrowth. Bear, moose, deer, mountain lion, and other large mammals live in the region. Small mammals, including mice, squirrels, and marten are common, as are numerous bird species. Fish and other marine animals are abundant offshore. 39 3 POPULATION AND ECONOMY Along with its unique environmental setting, Alaska has socioeconomic conditions that tend to be distinct from those in other states. In 1970, the entire state had slightly fewer inhabitants than the city of Trenton, New Jersey (U.S. Bureau of Census, 1980). The combination of sparse population, vast areas of wilderness, and undeveloped status is considered a valuable asset by many of the state's residents and forms the basis for the prized "Alaskan way of life." Conversely, other residents see an urgent need to use the abundant natural and energy resource base to promote economic growth and improve economic stability. Effecting a solution to this dilemma is an ardu- ous task for state officials. 3.1 POPULATION While Alaska is the largest of the 50 states, it is also the least populated. According to the 1970 census, the total population of the state was only 302,361. A 1978 estimate indicated that this number had increased to 416,400 (Harrison, 1979). Using the 1970 value, the average population density was approximately 0.2 persons per km2 (0.5 per mi2) as compared to an average of 22.2 (57.4) for the 48 conterminous states (U.S. Bureau of Census, 1980). This value increased to only 0.3 persons per km2 in 1978. Actually, these averages are misleading because more than 70% of the 1970 population lived in and adjacent to the three largest cities (Anchorage, Fairbanks, and Juneau), with an additional 14% distributed among the other larger towns (Table 3). Thus, the actual population density for most of the state is sig- nificantly less than the average values indicate. The 1970 census indicated that the population of Eskimos, Indians, and Aleuts was 27,797, 16,276, and 6,581, respectively; this represented approxi- mately 17% of the total populace (Harrison, 1979). Historically, Eskimos in- habited the northern coastal areas, Indians most often lived in the interior and southeastern panhandle, and Aleuts resided in the Aleutian and Kodiak islands. This native population lives primarily in a large number of vil- lages scattered throughout the state, and many individuals continue to pursue the traditional way of life. As many as 32 languages are spoken by Alaskan natives today (Harrison, 1979). 3.2 ECONOMY In the past, the Alaskan economy, and population growth as well, has experienced a number of "boom and bust" periods. Early economic activity centered on fur trade, and commercial fish-canning operations began in 1878 in southeastern Alaska. Gold prospecting increased in the late 1800s; the first major discovery in 1880 eventually led to the establishment of the town of Juneau (Alaska Division of Economic Enterprise, 1980). A few years later gold was discovered in the Klondike River Valley in the Yukon Territory of Canada, and Skagway became the major staging area and supply center for prospectors and miners during the rushes of 1896 and subsequent years. The Klondike discovery led to the search for gold throughout Alaska, and in 1898 a very large discovery was made where the town of Nome now stands. In just 40 Table 3 Alaskan Communities with 1976 Estimated Population Greater than 1000 Commun ity 1970 Census 1976 Estimate? Anchorage 48,029 --b Anchorage Borough 126, 385 179,865 Barrow 2,104 2,471 Bethel 2,416 3,166 Cordova 1,587 ay kad Dillingham 914 1,207 Fairbanks 27,5218 33,956 Homer 1,083 1,612 Juneau® 13,556 18,635 Kenai 3,533 4,837 Ketchikan 6,994 7,719 Kodiak 3,798 4,706 Kotzebue 1,696 2,060 Nome Qe so 2 Dae Palmer 1,140 1,930 Petersburg 2,042 2,334 Seward 1,587 25907 Sitka 6,073 7,216 Soldotna 1,202 1,631 Valdez 1,005 4,205 Wrangell 2,029 2,614 Source: Harrison (1979). @Revised estimate of 7/1/76 by U.S. Bureau of Census. binformation not available. City and borough population combined. two years, Nome grew to a population of 20,000. This discovery encouraged additional prospecting and gold was soon discovered in other parts of Alaska. One strike in the Tanana River Valley in 1902 led to the founding of Fair- banks, which grew to a population of 5,000 by 1905 (Alaska Division of Economic Enterprise, 1980). During the early 1900s, copper, lead, and zinc were mined periodically at scattered locations throughout the territory. The federal government constructed the Alaska Railroad during the period of 1914 to 1923. The line connects Seward to Fairbanks and employed a maximum of 4500 workers at one time; this created another boom to the economy and ensured the survival of Fairbanks. However, the total population of the territory declined significantly from 1910 to 1920 (Harrison, 1979). The start of war in Europe in 1939 signalled another boom period for the economy as Anchorage was chosen the site for military installations and related de- fense activity. In 1942 the Alcan Highway -- the first land transportation 41 link between Alaska and the U.S. -- was constructed to provide an overland supply route to these installations. During World War II, the number of military personnel rose to 150,000; the attendant construction, transporta- tion, and communication activities provided a stimulus that continues today as one of the economic mainstays in Alaska. It is estimated that approxi- mately one-third of all wage and salary jobs in the state in 1979 were sup- ported directly or indirectly by military funds (Alaska Division of Economic Enterprise, 1980). Fluctuating periods of rapid economic growth and relative stability or decline have continued since the 1950s. Much of this growth has been related to the discovery and development of petroleum resources. In 1957, substan- tial oil discoveries were made on the Kenai Peninsula; by 1969, annual pro- duction from the Kenai-Cook Inlet area totalled 7.4 x 10’ barrels. In 1968, ARCO discovered oil at Prudhoe Bay, but complete development of the field was delayed for several years because of land-ownership disputes by natives and environmental concerns regarding pipeline construction. Finally, construc- tion of the trans-Alaska pipeline began in 1974 and Alaska entered yet an- other economic boom. Total cost of the pipeline was in excess of $12 x109, much of which was expended for labor. Upon completion of the project in mid- 1977, thousands of jobs ended and workers left the state. Although the state government has received huge royalty and tax revenues from oil and gas pro- duction, significant general economic problems still remain. Today, the two main economic concerns are for the provision of sufficient jobs and develop- ment of energy and mineral resources (Alaska Division of Economic Enterprise, 1980). Most aspects of Alaska's present economy are directly related to the development of the state's abundant natural resources. The geologic condi- tions of Alaska are highly favorable for mineral occurrence and historically, several mineral commodities have been produced in significant quantities (Table 4). Although hard-rock mineral production has declined in recent years, exploration activities and expenditures have simultaneously increased significantly. Petroleum discoveries and field development in the state have more than compensated for revenues lost from the decline in metallic mineral production. Consequently, the dollar value of total mineral production dur- ing the last 20 years has increased progressively and dramatically following pipeline construction (Table 5). The domestic fishing industry is also a substantial component of the Alaskan economy and has also experienced general growth during recent years Total wholesale value of catch landed increased from $2.54 x 108 in 1974 to more than $1.24 x 109 in 1979. This represents a real value increase, based on the 1974 price index, of 332% (Alaska Division of Economic Enterprise, 1980). This trend is due to the combined results of increased catch weight and escalated market value. Thus, the fishing industry, like the mineral in- dustry, continues to be as significant to the present economy as it has in the past. A third major facet of the economy is the timber industry. Approxi- mately 57 x 10© ha of the state is forested with more than 9 x 10° ha of coastal forests and almost 48 x 10 ha within the interior. About 2.5 x 10° ha of coastal forest, consisting principally of western hemlock and Sitka spruce, together with more than 8.9 x 10° ha in the interior, are considered 42 Table 4 Historical Production of Minerals in Alaska Estimated Total Period of Quantity Mineral Production Units Produced Antimony 1928-1973 Short tons (approx. 53% Sb) 4,390 Coal 1951-1979 Short tons 28,460,000 Copper 1880-1973 Short tons 690,035 Chromite 1917-1957 Long tons (approx. 45% Cr203) 29,000 Crude Petroleum 1958-1979 42-gallon bbl (103) 1,884,857 Gold 1880-1979 Troy oz 30, 149,530 Lead 1906-1973 Short tons 25,028 Mercury 1902-1972 76-1b flasks 29,224 Natural Gas 1948-1979 Cubic ft (106) 1,643,250 Sand and Gravel 1958-1979 Short tons (103) 709,650 Silver 1906-1979 Troy oz 19,082,518 Stone 1921-1979 Short tons 26,000,000 Tin 1902-1979 Short tons 2,464 Tungsten 1916-1958 Short-ton units (W03) 7,000 Source: Alaska Division of Economic Enterprise (1980). commercial timber lands (Stenmark, 1978). Almost all of Alaska's forest pro- ducts are exported. From 1976 through 1979, the annual value of exported timber products averaged more than $192 x 106 and accounted for an average 38% of the total dollar value of exported materials. Almost 83% of these wood products were shipped to Japan (Alaska Division of Economic Enterprise, 1980). Although Alaska has more than 41 x 106 ha of land with some agricul- tural potential and an additional 7 x 106 ha that are potentially suitable for domestic livestock grazing, agricultural production is small. The total annual value of agricultural production for the entire state averaged only $9.4 x 106 from 1977 through 1979. Production of field crops, vegetable crops, and livestock averaged 38.9%, 17.4%, and 43.4%, respectively, of this total (Alaska Division of Economic Enterprise, 1980). Most of the state's soils are classed as "poor" and the remaining areas of greater potential are located in the interior drainages of the Tanana, Yukon, and Kuskokwim rivers and in the Bristol Bay and Cook Inlet areas (Stenmark, 1978). The majority of vegetable and field crop production is found in the Matanuska-Susitna Val- ley, and the principal local market is Anchorage. Although small farms in the Tanana Valley supply some products to Fairbanks, high costs, short grow- ing seasons, and transportation difficulties have kept the total farmland area to less than 31,000 ha. Similarly, livestock production is limited due to costs and climatic conditions with the result that more than 98% of the beef consumed in Alaska is raised outside of the state (Stenmark, 1978). 43 Table 5 Value of Alaskan Mineral Production from 1959-1979, in Thousands of Dollars Crude Natural Sand and Year Petroleum> Gas© Gravel Gold Other Total 1959 295 16 5,265 6,262 8,673 20,511 1960 15230) 30 5,483 5,887 9,230 21,860 1961 T7652 129 4,185 3,998 8,789 34,753 1962 31,187 467 3,399 5,784 11,399 54,192 1963 32,650 Led 22,005 3,485 8,589 67,840 1964 33,627 1,719 18,488 2,045 10,068 65,947 1965 34,073 1,799 34,467 14479 | 115637 83,455 1966 44,083 65335 215793 956 71 13,,,1135) 86,300 1967 88, 187 7,268 27,683 910 13,099 137,147 1968 186,695 4,388 20,366 835 9,416 221,700 1969 214,464 12,665 18,615 881 11,018 257,643 1970 232,829 18,164 41,092 13265 16, 782 310,132 1971 234,337 17,972 32,806 537 14,044 299,696 1972 220 7a, 17,989 15,214 506 16,293 271,749 1973 239,574 19,482 19,913 695 26,821 306,485 1974 347,408 22,505 52,788 1,461 14,861 439,023 1975 364,626 42,786 25,780 2,419 39,514 475,125 1976 318,788 60,455 204,738 2,868 34,191 621,040 1977 988,874 66,605 134,251 2,612) | 33,443 1,225,985 1978 2,701,522 89,626 145,300 33610) 114 752 2,954,810 19794 = 55,493,596 91533 150,000 —= 17,543 5,752,672 Source: Alaska Division of Economic Enterprise (1980). €@value of dollar assumed to be actual value for each year. byalue figures for Prudhoe Bay oil are values at the point where the oil enters the trans-Alaska pipeline and do not include pipeline transportation charges. CAl1 natural-gas values include both dry and liquid gas, including casing-head gas. dpreliminary estimate. Withheld. Table 6 contains summary wage and employment levels, by industry, for 1975 through 1979. This table clearly shows that, during this period, the number of jobs increased only slightly and unemployment remained above the national average. These figures also demonstrate that federal and state governments supply a large proportion of available jobs. In 1973, the year before pipeline construction began, government employment totalled 50% of all wage and salary jobs; from 1973 through 1978 this figure averaged 44%. If active-duty military personnel are not considered, this six-year average is 27%. In fiscal 1977 and 1978, federal spending in Alaska was $1,545 mil- lion and $1,753 million, respectively (Alaska Division of Economic Enter- prise, 1979). Total government spending in Alaska at federal, state, and 44 Table 6 Total Alaskan Wages and Employment by Industry, 1975-79 Wage and Employment Category 1975 1976 1977 1978 1979 Total Wage Payments ($103) 3,426,281 4,237,481 3,781,063 3,596,138 3,775,021 Total Wage and Salary Employment 187,556 197,269 189,229 188,091 190, 8134 Military and Related Federal Civilian Employees --b -- -- 32,001 31,4514 Military Personnel (active duty only) 25,667 25,354 24,958 24,658 24,1984 Military-Related Federal Civilian Employees == = - 7,343 7, 2534 Other Military-Related Employees - - = 7,343 2.2532 Federal Govt. (except military- related in 1978 and 1979) 18, 288 17,944 17,734 10,742 10,6844 State and Local Government 29,247 29,941 31,061 34,122 36,617 Mining 3,790 3,965 4,959 5,562 5,773 Construction 25,735 30, 233 19,545 12,240 10,092 Manu facturing 9,639 10,331 10,845 11,481 12,818 Transportat ion-Communications— Utilities 16,473 15,754 16,028 16,369 16,704 Wholesale Trade 5,908 6,098 5,900 5,721 5,011 Retail Trade 20, 300 21,465 22,553 23,128 23,877 Finance-Insurance-Real Estate 6,030 7,102 7,774 8,228 8,035 Services 25,136 27,670 26,981 27,564 28,345 Farm Workers 200 200 200 255 2554 Miscellaneous 1,143 1,212 691 678 652 Wage and Salary Employment Index (Number of Jobs in 1975 = 1.00) 1.00 1.05 1.01 1.00 1.02 Unemployment Rate 6.9% 8.4% 9.2% 11.0% 8.9% Source: Alaska Division of Economic Enterprise (1980). aPreliminary estimate. b-- = Information not available. local levels was reported to be $2,524 million and $2,845 million for these same years and was estimated to reach $3,147 million in 1979 (Alaska Division of Economic Enterprise, 1980). The strong relative position of government employment, coupled with the slow growth indicated by Table 6 suggests that the state has not yet attained economic stability with strong private-sector support. / Because almost all consumer products must be brought into Alaska, the state has a high cost of living. As examples, costs of food and energy for residents of Anchorage and other Alaskan communities are shown in Table 7. Generally speaking, the variations seen in these data are directly related to distance from source of goods and products and thus reflect the added costs 45 Table 7 Comparative Costs of Food at Home for a Week® and Energy Costs in December 1979 for Anchorage and Other Alaskan Communities Food and Energy Category Anchorage Barrow Bethel Fairbanks Juneau Kenai Nome Petersburg Food $85.80 $154.26 $120.44 $91.92 $77.55 $100.41 $124.62 $91.51 Percent of Anchorage 100 180 140 107 90 117 145 107 Percent of U.S. Average 135 243 189 144 122 158 196 144 Electricity (1,000 kwh) 32.00 121.55 131.80 70.44 --> 45.98 170.00 st Percent of Anchorage 100 380 412 220 me 144 531 = Heating Oil (55 gal) 45.10 92.40 48.40 44.18 45.10 36.85 58.85 44.00 Percent of Anchorage 100 205 107 98 100 82 130 98 Gasoline, Auto (55 gal) 58.30 95.70 57.20 66.00 55.18 68.20 64.90 55.00 Percent of Anchorage 100 164 98 113 95 iy lll 94 Source: Alaska Division of Economic Enterprise (1980). 4Food costs computed for family of four with elementary school children. b-- = Information not available. of transportation. Although wages and salaries are significantly higher in Alaska than in the U.S. as a whole, the differential is not sufficient to compensate for the inflated living costs. As shown in Table 8, recent per- capita income for Alaska residents has ranged from 15 to 65% greater than the U.S. average, but when these values are adjusted for the higher cost of liv- ing, real income falls below the national average by up to 26%. Data for 1974 through 1977 reflect the inflated salaries and wages resulting from pipeline construction activities. When this short-term effect is neglected, the adjusted per-capita income for Alaska falls below the U.S. average by a minimum of 11% and an average of 18% for the remaining seven years tabulated. 3.3 TRANSPORTATION The transportation system of a geographic region is a reflection of the combined needs of the population and the level of economic development. In Alaska, the transportation system is limited because of the low density and unequal distribution of population and restricted economic development. Presently, the system consists of land, water, and air modes, with most land transportation routes in central and south-central Alaska where most of the population is concentrated (Figure 19). The present road system has a total length of about 15,800 km, approx- imately 4000 km of which is paved. The system has more than doubled since 1973, when total road length was about 6500 km. Except for the 1300 km of the Prudhoe Bay haul road and some additional links to the network, much of the mileage increase is due to the construction of multilane highways. The actual area served by the road system is essentially the same as it was in 1937 (National Research Council, 1980). At present, there is still no ground access to western Alaska or the Alaska Peninsula, and the Prudhoe Bay haul road is the only established surface route, other than a few short local links, north of the Yukon River. 46 Table 8 Comparison of Alaskan and U.S. Per-Capita Income, with Adjust- ments for Cost-of-Living Differences Income Parameter 1969 1971 1973 1975 1977 1979 Alaska Per-Capita Personal Income ($) 4,205 4,939 6,046 9,673 10,455 11,252 U.S. Per-Capita Personal Income ($) 3,667 4,132 4,981 5,861 7,038 8,706 Ratio of Alaska Per-Capita Income to U.S. Per-Capita Income 1S 1.20 Loz 1.65 1.49 1 229) Alaskan Family Budget Required for a Moderate Standard of Living ($) 15,485 16,339 18,155 = 23,432, 26,512 += 30,698 Average Urban U.S. Family Budget Required for a Moderate Standard of Living ($) 10,064 10,971 12,626 15,318 17,106 20,517 Ratio of Alaskan Family Budget Requirements to Average U.S. Family Budget Requirements 1.54 1.49 1.44 1253 1.55 1.50 Alaska Per-Capita Income, Adjusted for Family Budget Requirements in Alaska ($) 2,731 3,315 4,199 6,322 6,745 7,501 Ratio of Alaska Per-Capita Income, Adjusted by Higher Family Budget Requirements in Alaska, to U.S. Per Capita Income 0.74 0.80 0.84 1.08 0.96 0.86 Source: Alaska Division of Economic Enterprise (1980). Railroad service in the state is also limited. The federally owned Alaska Railroad has been in service since 1923 and operates over a distance of 756 km from Seward on the Kenai Peninsula, through Anchorage, to the northern terminus at Fairbanks. A connection between Skagway, Alaska, and Whitehorse in the Yukon Territory is the only other rail line in the state. Access to most parts of Alaska, as well as between Alaska and other areas, is primarily by water and air. Barge service is provided to many coastal settlements along the thousands of kilometers of coastline; service to western and northern ports is seasonal. The Alaska Marine Highway System also has scheduled service to 25 coastal communities and connects Alaska with Seattle (National Research Council, 1980). The Yukon and other major rivers also serve as natural transportation routes. Anchorage is the center for air transportation, and there are airports suitable for jet traffic at 25 other locations throughout the state (Federal- State Land Use Planning Commission, 1977). All Alaskan communities with pop- ulations greater than 2500 have daily scheduled airline service (Engleman et al., 1978). Additionally, most of the more than 200 villages have gravel airstrips for air service on an irregular basis. st. aN, Island Qo a “HHH Kilometers © 100 200 300 (Geet 0 100 200 Miles Fig. 19 Major Components of Alaska's Transportation System (after Selkregg, 1974-76; Stenmark, 1978) Key Jet Ports Major Harbors Highway System Railroads Ferry and Barge Routes LY 49 4 LAND OWNERSHIP, JURISDICTION, AND USE The issues of land ownership and use classification are extremely im- portant considerations in determining the use of Alaska's natural resources and the future development of energy and mineral resources. During the 91 years following the acquisition from Russia, Alaska was a territory of the United States and land use policies were dictated by federal laws and govern- mental decisions made in Washington. Few restrictions were placed upon land use and most of the land remained within the general public domain. Alaskan native residents (Aleuts, Indians, and Eskimos) were generally undisturbed in their traditional use of the land, and private ownership was not a signifi- cant issue. When Alaska was admitted to the Union in January 1959, the federal government owned and controlled more than 99% of Alaska's land. The State- hood Act stipulated that title to more than 4.0 x 107 ha (9.9 x 107 acres) of land be transferred to the new state governments. As selection of these lands proceeded and the state attempted to obtain lands previously claimed by natives, the native leaders filed protests that prompted a general freeze in the selection and transfer process. These actions culminated in the passage of the Alaska Native Claims Settlement Act of 1971, which also carried stipu- lations for the withdrawal and preservation of large areas within federal conservation units. Thus began a nine-year period of often heated public and congressional debate over the quantity, location, and ultimate use of these withdrawn federal lands. This so-called "d-2 issue" created a situation in which pro-development and conservation/preservation interest groups assumed adversary positions in regard to the size and management classification of millions of hectares of land. Simultaneously, land transfer to the state and natives was essentially halted and the final distribution of land ownership remained much in doubt. The question of d-2 lands was finally resolved with the passage of the Alaska National Interest Lands Conservation Act in Decem- ber 1980. Redistribution of land ownership within Alaska has been a monumental and complicated task that is still incomplete pending final selections by the state and native groups, as well as completion of federal action to resolve conflicts and discrepancies. Congressional and executive actions undertaken to effect a suitable disposition of ownership and management classification have been instrumental in determining future patterns of energy and mineral resource development. The remainder of this section briefly summarizes the land ownership rights of the federal government, the State of Alaska, and Alaskan natives as related to potential land use and future surface coal mining possibilities. 4.1 EARLY FEDERAL INFLUENCE IN ALASKA With the signing of the Treaty of Cession in 1867, ownership of Alaska was transferred from Russia to the United States in exchange for $7.2 million in gold (Epps, 1978). No consideration was given to the issue of native lands at that time because of the isolated location of the new possession and because there was no compelling need for natives to secure land entitlements 50 at that time. Anticipating potential Indian wars in Alaska, however, Con- gress passed a bill in 1883 that established civil and judicial districts and addressed the rights of natives on the lands they occupied and used. This bill, by deferring to future legislation the terms under which title to these lands could be obtained, provided the basis for the Alaska Native Claims Settlement Act of 1971 (Epps, 1978). In 1884, passage of the Organic Act for Alaska was significant in that the Department of the Army was relieved of its duty to govern Alaska; the general laws of the State of Oregon were applied to the territory. Also im- portant in the passage of this Act was that the U.S. mining laws were ex- tended to Alaska and the area was designated as a land district. Provision was made for railroad rights-of-way and the homestead laws were extended to Alaska in 1898; in 1899 the public land survey system was initiated. The Civil Code for Alaska, which established additional aspects of local govern- ment and property taxation, was adopted in 1900 (Selkregg, 1974-76). A territorial government in Alaska was established through the second Organic Act of 1912, and four years later the Native Allotment Act enabled any Alaskan native to claim 65 ha (160 acres) of land. This was the first time that settlement of native claims was directly addressed in legislation per- taining to Alaska. There was little response to this Act by natives, who were not interested in private ownership of land and who considered that such small plots of land were of little or no use. During the early 1900s, the conservation movement began and grew steadily. In response, the federal government withdrew from the public do- main and designated for permanent federal ownership several large tracts of land in the territory. By 1909, Tongass National Forest had been estab- lished. Subsequently, Mt. McKinley National Park, Glacier Bay and Katmai National Monuments, and several national wildlife refuges were created. A list of all pre-1969 federal land withdrawals that individually exceeded 404,700 ha (1 million acres) is given in Table 9. In addition to executive and congressional actions pertaining directly to land use management and jurisdiction, several other legislative actions will affect future mining activities on federal lands. Relevant legislation includes, among others, the Mineral Leasing Act of 1920, Mineral Materials Act, Multiple-Use Sustained-Yield Act of 1960, Wilderness Act of 1964, Fed- eral Coal Leasing Amendments Act of 1975, and the Federal Land Policy and Management Act of 1976. Norman and Silver (1978) and DeStefano and Foss (1978) discuss in some detail the legal and legislative relationships govern- ing federal lands within Alaska. Naske and Triplehorn (1980) summarize early federal legislation and political decisions that directly affected mining and use of coal in the state. 4.2 THE ALASKA STATEHOOD ACT The passage of the Alaska Statehood Act (P.L. 85-508) of 1958 provided for the transfer of large quantities of land from federal to state ownership. In so doing, Congress intended to provide the new state with a sound economic basis, principally through the natural resources contained within these_land areas. When selection is complete, Alaska will own more than 4.1 x 107 ha, an area larger than the state of California. SI Table 9 Federal Withdrawals of Land before 1969 (Areas of more than 1 million acres only)@ System and Name Acreage National Park System McKinley National Park 1,939,000 Glacier Bay National Monument 2,804,000 Katmai National Monument 2,792,000 National Forest System Chugach National Forest 4,726,000 Tongass National Forest 16,000,000 National Wildlife Refuges Aleutian Islands 2,720,000 Nunivak Island 1,109,500 Kenai National Moose Range 1,730,000 Kodiak Island 1,820,000 Clarence Rhode 2,888,000 Arctic 8,900,000 Reindeer Stations St. Lawrence Island 1,205,000 Military Naval Petroleum Reserve No. 4 23,029,000 Water Power Rampart Project 8,956,000 TOTAL 80,618,500 Source: Selkregg (1974-76). 41 acre = 0.4047 ha. The Statehood Act conveyed to the state government of Alaska (1) 41,481,750 ha, including subsurface rights, of "vacant, unappropriated, and unreserved" lands from the public domain to be selected within 25 years; (2) 161,880 ha of “vacant, unappropriated" federal land from the Tongass and Chugach National Forests for community expansion, development, and recreation purposes; and (3) 161,880 ha from other public lands that were "vacant, unap- propriated, and unreserved," also for community expansion, development, and recreation. In addition to the land grants received at statehood, Alaska had also obtained rights to other federal lands before statehood. These prior selections set aside land for public schools (approximately 70,800 ha), the University of Alaska (40,470 ha), and the Territory's mental health program (404,700 ha authorized to be selected by the Mental Health Enabling Act of 52 1956); details are provided in Meacham (1978). The Statehood Act also ex- tended the Submerged Lands Act of 1953 to Alaska, which entitled the state to submerged offshore lands to the limits of the territorial sea and lands be- neath inland navigable lakes and streams. Transfer of lands to the state involves three steps: selection, tenta- tive approval, and patent. "Selection" does not include state management authority; however, the lands are withdrawn from other forms of entry under the public land laws. Before "tentative approval," publication of notice of the state's selection must be given to provide opportunity for claims of com- peting rights and to allow time for surveying the selected lands. Following "patent" approval, the state assumes management responsibility, and the lands have then been legally transferred. A detailed discussion of the selection and transfer process is given by Meacham (1978). In 1959, Alaska began reviewing and selecting lands from the 1.19 x 108 available hectares of federal public domain and from the 8.4 x 10© ha of national forest lands within the state. Selection was halted in response to native protests and a general freeze in transactions in the late 1960s persisted until Congress passed the Alaska Native Claims Settlement Act in 1971. Since that time, selection and transfer has been slow due to federal designations of national interest (d-2) lands and debate over their location and extent, the complex legal framework involved, and over-selection by na- tive groups. To date, more than 3.2 x 107 ha have been selected and approxi- mately 1.2 x 107 ha have been patented to the state (Kennedy, 1981). Al- though no formal policy of land selection has been presented by the state, it is apparent from the pattern of previous selections that three main goals have been emphasized: (1) obtaining of land to meet existing and future set- tlement requirements, (2) controlling of lands along major highway corridors, and (3) obtaining of lands with a high potential for natural resource devel- opment (McConkey et al., 1977a). 4.3 THE ALASKA NATIVE CLAIMS SETTLEMENT ACT In almost every congressional act affecting land title in Alaska, na- tive land rights were protected. However, the exact nature of land rights of these people was never addressed. Some of the lands selected by the state under the Statehood Act were those. that had been traditionally used and oc- cupied by natives. This encroachment on native areas, as well as the oil and gas leasing royalty payments to the Tyonek reservation, generated native in- terest in the benefits of land ownership. In response, the natives formed the Alaska Federation of Natives and began filing land claims and requesting native claim settlement in the mid-1960s. Eventual settlement was hastened when the Secretary of the Interior placed a general land freeze on the trans- ferral of all public lands in Alaska until a final settlement was reached. As a consequence, oil companies found that petroleum exploration and develop- ment could not proceed until land ownership had been established. Pressure by these companies for settlement of the issue, together with the efforts of native interest groups, resulted in the passage and signing of the Alaska Native Claims Settlement Act, or ANCSA (P.L. 92-203) in December 1971 (Shively, 1978). 53 The provisions of ANCSA were unique, controversial, and complex. In essence, the Act gave Aleuts, Indians, and Eskimos authority to select nearly 1.8 x 107 ha of land from the public domain and provided disbursement of nearly $1 billion ($109) to natives. Instead of individual monetary allo- cations, most of the revenues were to be given over a period of years to 13 native-controlled profit-making corporations. Under terms of ANCSA, an individual who was determined to be eligible for benefits (which requires 1/4 or more native ancestry), was enrolled in a regional corporation and, in most instances, a village corporation. Village corporations were established for villages having 25 or more native resi- dents. Regional corporations were delineated by boundaries based on the old regional native associations that were the forerunners of the Alaska Federa- tion of Natives. Figure 20 shows the location of native villages and 12 regional corporations. The 13th regional corporation was formed to include all eligible natives residing outside of Alaska. The regional native corporations have the same function and responsi- bility as any other corporate entity in Alaska. The corporations are in business to make money for their shareholders (enrolled natives) and have no legal responsibility to provide housing, health, education, or legal ser- vices. However, for each profit-making regional native corporation, there is a nonprofit native regional association that is not involved with management of the lands azd resources, but that receives federal, state, and private funds and grants. These revenues provide some health, educational, and so- cial services to the people within the region (Shively, 1978). Approximately 200 village corporations were established with the op- tion of becoming either profit- or nonprofit-making entities. All chose to become profit-making in order to allow cash distribution to shareholders from profits obtained from their capital base. The village corporations are over- seen to a certain extent by the regional corporation even though the village corporations, located within the regional corporation boundaries, are incor- porated independently of the regional organization. Matters relating to fis- cal management and land transactions associated with the village corporations can be controlled and monitored by the regional corporation (Shively, 1978). The Settlement Act specified that approximately 18 x 106 ha be dis- tributed to natives in the proportions shown in Table 10. Village corpora- tions are entitled to approximately 10.4 x 10 ha of land. A portion of this amount (approximately 1.5 x 106 ha) consists of reservation lands that six villages chose to retain in lieu of other benefits of ANCSA. In this case, natives obtain fee-simple title (both surface and subsurface rights) to the lands. The remaining 8.9 x 10® ha are available to the other villages and will be allocated based on the number of shareholders enrolled in a particu- lar village. The villages will receive surface rights to this land, while the regional corporations obtain the subsurface rights. The village corporations located in southeastern Alaska do not receive land on the same basis as the other village corporations. This is because the Tlingit-Haida Indians of southeastern Alaska received a cash settlement for claims in the mid-1960s. Approximately one township, or 9325 ha, will therefore be allotted to 10 villages in southeastern Alaska. In this case, the regional corporation will receive subsurface rights to the lands, while the villages will receive surface rights (Shively, 1978). POINT Lay Arctic Slope Regional Corp. colite Rive? ie esse, Fig. 20 Native Villages and Regional feanctic VILLAGE ALASKA'S NATIVE REGIONS Corporations (from Nelson, 1979) 9S a Approximately 6.5 x 106 ha will Table 10 Summary of Land Distribu- be available for regional corporation tion Specified by ANCSA entitlement. Regional corporation al- lotments are based on each regional Ent it lement corporation's (excluding the south- es 6 a eastern region, i.e., Sealaska Native Recipvent (10° acres) Corporation) percentage of the land 5 . base of Alaska. The complicated dis- Sitios Gorpeeaienn 22 tribution system is based on what is Regional Corporations 16 termed the "land lost" formula Vill c ety 3 (Shively, 1978). illage Corporations 7 Miscellaneous Select ions© 2.0 The S = e Settlement Act also pro Total 43.7 vides nearly $1 billion ($109) to Alaskan natives. Of this amount, gource: Modified from Shively (1978). $462.5 million was to be appropriated by Congress to the Alaska Native Fund 71 acre = 0.4047 ha. (established in the U.S. Treasury by bReservation lands retained in lieu the Act) for distribution over a of further benefits from ANCSA. period of ll years. — The remaining croy special purposes such as ceme- emia million 1s cerived from a 2% tery sites, historical sites, grants overiding royalty" on all minerals to native groups with less than 25 covered under the Mineral Leasing Act enrolled members, individuals whose of 1920 (primarily coal, oil, and gas) primary place of residence is sepa- that are developed on state or federal ated from a village, etc. lands in Alaska. Money in the Alaska Native Fund is to be subsequently dis- tributed to the regional and village corporations based upon the number of enrolled natives (shareholders) in each region or village. A complete discussion of the complex provisions of ANCSA can be found in Shively (1978). 4.4 THE ALASKA NATIONAL INTEREST LANDS CONSERVATION ACT Aside from providing land entitlements and monetary benefits to native groups, ANCSA also was instrumental in determining the distribution of owner- ship and use classification of the remainder of Alaska's land. Section 17 of ANCSA established the Federal-State Land Use Planning Commission for Alaska and initiated a series of land use studies and decisions regarding the need to protect large tracts of land under federal conservation systems. Section 17(d)(1) authorized the Secretary of Interior to examine all public lands in Alaska to determine if areas should be withdrawn from possible state and native selection to allow preservation in the public interest and to classify or reclassify any withdrawn lands for conservation purposes. Section 17(d)(2) of ANCSA gave the Secretary authority to withdraw up to 3.24 x 107 ha (or more at the discretion of Congress) for study and possible inclusion within the National Park, Forest, Wildlife Refuge, and Wild and Scenic River systems. The d-2 issue has been the subject of extensive consideration and often emotional debate since the passage of ANCSA in 1971. Pro-development groups (petroleum, mining, and other natural-resource-based industries, na- tive corporations, and some state and federal agencies) expressed concern 56 about the future economic growth of Alaska through the utilization of an abundant natural resource base if millions of hectares of land were withdrawn into conservation units with restricted land use. On the other hand, con- servation and environmental groups have argued that the unique and unspoiled environmental setting, with its abundant wildlife and scenic qualities, should be preserved in its natural condition for future generations. These two opposing land use philosophies have provided the central theme of a lengthy series of public and governmental hearings and numerous legislative, administrative, and judicial actions during the 1970s (Norman and Silver, 1978). After debate on several preliminary House and Senate versions, a com- promise measure was finally drafted and the Alaska National Interest Lands Conservation Act (P.L. 96-487) was approved in December 1980. The principal effect of this law was to withdraw and place within the federal conservation systems (National Park, National Forest, National Wildlife Refuge, and Wild and Scenic Rivers systems) more than 4.29 x 107 ha of land (Table ll, Figure 21). Of this total, 1 daecooeanmae 1.8 x 107 ha were placed under the National Park System, 2.2 x 10’ ha were incorporated as National Wildlife Refuges, 2.6 x 106 ha were added to the National Forest System, and the remaining 0.9 x 106 ha were placed under the jurisdiction of the Bureau of Land Management (BLM) . Numerous river segments were also incorporated into the Wild and Scenic Rivers System and approximately 2.3 x 10’ ha, or about 53% of the total 4.29 x 10/ ha, were designated as wilderness areas and placed under the National Wilderness Preservation System. In addition to these actions, Con- gress incorporated into the Act several stipulations and procedures for the administration and management of the new and expanded conservation units. Several provisions were included to (1) limit the impact of the Act upon prior resident rights and existing land use within withdrawn areas and (2) prevent the impeding of alternative land uses and natural resource develop- ment in nonfederal areas adjacent to convervation units. Because of the multiple-use management scheme applicable to National Forest lands, those areas within the Tongass and Chugach National Forests (except within the boundaries of National Monuments) are subject to mineral leasing and mining with the approval of administrative authorities. The same is true for those lands under BLM jurisdiction. However, all areas placed within the National Park and Wildlife Refuge Systems were essentially removed from all future mineral exploration and development activities. Although exception for exploration in certain areas within the conservation units (National Parks and the Arctic NWR excluded) could be granted by Congress at the recommendation of the President, it is extremely unlikely that any coal resources will be developed in the forseeable future on lands affected by this Act. 4.5 SUMMARY In the years following the passage of the Statehood Act in 1958, land status and ownership in Alaska has remained unsettled and has become an issue of singular importance. Unquestionably, this has been a major obstacle to economic growth through development of the state's natural resources. Pas- sage of P.L. 96-487 was a major step toward resolving the land distribution problem and paved the way for completion of state and native selections and 57 Table 11 Land Conservation Units and Acreages Established by the Alaska National Interest Lands Conservation Act of 1980 (P.L. 96-487); Unit Locations and Numbers Are Shown in Figure 21 Category and Area Unit Number Conservation Unit (acres)@ NATIONAL PARK SYSTEM: A. New Areas 1 Aniakchak Natl. Mon. 138,000 Aniakchak Natl. Pres. 376,000 2. Bering Land Bridge Natl. Pres. 2,457,000 3 Cape Krusenstern Natl. Mon. 560,000 4 Gates of the Arctic Natl. Park 7,052,000 Gates of the Arctic Natl. Pres. 900,000 5 Kenai Fjords Natl. Park 567,000 6 Kobuk Valley Natl. Park 1,710,000 7 Lake Clark Natl. Park 2,439,000 Lake Clark Natl. Pres. 1,214,000 8 Noatak Natl. Preserve 6,460,000 9 Wrangell-Saint Elias Natl. Park 8,147,000 Wrangell-Saint Elias Natl. Pres. 4,171,000 10 Yukon-Charley Rivers Natl. Pres. 1,713,000 B. Additions to Existing Areas> il Denali Natl. Park© 2,426,000 Denali Natl. Pres. 1,330,000 12 Glacier Bay Natl. Parkd 523,000 Glacier Bay Natl. Pres. 57,000 13 Katmai Natl. Park? 1,037,000 Katmai Natl. Pres. 308 , 000 58 Table 11 (Contd.) Category and Area Unit Number Conservation Unit (acres)@ NATIONAL WILDLIFE REFUGE (NWR) SYSTEM: A. New Areas 14 Alaska Peninsula 3,500,000 15 Becharof 1,200,000 16 Innoko 3,850,000 17 Kanuti 1,430,000 18 Koyukuk 3,550,000 19 Nowitna 1,560,000 20 Selawik 2,150,000 21 Tetlin 700,000 22 Yukon Flats 8,630,000 B. Additions to Existing Areas> 23 Alaska Marit ime® 460,000 a. Alaska Peninsula Unit b. Aleutian Islands Unit c. Bering Sea Unit d. Chukchi Sea Unit e. Gulf of Alaska Unit 24 Arcticf 9, 160,000 25 Izembek& 26 Kenaih 240,000 27 Kodiak 50,000 28 Togiaki 3,840,000 29 Yukon DeltaJ 13,400,000 59 Table 11 (Contd.) Category and Area Unit Number Conservation Unit (acres)4 BUREAU OF LAND MANAGEMENT SYSTEM: A. New Areas 30 Steese Natl. Cons. Area 1,220,000 31 White Mountains Natl. Rec. Area 1,000,000 NATIONAL FOREST SYSTEM: B. Additions to Existing Areas> 32 Chugach Natl. Forest 1,900,000 33) Tongass Natl. Forest 1,450,000 a. Misty Fjords Natl. Mon. 2,285,000 b. Admiralty Island Natl. Mon. 921,000 TOTAL 106,081,000 4area expressed in acres as per P.L. 96-487; 1 acre = 0.4047 ha. barea tabulated is only that added by P.L. 96-487 and does not in- clude pre-law acreage. CFormerly Mt. McKinley Natl. Park. dFormerly Natl. Monument. Includes all public lands in coastal waters and adjacent seas of Alaska consisting of islands, islets, rocks, reefs, capes, and spires. fIncludes former Arctic National Wildlife Range. 8Formerly National Wildlife Range. hincludes former Kenai National Moose Range. 1Includes former Cape Newenham National Wildlife Refuge. Jincludes former Clarence Rhode National Wildlife Range, Hazen Bay National Wildlife Refuge, and Nunivak National Wildlife Refuge. l St. Lawrence Istand Dec. 2, 1980, P.L. 96-487 [1980]) ee ee Fe Island Kilometers OQ 100 200 300 0 100 200 Miles Areas Withdrawn by the Alaska National Interest Lands Conservation Act of 1980; Numbers Are Defined in Table 11 (Adapted from USGS map Alaska National Interest Lands Conservation Act, 09 61 transfers. Although numerous complexities still exist that may affect smaller parcels of land, the final ownership patterns, and thus potential land use, should be established in the near future. When all legislated transfers are complete, the federal government will retain control of an estimated 60% of the total land area. State government will own approxi- mately 28%, natives will control about 12%, and non-native, private ownership will involve less than 1% of Alaska. Stated and implied policies by both the Alaskan government and native groups suggest that one of the major objectives of past land selections has been to acquire land with good potential for natural resource development. The transfer of additional acreages to these two entities should thus stimu- late planning for the development of energy and mineral resources. In par- ticular, the recent increased interest in the coal resources located on state and native holdings should be sustained and enhanced. Thus, if economic con- ditions are favorable, coal production should increase significantly on non- federal Alaskan lands. 63 5 COAL RESOURCES AND DEVELOPMENT Coal deposits are found in all parts of Alaska from the North Slope to the southeastern panhandle and from the eastern border with Canada to the western coast on the Bering Sea and on the Alaska Peninsula. Because of this widespread distribution, coals are found under a wide variety of environ- mental conditions and geologic settings. The geologic ages of deposits range from Carboniferous to Tertiary and rank varys from lignite through anthra- cite. Although the locations of major coal basins and deposits are known, an accurate evaluation of existing reserves and resources is generally lacking due to incomplete geologic data. Estimates of in-place coal tonnages vary significantly and are subject to revision as additional data become avail- able. However, it is accepted that tremendous quantities of coal exist with- in the state, perhaps as much as that found in the remainder of the U.S. Due to variations in geologic characteristics and environmental conditions, not all Alaskan coal can be mined with existing technologies. Some deposits ap- pear likely to be mined in the very near future, whereas others may be ex- ploited at a later date in response to regional and world-wide energy demands or as liquefaction and gasification technologies advance. Some may never be mined. Coal has been mined in Alaska for more than 100 years; for several reasons, however, production has always been very low in comparison to the potential. As a result of uncertain energy supplies and increasing demands for petroleum that is less abundant and more expensive, and because of other economic and political considerations, Alaska's coal resources have received an increasing amount of attention and consideration since the early 1970s. Several recent events and continuing activities indicate that increased and expanded coal production is probable, if not imminent. This section briefly discusses the geologic and geographic settings for major coal deposits, past and present mining activities, and anticipated future development of Alaskan coal resources. 5.1 COAL RESOURCES The accuracy and reliability of estimated quantities of in-place coal are dependent upon the amount and type of basic geologic data available. Ob- viously, more accurate estimates will be possible with more detailed geologic exploration data. The U.S. Bureau of Mines and U.S. Geological Survey have devised a coal resource classification system based upon the degree of geo- logic identification and economic and technologic feasibility of recovery (U.S. Geological Survey, 1976). Within this system, the term coal resources refers to the "...estimated quantity of coal in the ground in such form that economic extraction is currently or potentially feasible." Coal reserve is "...that part of the resource for which rank, quality, and quantity have been reasonably determined and which is deemed to be minable at a profit under existing market conditions." Because an understanding of this classification scheme and associated terminology is pertinent to interpreting published coal resource estimates, the original publication is reproduced in the appendix to this report. 64 5.1.1 Rank and Grade In addition to the quantity of coal present, the rank and grade of the deposits are important considerations. Coal is classified by rank according to heat content and quantity of fixed carbon, as calculated on a mineral- matter-free basis (Averitt, 1973, 1975). Both the heat content and percent- age of fixed carbon increase from lignite to low-volatile bituminous coal as the quantities of moisture and volatile materials decrease (Figure 22). Generally speaking, older coal deposits tend to be of higher rank than younger ones. Because coals of different rank have different heat values and ultimate uses, variation in rank of large deposits become an important con- sideration in resource evaluation. In this manner, a field containing a lesser quantity of bituminous coal, for example, could perhaps be more eco- nomically attractive for early development than an area containing larger tonnages of lower-rank subbituminous coal or lignite. Coal is also classified by grade primarily according to the content of ash, sulfur, and other undesirable constituents (Averitt, 1973, 1975). Grade therefore becomes an obvious consideration in coal utilization in terms of waste and combustion products and environmental considerations. In partic-— ular, sulfur content is an important characteristic because it lowers the quality of coke and resulting steel products, contributes to corrosion of equipment and the formation of boiler deposits, and results in air pollution from oxides formed during combustion. During mining, sulfur exposed to the atmosphere oxidizes, combines with water to form sulfuric: acid, and produces acid mine drainage that causes significant adverse effects on aquatic and hydrologic systems. Thus coal grade, as well as rank, affects the market- ability of coal reserves. 5.1.2 Estimated Coal Resources The locations of the major known coal deposits in Alaska are shown in Figure 23. In addition to those illustrated, coal outcrops are found at several sites along the Yukon and Kuskokwim rivers, on the Seward Peninsula, at scattered locations throughout the interior, and in the southeastern pan- handle. Very little is known of the extent and geologic continuity of most of these scattered deposits and an accurate evaluation of the quantity of coal present cannot be made without detailed geologic data. Early estimates of the coal resources of Alaska include those of Brooks (1909), which were soon revised (Brooks and Martin, 1913), and those by Wahrhaftig in 1944 (in Buch et al., 1947), which were later revised by Gates (1946). These estimates used geologic information available at the time and therefore gave only a general order of magnitude of the resources on a regional basis. Following World War II, detailed geologic investigations in the principal coal areas were undertaken and resource estimates were modified accordingly. Results of investigations in the Nenana, Broad Pass, Matanuska, Jarvis Creek, and Kenai areas obtained before to the mid-1950s were incorporated into revised state estimates by Barnes (1961) and Averitt (1961). During the 1960s, a detailed reconnaissance survey of the Beluga- Yentna area was completed (Barnes, 1966), and an evaluation of North Slope (SL61 ‘39}4eay wWoIz) (woz}0q) yUeY JUuerezJTG Jo [eoD Jo stskTeuy ojewpxXorg pue (do) sente, Jeoq STSeg-901qJ-19}IeW-TereuTW ‘stow Jo uostTaedwog 77 ‘8Ty PERCENT 3 & & Lignite B % SSI Lignite A ‘ Subbityminous\ ‘ Q te 7 SS Subbituminous B/’ im \ ‘ oO Subbituminous A + Im O igh-volatite c| | bituminous om Oo! High-volatile B bituminous , O High-volatiie A | bituminous \ \ > Medium volatile bituminous \ a Low-volatile bituminous [9] Semianthracite Anthracite Meta-anthracite oO RX SSSA 3 RGSS SSS SSS EIN RSS SSAA RSS ASS SSSA BRITISH THERMAL UNITS g 8 § 4000'01 XN Lignite B \ Lignite A ‘ Subbituminous C Ss \ \ \ Subbituminous B \ \ Subbituminous A 5 KARAS RAMS SSS High-volatile C bituminous High-volatile B bituminous High-volatile A bituminous Medium-volatile bituminous Low-volatiie bituminous \ \ X KASS SE KARAM AAMAS SES MMS ASS RXR AAAS AAAS SSS SSS SSSA 00021 {000'r1 \ \ \ \ ‘nN KRAEMER SSS \ \ \ RIERA RRA SSS SSS ip t SSS \ \ \ SEEMS SSS SSS / / 1 000'9T S9 ° Rampart e Nulato =~ 3 Eagle-Circle \s s Nenana Basin (includes Healy Creek, Lignite Creek, Jarvis Creek, Wood River, Tatlanika, and Teklanika fields) - @ Matanuska Bering '__- 0p? River L/ Kilometers © 100 200 300 ° 100 200 Miles Fig. 23 Location of Principal Coal Deposits in Alaska (after National Research Council, 1980) 99 67 coal was undertaken from data collected during petroleum exploration of what is now the National Petroleum Reserve, Alaska (NPR-A) and adjacent areas. Petroleum exploration and development wells on the North Slope and in the Kenai Peninsula-Cook Inlet areas, as well as increased geologic exploration in other sectors of the state during the 1960s and 1970s, provided valuable information on the geologic setting and areal extent of coal units and thus allowed further refinements in resource estimates. Most of the published information on the geology and coal resources of the individual coal basins and fields, and for the entire state as well, has been reported by the U.S. Geological Survey, U.S. Bureau of Mines, and the Alaska Division of Geological and Geophysical Surveys. Table 12 contains a listing of selected publications that detail the geology and coal resources of most significant known coal occurrences in the state. In addition to these publications dealing with specific areas, reports by Averitt (1970, 1975), Barnes (1976a), Geer (1973), McConkey et al. (1977a, 1977b), McGee and Emmel (1979), McGee and O'Connor (1975b), Sanders (1975a), and Warfield (1967) contain information on the coal resources of the entire state and, in some cases, discussions of specific coal fields as well. References to addi- tional publications on these subjects can be found in Averitt and Lopez (1972), Cobb (1974a, 1974b), and Walker (1976). Using all available information collected before 1967, Barnes (1967a) estimated the original coal resources of Alaska to be 1.18 x lol! G,, or metric tons (1.30 x 10!! short tons). Averitt (1975) used Barnes' estimate and expanded it to 2.40 x 101! t to include all remaining identified and hy- pothetical resources to a depth of 1830 m (6000 ft). That same year, Sanders (1975a) estimated that the resource base, to a depth of 915 m, was 5.30 x 1011 ¢ (5.82 x 101! short tons), more than twice the value reported by Averitt. Tailleur and Brosge (1975) reported that the total resources of the Arctic region alone could be as much as 4.54 x 1012 t. Using data from a number of sources, McConkey et al. (1977b) subsequently presented a range of approximately 1.69 to 4.53 x 1ol2 ¢ for their estimate of the state's coal resources. More recently, McGee and Emmel (1979) estimated the measured, in- dicated and inferred, and hypothetical coal resources of Alaska to be 1.27 x 109, 1.00 x 1011, and 1.61 x 1012 respectively, for a total of approxi- mately 1.71 x 1012 t. This total did not include the estimated 9.07 x loll t beneath Cook Inlet. McGee and Emmel (1979) indicated that more than 95% of the total coal resources are located in northern Alaska, with about 84% of the total located within the boundary of NPR-A. Exploration activities in coal-bearing areas are continuing and, as information becomes available, additional revisions of the estimated coal resource base will be required. Earlier resource estimates indicated that most of Alaska's coal was subbituminous. For example, Sanders (1975a) estimated that only about 4% of the state's coal was bituminous, with the remainder principally subbituminous and lignite. However, based on a ratio of 60% bituminous to 40% subbitumi- nous for coals in NPR-A reported by Martin and Callahan (1978), McGee and Emmel (1979) estimated that Alaskan coal resources are 57% bituminous, 42% subbituminous, and about 1% lignite. McGee and Emmel also noted that most of the easily accessible coal, which is also the coal that will be mined in the near future, is subbituminous in rank. 68 Table 12 Selected Publications on Geology and Coal Resources of Alaskan Coal-Bearing Areas Coal Basin or Field Publication North Slope Basin Barnes (1967b), Callahan (1971, 1975), Martin and Callahan (1978), Sanford and Pierce (1946), Tailleur (1965), Tailleur and Brosge (1975), Toenges and Jolley (1947) Nenana Basin Toenges and Jolley (1949), Wahrhaftig (1951, 1973), Wahrhaftig and Birman (1954), Wahrhaftig and Hickcox (1955), Wahrhaftig, Hickcox, and Freedman (1969) Broad Pass Field Hopkins (1951), Wahrhaftig (1944) Matanuska Field Barnes and Payne (1956), Barnes and Sokol (1959), Toenges and Jolley (1949), Warfield (1962) Bering River Field Barnes (1951), Sanders (1975b) Beluga-Yentna (Susitna) Fields Barnes (1966), Maloney (1958), McGee (1972), Patsch (1975, 1977), Warfield (1963) Kenai Field-Cook Inlet Basin Barnes (1962), Barnes and Cobb (1959), McGee (1975), McGee and O'Connor (1975a), Toenges and Jolley (1949) Information on the quality or grade characteristics of coals sampled at several locations throughout the state is given in Table 13. Most of the samples are ranked as subbituminous except for those from the Broad Pass and Yentna fields, which are lignite. When Alaskan coal characteristics are com- pared with those of coals from the conterminous 48 states (Table 14), it is seen that the Alaskan coals are similar to those found in the Northern Great Plains and Rocky Mountain coal provinces. Coals sampled from all three re- gions are primarily subbituminous with a modest heat value, relatively high moisture, and low sulfur content. Similarly, Alaskan coals -- particularly those in central and southern basins -- tend to occur in thick beds that form a large portion of the stratigraphic section and provide favorable stripping ratios. Additional information on the quality characteristics of Alaskan coals can be found in Conwell (1972), Cooper et al. (1946), Rao (1975, 1980), and Sanders (1975a). 5.2 COAL BASINS AND GEOLOGY Alaska can be divided into three broad geographic regions: the Arctic (area north of the Brooks Range), the Interior (area between the Brooks and Alaska ranges), and the South-Central (area south of Alaska Range) (see Fig- ures 2-4). Each of these regions has a major coal basin, several coal fields, and additional outcroppings of coal units. 69 Table 13 Quality Characteristics of Alaskan Coals Concentration (%) Coal Field and Volatile Fixed Heat Content Number of Samples Moisture Matter Carbon Ash Sul fur (Btu/1b)# Northern Alaska 19.08 30.25 46.49 4.19 0.36 9,859 (2) Nenana® 23.60 34.84 31.70 9.85 0.27 8,129 (7) Nenanad 22.6 41.3 36.4 10.7 0.2 8,530 (typical analysis) Broad Pass? 28.32 33.53 24.08 14.07 0.15 6,395 qa) Matanuska® 3.83 31.98 48.08 16.11 0.41 11, 680 (2) Matanuskad 3.9 38.9 43.0 14.2 0.4 11,650 (typical analysis) Beluga? 23.65 35.20 33.34 7.81 0.14 8,327 Q) - Beluga? 20 40 30 10 0.2 7,500 (typical analysis) Yentna? 29.83 38.78 28.52 2.88 0.08 7,980 (2) Kenai 23.01 35.63 32.71 8.65 0.23 8,028 qa) Multiple Fields® 24.1 34.9 30.2. 10.7 0.2 8,080 (9) 4] Btu/1lb = 2325.8 joule/kg. bpata from Rao and Wolff (1980); analyses on equilibrium-bed-moisture basis. Data from Rao and Wolff (1979); analyses on equilibrium-bed-moisture basis. dProm Geer (1973). Samples from Healy Creek, Chignik, and Herendeen Bay fields. Data from Swanson et al. (1976) and compiled by National Research Council (1980). Table 14 Comparison of Average Characteristics of Coal from Major Coal-Bearing Regions of the Con- terminous U.S. (in percent, except for Btu) Concentration (%) Coal Field and Volatile Fixed Heat Content Number of Samples Moisture Matter Carbon Ash Sul fur (Btu/1b)4 Pennsylvania Anthracite Region 1.4 6.5 79.5 12.6 0.8 12,780 (38) Appalachian Region 2.8 31.6 54.6 11.0 2.3 12,890 (158) Interior Province 7.2 32.2 48.0 12.6 3.9 11,580 (90) Northern Great Plains Province 24.5 31.7 35.4 8.3 1.2 8,400 (40) Rocky Mountain Province 12.9 36.0 42.0 921 0.6 10,400 (86) Source: Modified from National Research Council (1980). Data from Swanson et al. (1976). 4] Btu = 1055.1 joule. 70 5.2.1 Arctic Region The largest concentration of coal in the state is in the North Slope Basin (see Figure 23), which extends essentially from the north flank of the Brooks Range to the Arctic Ocean coastline and from the Itkillik and Colville rivers on the east to Cape Lisburne on the west. These same coal units ex- tend beneath the floor of the Chukchi Sea (National Research Council, 1980). This large basin contains Cretaceous sediments, primarily in alternating sandstone and shale units, that have been folded into alternating, east-west- trending anticlines and synclines in the foothills of the Brooks Range. To the north, the degree of structural deformation diminishes rapidly and the rock units become essentially horizontal. Sanders (1975a) and the National Research Council (1980) report that at least 60% of the region's numerous coal beds are more than 1 m thick, with 3-m beds being common and 6- to 12-m beds known to be present. In some locations, coal makes up 10% of the entire stratigraphic column. Rank of the coals increases from subbituminous (9800 Btu/lb)* in the north to bituminous (11,000 Btu/1b) in the Brooks Range foot- hills (National Research Council, 1980). In the eastern portion of the basin, between the lower Colville and Kavik rivers, additional Cretaceous subbituminous coals and Tertiary lignites have been observed in oil exploration wells. To the west, the Point Hope de- posits contain low-volatile bituminous coals (14,000 Btu/1lb) in a highly de- formed Mississippian section (Sanders, 1975a). This deposit is geologically independent from the younger coals in the North Slope Basin. 5.2.2 Interior Region The coal-bearing strata of the Nenana Basin crop out intermittently in a 130-km-long zone that ranges in width from 1.6 to 48 km and parallels the structural trend of the Alaska Range. The Tertiary rocks of the basin are primarily weakly indurated terrestrial sandstones and siltstones interbedded with and overlying the coal units. These units have been folded and faulted into a series of smaller, separately named basins (Jarvis Creek, Healy Creek, Tatlanika, Hood River, Suntrana, Lignite Creek, Teklanika, etc.). The sub- bituminous coal beds, which range in age from Oligocene to Miocene, vary in thickness from a few centimeters to more than 18 m; outcrops exposing more than 60 m of coal have been reported (National Research Council, 1980). In addition to the extensive deposits in the Nenana Basin, coal units of various age, thickness, and rank crop out at numerous locations throughout the Interior Region. These locations include the area adjacent to a 190-km length of the Kobuk River; a location on the middle fork of the Koyukuk River near Bettles; along the lower Yukon, particularly near Nulato; at several sites along the Kuskokwim River; on the Seward Peninsula; near the confluence of the Yukon and Tanana rivers; and along the upper Yukon River in the vicin- ity of the towns of Eagle and Circle (McConkey et al., 1977b; Sanders, 1975a). *] Btu/1b = 2325.8 joule/kg. 71 5.2.3 South-Central Region The largest accumulations of coals in south-central Alaska is in the Cook Inlet Basin (see Figure 23). The basin is a_ structural trough that contains Tertiary deposits and is about 515 km long and up to 130 km wide; the total coal-bearing area is 31,080 km? or more. Coal units are exposed along the basin margins surrounding Cook Inlet on the northwest, north, and northeast, and along the western portion of the Kenai Peninsula. The coals vary in rank and thickness and range in age from Paleocene to Miocene; total resources of the basin, including those in the deep subsurface beneath Cook Inlet, may reach 1.4 x 1012 t (National Research Council, 1980). Due to local geographic and geologic conditions, the areas of coal ex- posure are separated from one another. Consequently, the region has tradi- tionally been divided into four principal coal fields: Yentna, Beluga, Kenai, and Matanuska (Figure 24). Some authorities combine the Beluga and Yentna fields into one field called the Susitna, and the deposits in the Broad Pass area are considered by some to be a separate field. The Yentna and Beluga fields contain a good grade of low-sulfur, lig- nite-to-subbituminous coal in the Tertiary Kenai Group. Lithologies associ- ated with the coals range from a massive sandstone and conglomerate to silt- stone and claystones. Toward the southern end of the fields, in the drainage areas of the Beluga and Chuitna rivers, conditions are particularly favorable for surface mining because the geologic strata are relatively flat or gently folded and thick coal seams occur near the surface. Some faulting in the area has disrupted lateral stratigraphic continuity. Several coal beds rang- ing in thickness to more than 15 m are found in these fields; one nearly horizontal coal bed 9 to 15 m thick has been traced for more than 11 km along the Chuitna River. Another bed about 15 m thick crops out near Capps Glacier, and other thick units are known to occur in the area (see McConkey et al., 1977b; National Research Council, 1980; Patsch, 1975, 1977; Sanders, 1975a). In the Kenai field, at least 30 coal beds ranging in thickness from 0.8 to 3 m are interbedded with sandstones, siltstones, and claystones in a coal-bearing section at least 1525 m thick. Although some faulting is pre- sent, the units are essentially flat, with dips of 10° or less on the limbs of broad, gentle folds. The coals are subbituminous to lignite, low in sul- fur, high in moisture, and ranging in heat content from about 6000 to 8000 Btu/lb. Despite low overburden ratios, simple geologic structure, and other favorable factors, these coal units appear to be too thin to warrant com- mercial development (National Research Council, 1980). The Matanuska field east of Palmer contains mostly bituminous coals with some anthracite. Geologic information indicates that the coal units vary in quality and thickness and have been extensively deformed through faulting and folding. Although coal was produced from this field from 1914 to 1968, primarily from the Eska and Evan Jones underground mines, a compara- tively small coal reserve is indicated and the complex geologic conditions make mining difficult and expensive (National Research Council, 1980; Sanders, 1975a). 72 Yentna Coal Field Za Matanuska Coal Field cs% Beluga Coal Field Coal Field Kilometers o 100 o 50 Miles Fig. 24 Coal Fields of the Cook Inlet Basin (after National Research Council, 1980) 5.2.4 Other Coal Areas Coal deposits occur in two other areas of southern Alaska that are outside the three main basins. In the southwestern portion of the state on the Alaska Peninsula, there are small coal fields at Unga Island, Herendeen Bay, and Chignik. The Unga Island field contains Tertiary lignite of fairly poor quality. Beds _up to 2.5 m thick in simple geologic structure occupy approximately 104 km2 (McConkey et al., 1977b; Sanders, 1975b). The Chignik field consists of high-volatile bituminous coals in impure beds ranging from 0.3 to 1.5 m in thickness and occurring within a moderately folded and faulted Cretaceous sequence. Little is known of the quality and extent of the resources (Sanders, 1975a). Within the Herendeen Bay field, the largest of the three, the same Cretaceous coal-bearing sequence occurs; Tertiary lig- nites are found to the south. The bituminous coals occur in beds up to 2.1m thick that have been moderately folded and faulted. The extent of resources in this field is essentially undefined. In the southeast, the Bering River field covers about 200 km2 and is located on the Gulf of Alaska approximately 320 km east of Anchorage and 95 73 km east of Cordova. The early Tertiary coals in this field range from low- volatile bituminous in the west to semianthracite in the east. Due to the extremely complex geologic characteristics of the area, an accurate evalua- tion of the coal resources is still not available after 60 years of study by several geologists (Sanders, 1975b). Coals occur within an alternating se- quence of graywacke, sandstone, siltstone, and shale that has been highly de- formed through complex faulting and folding. It is difficult to identify and correlate the coal units because of the extreme lateral variations resulting from thinning and truncations as well as facies changes. Although prior in- vestigations indicated the presence of a significant quantity (hypothetical resource of 3.3 x 106 t to a depth of 915-m) of good-quality coal, the com- plex geology is an obvious detriment to future mining (Sanders, 1975a, 1975b) Additional information on the geology of particular coal fields can be found in the publications referred to in Table 12. A summary of pertinent geologic characteristics of Alaska's principal coal deposits is given in Table 15. 5.3 HISTORY OF ALASKAN COAL MINING 5.3.1 Pre-World-War IL Early records indicate that English explorers were aware of coal de- posits at Coal Cove (Port Graham) on Cook Inlet as early as 1786. Later, coals were of only casual interest to the Russians until the mid-1800s, when samples from several coastal locations, including Cape Lisburne in the Arc-— tic, were collected and examined. In 1855, the Russian-American Trading Com- pany opened the first coal mine in Alaska at Port Graham on the Kenai Penin- sula. Later in the 19th century, whaling ships and U.S. revenue cutters used coals from Cape Sabine and Corwin on the Arctic coast, and coals along the Yukon and Kuskokwim rivers were mined in small quantities for local use by miners and trappers and for riverboats. The Wharf Mine near Port Graham be- gan supplying lignite commercially in 1888. During this same period and into the early 1900s, coal was mined intermittently at several locations in the Homer area. Subsequently, production fluctuated more and eventually dimin- ished as the Pacific Coast markets were supplied by less expensive coals from Washington, California, and British Columbia (McConkey et al., 1977b; Naske and Triplehorn, 1980; Sanders, 1975b). In 1900 the coal land laws (originally passed by Congress in 1873) were extended to Alaska, making it possible for prospectors to file a claim for coal on public lands. Many coal claims in the Bering River and Matanuska fields were filed under this act and were soon found to be invalid due to the lack of a requisite land survey in the Territory. The Coal Act of 1904 re- moved the survey stipulation and about 900 claims were relocated and filed under this act. These claims became suspect however, because of fraudulent activities to obtain larger tracts of land than allowed under the law. What followed was a much publicized national scandal and a political and ideo- logical feud between R.A. Ballinger and G. Pinchot. In response, President Theodore Roosevelt withdrew all Alaska public lands from entry under the coal laws in November 1906. Consequently, the status of Alaska coal lands and Table 15 Geologic Characteristics of Alaskan Coal Deposits Region Basin Field Age of Geologic Strata Rank of Coal Geologic Structure Thickness of Coal Seams Arctic North Arctic coastal Mainly Cretaceous; Subbituminous Flat-Lying 10-ft beds common; 20- Slope plain fields: minor Tertiary to 40-ft beds known; Meade River, most beds greater than Colville River, 42 in. etc. Foothills fields Bituminous Broad folds Other Point Hope Mississippian Bituminous Highly deformed Maximum known thick- ness 6 ft Interior Nenana Healy Creek Tertiary Subbituminous Moderately dipping Considerable variation Lignite Creek (Oligocene-Miocene) fault blocks and between 2.5 and 60 ft Jarvis Creek gentle folds Wood River Tatlanika Teklanika Other® Eagle-Circle Tertiary Subbituminous to Open folds One 22-ft bed bituminous Sout h- Cook Broad Pass Tertiary Lignite Narrow graben 5 to 10 ft central antet Yentna> Subbituminous Flat-lying to 6 to 50 ft; several Beluga to lignite gentle broad folds, beds in excess of 20 minor faulting ft Matanuska Anthracite to Complexly folded 2 to 23 ft subbituminous and faulted Kenai Lignite to Predominantly 2.5-to 10-ft beds Kenai offshore subbituminous flat-lying Other Bering River Tertiary Bituminous to Extremely deformed Unknown; thick pod- Chignik Herendeen Bay Late Cretaceous and Tertiary semianthracite Bituminous and subbituminous Moderately folded and faulted like masses that thin rapidly Numerous beds less than 2 ft thick; composite zones of coal and thin shale interbeds in ex- cess of 8 ft oL Table 15 (Contd) Indicated and Inferred Hypothetical Resources Region Basin Field Resources (Short tons) (Short tons) Probable Mining Method Arctic North Arctic coastal 60 to 146 billion 402 billion to 4.0 Surface mining; possible Slope plain fields: trillion (includes underground mining in Meade River, U.S. Geological Survey permafrost Colville River, estimates for National etc. Petroleum Reserve, : : (NPR-A) plus 22% Foothuiite. fields added for coal outside NPR-A Other Point Hope Unknown Unknown Surface mining Interior Nenana Healy Creek 440 million to 6 billion 8.7 billion maximum, Surface mining; possible Lignite Creek (reserves estimated at based on area and underground mining Jarvis Creek 120 million) outcrop patterns Wood River Tatlanika Teklanika Other# Eagle-Circle Unknown 100 million Surface and underground mining South- Cook Broad Pass 64 million 50 to 100 billion Surface mining Central Inlet yentnab 207 tol 10! 2\bildion 27, billion Surface mining Beluga? Matanuska 108 to 130 million 149 million Surface and underground mining Kenai 318 million in coastal 100 billion (to Surface mining; under- Kenai offshore areas; 200,000 tons of 2000-ft depth) ground mining in selected stripping coal areas Other Bering River Unknown 36 million to 3.6 Surface and underground billion (to 3000-ft mining depth) Chignik Unknown 300 million Small underground mines; Herendeen Bay less than 300 local small surface mines million Source: National Research Council (1980). @Includes coal occurrences at Nulato, Rampart, etc. Late Cretaceous or Tertiary in age and bituminous and subbituminous in grade. bBeluga and Yentna fields are considered to be portions of the larger Susitna field. Nothing is known about the extent of these deposits; they are SL 76 associated land claims were in dispute and unsettled. In an attempt to re- solve the issue, Congress passed the Alaska Coal Act in 1908 (McConkey et al., 1977b; Naske and Triplehorn, 1980; Sanders, 1975a). Due to the inability to obtain patent on coal claims and the unsettled questions of federal coal land policies, the coal industry in Alaska was sty- mied. At this time, Alaskan production supplied only 2% of the coal con- sumed in the territory, with the remainder imported from Washington, British Columbia, Japan, or Australia. Imported coal cost the consumer about $15 per ton whereas local coal could have probably been supplied for approximately $3 per ton. From 1888 to 1914, a total coal production of only 43,526 t (47,979 tons) was reported for the entire territory (McConkey et al., 1977b; Sanders, 1975a). Impetus for coal development came with the passage of the Alaska Rail- road Act and Alaska Coal Leasing Act in 1914. Mines were opened in McKinley National Park and in the Nenana, Matanuska, and Bering River coal fields under the new leasing act. The building of the Alaska Railroad to the Matanuska field in 1916 and the Nenana field in 1918 provided both transpor- tation and a market for new mines. Between 1916 and 1940, annual coal pro- duction increased progressively to 157,850 t (McConkey et al., 1977b; Sanders, 1975a). 5.3.2 Post-World-War IIL Escalated military activities in the Anchorage and Fairbanks areas during and after World War II expanded the local market, and new mines were opened at Healy, Jarvis Creek, Nenana, Costello Creek, Broad Pass, and the Matanuska Valley. Although many individual mines were short-lived, annual production increased steadily to 373,760 t (412,000 short tons) by 1950. Even though the railroad converted to diesel locomotives in the early 1950s, the military and other markets were sufficient to increase production to a peak of about 781,990 t in 1953. Fluctuating and somewhat uncertain market conditions have predominated since the mid-1950s, althouth the reduced market in the Anchorage area resulting from conversion to natural gas has been par- tially offset by increased usage of coal to generate electricity for the Fairbanks area. Annual production peaked at about 840,950 t in 1966; since 1970, annual production has remained fairly stable at about 635,000 t. From 1880 through 1977, a total of only 2.4 x 107 t of coal was produced in the state (McConkey et al., 1977b). A chronology of significant events affecting coal production and uti- lization in Alaska is given in Table 16. More detailed information on this subject can be found in Naske and Triplehorn (1980). The following quote from Naske and Triplehorn (1980, p. 21) summarizes the federal influence upon past coal resource development and explains, in part, the present status of the coal industry in the state. The pattern of commercial coal mining here has had a certain consistency for over 70 years. Whether by President Teddy Roosevelt or President Jimmy Carter, essential policy decisions made by national leaders in the perceived national 77 Table 16 Significant Events in Coal Development and Production in Alaska Year 1786 1855 1862 1879 1888 1898 1900 1902 1904 1906 1911 1912 1914 1916 1919 1922 1924 1940 1942 1943 1944 1946- 1954 Event Capt. Nathaniel Portlock, English trader, finds coal at Coal Cove (presently Port Graham) on the Kenai Peninsula. First Alaska coal mine opened by the Russian-American Company at Coal Cove. First coal mined in S.E. Alaska (Sepphagen mine, Admiralty Island). Whaling ships and U.S. revenue cutters start using coal from the Corwin mines along the Arctic Coast. Wharf Mine opens near Port Graham. Yukon sternwheelers use coal as fuel to transport gold seekers to gold fields. Extension of coal laws to Territory of Alaska. Yukon River steamers convert coal and wood burners to petroleum-fueled engines. Coal Act enacted, allowing coal claim locations without previous surveys. President Theodore Roosevelt closes Alaska public land to entry under coal laws due to Pinchot-Ballinger feud. Cordova "Coal Party": imported coal shoveled into the harbor in pro- test of federal coal policies. Pinchot burned in effigy. U.S. Navy investigates Bering River field. U.S. Congress passes Alaska Coal Leasing Act. Alaska Railroad is built to Matanuska Coal Field. Alaska Railroad reaches Nenana Coal Field. Completion of 4.4-mile railroad spur up Healy Creek; Suntrana mine established. U.S. Navy begins converting its coal-burning ships to oil. Coal used to power dredges and large placer-mining operations near Fairbanks. Nearly all coal mined in Alaska comes from Evan Jones Mine in the Matanuska field and Healy River Mine in the Nenana field. Alaska Railroad reopens Eska Mine. Coal needed for new Army posts and military airfields. Traditional underground coal mining in Alaska gives way to surface mining methods. Usibelli Coal Mine, Inc., begins stripping coal under U.S. Army license. Alaska Railroad replaces coal-burning engines with diesel locomotives; Eska Mine closes in Matanuska field. 78 Table 16 (Contd.) Year Event 1968 Fort Richardson and Elmendorf Air Force Base convert coal-fired steam power plants to natural gas. Matanuska Field shuts down except for small local needs. Golden Valley Electric Association opens mine- mouth power plant at Healy. 1971 Usibelli Mine only remaining commercial coal mine in Alaska. Passage of the Alaska Native Claims Settlement Act and start of d-2 contro- versy. 1973 OPEC oil embargo and severe winter result in oil and gas shortage. Increased interest in and demand for other energy sources, including coal. 1977 President Carter's energy policy includes conversion of utilities and industry to coal, prompting increased interest in Alaskan coal. Pass- age of Surface Mining Control and Reclamation Act. 1980 Passage of Alaska National Interest Lands Conservation Act. 1981 Contract signed to supply Korea with coal from the Usibelli Mine. Source: Modified from McConkey et al., 1977b. Additional information obtained from Naske and Triplehorn (1980) and Sanders (1975a). interest have had prodigious effects on coal mining in Alas- ka. (Granting that the same pattern was operating in the interest of a different nation when the Russian-American Company ordered the Port Graham mine liquidated in 1865, the pattern is more than a century old.) These national deci- sions have affected coal production -- positively or nega- tively -- even more than have the challenges of geology or transportation. Although some small quantities of coal may be mined in some areas for local use, the only commercial coal mine operating in Alaska today is the Usibelli Mine near Healy, about 160 km south of Fairbanks. The mine was founded in 1943 with a contract to supply 9000 t of coal to the U.S. Army (Conwell, 1977). From the time that the railroad spur was completed to Healy in 1922 until the Usibelli Mine opened, production from the Nenana field had been primarily from the underground Suntrana Mine operated by the Healy Creek Coal Company. The Usibelli Company purchased the Suntrana Mine in 1952 and the Vitro Mine in 1970, making Usibelli Coal Mine, Inc., the largest coal mining operation in Alaska. Since the late 1960s, this mine has supplied coal to generate the majority of the power used in Alaska's interior with an annual average production of about 635,000 t (Usibelli Coal Mine, Inc.). Currently, the rate of production is about 700,000 t/yr (Denton, 1982), and estimated reserves are 1.8 x 108 t. The Healy coals occur within three Miocene formations: the Healy Creek, Suntrana, and Lignite Creek. More than 30 individual coal beds have been identified, but mining has been concentrated on three main seams that 79 have average thicknesses of about 4.9 m, 7.3 m, and 12.2 m in the Healy Creek field (Conwell, 1976, 1977). Both the Healy Creek and Lignite Creek fields (Figure 25) are found in east-west-trending synclinal structures. In the Healy Creek field, the north limb of the structure has been truncated by a near-vertical fault, and the beds in the south limb dip at 45°. Although folding has also occurred in the Lignite Creek field, dips are generally less than 20°, which facilitates stripping and extraction. Mining in the Healy Creek field has used a long open-pit method fol- lowing the thickest coal seams down-dip to a depth of about 90 m. Pit ad- vance stops when the stripping ratio exceeds 3:1 (Conwell, 1977). Very little stripping coal is left in this field and the mine operation has moved north into the Lignite Creek field. At the Gold Run Pass Pit (Figure 25) on upper Lignite Creek, mining uses a modified stripping plan. Unit areas of about 10 ha are stripped of overburden and the coal extracted. As mining progresses to the west, overburden is regraded into level benches and seeded for revegetation (Conwell, 1977). The Usibelli Mine opened the Poker Flats Pit in 1979 and will continue to produce coal from both pits in the Lignite Creek basin until reserves have been depleted in the Gold Run Pass Pit. The geologic strata in the western sector of the field are nearly horizontal and a true box-cut type of strip mining can be used. More than one pit is worked in any given year, with coal being ex- tracted from one pit while stripping is carried out at others. Presently, about 95% of the production is from the newer Poker Flats Pit (Denton, 1982). Pits may be expanded in subsequent years until the economic stripping limit is reached. Overburden is either ripped or blasted to facilitate removal. After preparation, this material is loaded onto trucks with front-end loaders and hauled to spoil piles near the pit for regrading and revegetation (Usi- belli Coal Mine, Inc.). The coal is usually blasted and loaded onto trucks by front-end loaders. Trucks then haul the coal to one of two tipples oper- ated by the mine where it can be crushed, sized, and/or washed as determined by coal characteristics and customer requirements. The Lignite tipple (Fig- ure 25), which began operating in January 1982, has unit-train loading capa- bilities and can crush and load more than 3.6 x 106 t/yr. The older Suntrana tipple has sizing or screening capabilities and is located to the southeast in the Healy Creek field. Most of the coal from the Gold Run Pass Pit will be processed by the Suntrana tipple for the home heating market, which re- quires screened coal (Denton, 1982). In 1977, the heavy equipment operated and maintained by the mine in- cluded seven crawler tractor dozers, a production dozer, four front-end loaders with 7.6- to 9.2-m3 buckets, a front-end loader with a 4.6-m3 bucket, 13 end-dumping trucks with 45-t capacity and interchangeable rock and coal boxes, a 68,000-kg lowboy for transporting dozers, and a fleet of pickups and service trucks, as well as numerous drills and other support equipment (Con- well, 1977). In November 1978 the mine began operating a walking dragline with a 56-m mast, 99-m boom, and 25.2-m? bucket (Andert, 1978). This piece of equipment is used to strip overburden in the Poker Flats area and has in- creased the potential annual production of the mine to 1.1 x 106 t or more (McFarland, 1978). The number of personnel at the mine varies both sea- sonally and in response to production requirements. Employment averages 80 za Lignite Tipple Usibelli Mine Usibelli Mine Gold Run Pass Pit Poker Flats Pit Kilometers Oo 5 10 15 a Oo 5 10 Miles Fig. 25 Healy Creek and Lignite Creek Coal Fields and Locations of Current Mining Operations (after National Research Council, 1980) about 70 during winter and increases to more than 90 in the summer when stripping operations are in progress. The summer crew consists of approxi- mately 40 equipment operators, 30 shop personnel, 4 tipple operators, 4 maintenance men, and 10 supervisors (Usibelli Coal Mine, Inc.). The Usibelli Mine began an active reclamation program in 1971 when several test plots were established to examine the revegetation potential of mine spoil piles under various fertilization rates. About 365 ha of pre- viously mined land were seeded in the spring of 1972, and an additional 325 ha were similarly treated the following year (Usibelli Coal Mine, Inc.). A good cover of grasses was reportedly established in each area by the fall following seeding (Conwell, 1977). In addition to the apparently successful revegetation efforts, Conwell noted that wildlife had reinhabited the re- claimed areas. He reported that a herd.of at least 150 Dall sheep was pre- sent in the vicinity of Healy Creek and that caribou had returned to the area after a long absence. Soils in the area tend to be thin, poorly developed, and contain limited quantities of necessary plant nutrients such as phosphorus, potas-— sium, and nitrogen. The sands and clays found in the geologic units that separate and overlie the coal beds in the basin provide an adequate plant growth medium when disaggregated somewhat and supplemented by fertilizer. As 81 a consequence, topsoil segregation, storage, and reapplication are not neces- sary or desirable for successful revegetation, and fertilization is not usually required beyond the second year after seeding (Conwell, 1976, 1977). Current reclamation activities at the mine consist of regrading spoiled material into an essentially flat surface, followed by aerial application of a mixture of seed and fertilizer. During early revegetation efforts, areas were seeded with a mixture of grasses and legumes (principally alfalfa). Hay was harvested from the reclaimed areas in 1976, and initial yields were suf- ficient to suggest that expanded hay production would be possible (Conwell, 1977). However, it was subsequently observed that yields decreased dramati- cally after two years without plowing and reseeding (Denton, 1982). Because these tillage practices are contrary to the primary goal of mined-land rec- lamation, that is, the rapid establishment of a stable vegetative ground cover for erosion control, the proportion of high-nutrient plants used for revege- tation was reduced. More recently, a seed mixture designed to provide a quick cover that consists primarily of grasses has been used; the mixture to be used in 1982 contains no alfalfa (Denton, 1982). Summary information for seed and fertilizer applications at the mine during the last five years is given in Table 17. During this period a total of 335 ha was treated. The successful operation of the Usibelli Coal Mine for almost 40 years is proof that conventional surface-mining techniques can be used in interior Alaska on a year-round basis in an economically viable enterprise. However, Usibelli (1975) has pointed out several factors that make mining in Alaska different from operations elsewhere in the country. The basic differences occur because of the labor market, the remote location of the mine, and cli- matic conditions. The Alaskan labor market is much younger, less experi- enced, and consequently less efficient than that available to mines in the conterminous 48 states. A second labor difference is a higher turnover rate, with the average length of service for all employees of less than cnree years. This situation, coupled with the inexperience of the labor force, re- quires additional training time and costs. However, the turnover problem has improved considerably in recent years (Denton, 1982). Labor costs are about 25% higher at the Usibelli Mine than at western U.S. mines and about 35 to 40% greater than those in West Virginia or Kentucky. Furthermore, because shifts normally work a 50-hour week, the overall wage level approaches twice that of eastern mines (Usibelli, 1975). The remote location of the mine contributes to the higher turnover in employees and adds to the mining costs in other ways. In the past, the com- pany provided housing for almost the entire crew. Recently Usibelli devel- oped a subdivision near the mine property where employees are building their own houses (Denton, 1982). Other costs result from increased freight charges on supplies and equipment and the necessity to maintain a larger inventory of repair parts to prevent long delays in delivery. The remote location also limits the market availability. Coal exported from the Usibelli Mine would require the additional costs of rail transportation to a port facility (Usi- belli, 1975). Climate affects the operation in several ways. Snowfall increases haulroad maintenance; if large accumulations occur, snow must be removed from the pit working areas. Summer efficiency is high due to the long hours of daylight. However, during the winter when only a few hours of daylight occur Table 17 Revegetation Data for the Usibelli Coal Mine, 1977-1981 82 Species Seeded and Fertilizer Applied Canada Bluegrass (Poa compressa) Kentucky Bluegrass (Poa pratensis) Smooth Bromegrass var. Manchar (Bromus inermis) Reed Canarygrass (Phalaris arundinacea) Hard Fescue var. Durar (Festuca ovina var. duriuscula) Meadow Fescue (Festuca elatior) Red Fescue (Festuca rubra) var. Boreal var. Creeping Red Tall Fescue var. Alta (Festuca arundinacea) Creeping Foxtail var. Garrison (Alopecurus arundinaceus) Meadow Foxtail (Alopecurus pratensis) Annual Ryegrass (Lolium multiflorum) var. Common var. Tetraploid Perennial Ryegrass (Lolium perenne) var. Common var. Tetraploid Timothy var. Climax (Phleum pratense) Crested Wheatgrass (Agropyron desertorum) Alfalfa var. Ceres (Medicago sp.) Alfalfa var. Rambler (Medicago falcata) Alfalfa var. Rhizoma (Medicago sp.) Alfalfa var. Romer (Medicago sp.) Alfalfa var. Thor Flemish (Medicago sativa) Alsike Clover var. Aurora (Trifolium hybridum) Clover var. Altaswede (Trifolium sp.) 1977 2.8 2.8 2.1 2.8 2.1 6.3 2.1 3.5 mh Re 4.9 2.1 2.8 3.5 2.1 Application Rate (1lb/acre)4 1978 1979 ~—-1980 1981 4.0 -b 3.5 3.0 - 3.4 - - 352 2.7 2.8 3.0 2.4 - 1.4 - 302 2.7 258 3.0 2.4 2.7 al - 7.2 8.0 5.6 - 1 - = 4.0 - - - 3.0 - - - 2.0 a2. - ell - 2.4 - - - 1.6 - - - - - = 4.0 5.6 8.0 6.3 - 2.4 QaT, 2.8 2.0 352 2.0 2.8 2.0 - 6.0 6.3 ~ - 6.0 6.3 ~ 9.6 - - - 4.8 = = = 2 = = 4.0 3m? 6.7 7.0 4.0 2.4 - - - 83 Table 17 (Contd.) Species Seeded and Application Rate (lb/acre)4 Fertilizer Applied 1977 1978 1979 1980 1981 Perennial Lupine (Lupinus sp.) - 1.6 - - - Cicer Milkvetch (Astragalus cicer) 1.4 - = - ~ Field Pea var. Austrian Winter (Pisum sativum subsp. arvense) 1.4 1.6 - = - Sainfoin (Onobrychis viciifolia) var. Melrose 2.1 4.0 6.7 9.8 = var. Remont = - - = 6.0 Birdsfoot Trefoil var. Leo (Lotus corniculatus) 2.8 1.6 - = = Hairy Vetch (Vicia villosa) 3.5 4.0 6.7 5.6 4.0 Buckwheat var. Winter (Fagopyrum esculentum) - 3.2 - - - Field Mustard (Brassica campestris) 2.1 352 3.4 2.8 2.0 Fertilizer Applied N 160 80 160 160 112 P205 80 40 75 80 60 K 90 30 15 30 30 24c Source: Compiled from information provided by J.D. McKendrick and W.W. Mitchell, University of Alaska, Agricultural Experiment Station, and C. Boddy, Reclamation Director, Usibelli Coal Mine, Inc. 41 1lb/acre = 1.121 kg/ha. b- = Species not included in seed mixture. ©9.6 1lb/acre of sulfur also applied. (see Figure 11), efficiency declines and the required lighting systems in- crease capital and operating costs. Temperature affects operations at Usi- belli more than any other climatic factor. Efficiency of both equipment and men is much lower under extremely cold conditions; lubricants do not function as well and metal fracturing is more common. Maintenance costs are signifi- cantly greater under these conditions and replacement costs rise because equipment must be replaced more frequently. Additionally, heated storage space is required for all equipment when not in use (Usibelli, 1975). Because Usibelli coal is used for heating and electricity generation, demand fluctuates seasonally. Coal shipments range from 26,300 t per month during the summer to 81,860 t in winter. Previously, in order to compensate for this variation, all stripping was done during the summer so that an ample quantity of coal was exposed for extraction the following winter. Since the 84 dragline was placed in service, stripping is carried out the year around (Denton, 1982). Some of the same equipment is converted for use in both overburden removal and coal production. These operational adjustments re- quire less equipment, but some efficiency is lost because such compromise equipment may not be ideally suited for either function (Usibelli, 1975). 5.4 PROSPECTS FOR ALASKAN COAL DEVELOPMENT As discussed previously, interest in Alaskan coal has increased pro- gressively since the early 1970s. Numerous studies have been conducted to evaluate the potential for expanded coal production, the economics involved, and the ramifications of such activities under a number of different sce- narios. Most of these studies have concluded that Alaskan coals hold great promise for near- and long-term development. However, coal production in the state has remained at essentially the same level since 1971, and no addi- tional mining operations have been started to date. 5.4.1 Constraints on Coal Development A number of complex conditions and factors can be expected to influ- ence future development of Alaska's coal resources, just as they have influ- enced past and present coal industry activities. These developmental con- straints can be placed into one of four broad categories: political and legal, socioeconomic, environmental, and technological. Within each category are numerous issues and considerations, many of which could obviously be placed in more than one category. Each of the four categories of constraints is discussed below in very general terms. Detailed information can be found in many of the references cited. Political and Legal Constraints. Future coal mining operations will fall under the jurisdiction of numerous federal and state laws and regula- tions. Most of these controls are directed toward environmental protection, regulation of mineral resource leasing and development, or generation of revenue through taxation and royalty payments (Alaska Division of Geological and Geophysical Surveys, 1979; Alaska Division of Lands, 1974; Brelsford, 1978; Brody and DeVries, 1981; Lowenfels, 1981; National Research Council, 1980, Appendix B; Smith and Tilsworth, 1975; Tillinghast, 1978; Williams, 1978). These statutes are definite considerations when evaluating the feas— ibility of future mining ventures and planning and conducting a successful operation. Two recent federal laws appear to have particular impact on future coal mining in Alaska. The federal Surface Mining Control and Reclamation Act of 1977 (P.L. 95-87) addresses the issue of minimizing environmental damage during and after surface coal mining. The Act established the Office of Surface Mining and Enforcement (OSM) as the principal federal regulatory authority charged with devising and promulgating a regulatory program and establishing minimum performance standards for mining and reclamation. The Act further stipulates that, due to the diversity of environmental conditions in areas subject to surface mining, primary regulatory authority should be with the individual states involved. Recognizing that conditions in Alaska are significantly different from those in the other coal-bearing states, Congress requested 85 that the National Academy of Sciences and the National Academy of Engineering conduct "an in-depth study of surface coal mining conditions in the State of Alaska in order to determine which, if any, of the provisions of this Act should be modified with respect to surface coal mining operations in Alaska" (P.L. 95-87, Sec. 708[a]). This study indicated that several modifications and allowances would be appropriate for Alaskan conditions, particularly those in the arctic. It was recommended that regulatory procedures for oper- ations in the interior should focus on innovative and effective methods of meeting the objectives of the Act and that best standards should be related to results achieved using the best available technology for dealing with special local and regional conditions. Finally, it was concluded that the similarity of conditions in the south-central region (Cook Inlet Basin) to those in certain of the conterminous states is sufficient to warrant initial regulation under the provisions of the Act (National Research Council, 1980, p. xxvii). In accordance with P.L. 95-87, Alaska is developing and adopting a state surface mining program (Brody, 198la). At the time of this writing, this program is being modified and revised for consideration by state legis- lators during 1982 (Brody, 198lb). Future mining will be required to meet the standards and criteria that are specified in this regulatory program. The second federal law that should prove to be a major influence in future coal industry developments is the Alaska National Interest Lands Con- servation Act of 1980 (P.L. 96-487). Details of this Act and the events leading to its enactment were discussed in Section 4. However, it should be restated here that from 1971 until P.L. 96-487 was passed, the status of land ownership and management jurisdiction was unsettled and confused. This sig- nificantly retarded development of Alaskan natural resources, including coal. Even if other developmental constraints had been favorable, it is unlikely that new mining ventures could have been initiated during this period because most public lands had been "locked up" and state and native selections halted. Consequently, land use alternatives were in doubt. Two aspects of P.L. 96-487 should have a favorable influence on coal development. First, the Act placed specified acreages within units of the federal conservation system and therefore eliminated these areas, excepting those placed within national forests, from practical consideration for future mine locations (see Figure 21). Conversely, this action defined where future mining activities could be permitted, namely anywhere in the state not with- drawn under the Act. Second, P.L. 96-487 allows finalization of state and native land selection and title transfer. This will eventually resolve the questions of ownership and jurisdiction for all lands and allow, at the dis- cretion of the owner, an orderly and planned development of the natural re- sources within an area. As stated previously, prior land selections by the state and native corporations indicate that both parties are interested in obtaining lands with high potential for future natural resource utilization. These selections include known coal-bearing areas. Finally, the state and federal political attitude toward coal mining and use can be very important in determining the timing and direction of future developments. The following would clearly stimulate the growth of the Alaskan coal industry: a favorable governmental position and assistance in pursuing and establishing a foreign coal market; a federal and/or state 86 energy policy stressing increased coal utilization; financial assistance and cooperation with industry in developing the costly infrastructure required for coal production and transportation in remote areas; and government sup- port of research and technology development applicable to coal extraction, combustion, and conversion. Socioeconomic Constraints. The socioeconomic factors that must be considered in regard to coal development are those that could be adversely affected by a rapid and large expansion of coal production facilities and those that will exert some measure of control over future developments. A summary of the general socioeconomic conditions found in the coal-bearing regions of Alaska is presented in Table 18. Alaska has a large number of natives living in numerous isolated and scattered villages (see Figure 20); many of these individuals pursue, at least in part, a subsistence living while maintaining traditional values and customs. The villages are generally small and have only rudimentary housing and utilities, as well as minimal community, business, and institutional ser- vices. If a large mining operation and its attendant processing and trans- portation facilities were located near one or more of these villages, a sig- nificant alteration of the local economy and day-to-day functioning could be expected. The influx of a large number of non-native construction and mining workers and their families, increased employment of local village members, possible ground transportation ties with urban centers, and even the creation of a new community for the mining personnel would unquestionably change the socioeconomic environment of the area and perhaps the region. Not all of these changes would necessarily be viewed as beneficial by the local popula- tion. Therefore, a careful evaluation of the existing conditions, as well as predictions of probable changes resulting from coal development, should be undertaken before actual development so that adequate planning with local input can be provided (Olsen et al., 1979; Rutledge et al., 1980). Such an evaluation and careful planning could minimize adverse effects on the local population and prevent "boom-town" effects similar to those experienced by smaller communities in the western U.S. because of rapid energy-resource de- velopment. Due to the lower level of socioeconomic development in rural Alaskan communities, this effect could be significantly more pronounced than in the West. Many non-native residents of the state place a high value on the "Alaskan way of life." These individuals might not view a large mining oper- ation as a favorable development for their area of residence or for the state as a whole. Their input and opinions would also require careful considera- tion and incorporation into the planning for an operation that could be ac-— tive for several decades. Two significant obstacles to be overcome by an expanding coal industry are inadequate transportation facilities and the lack of an adequate market to justify expanded production. At present, only those coal deposits near the railway between Seward and Fairbanks have transportation access. The operating costs and expenditure required to provide transportation for unde- veloped fields will vary considerably depending upon location. The costs of developing a mine and transportation facilities on the North Slope would be much greater than in the interior or south-central regions. Several studies (Bottge, 1975; Clark, 1973; Engleman et al., 1978; Kaiser Engineers, 1977; Table 18 Socioeconomic Conditions in Alaskan Coal-Bearing Regions Field Population Transportation Land Use Native Economy Region Basin Arctic North Slope Other Interior Nenana Other South- Cook Central Inlet Other Arctic coastal plain fields: Meade River, Colville River, etc. Foothills fields Point Hope Healy Creek Lignite Creek Jarvis Creek Wood River Tatlanika Teklanika Eagle-Circle Broad Pass Yentna® Beluga? Matanuska Kenai Kenai offshore Bering River Chignik Herendeen Bay Population 2600 to 3000, of which 87% is native. Popu- lation concentrated largely at Barrow, Wainwright, and Anaktuvuk Pass. Most of region un- inhabited. Popula- tion density 0.04/ mi? (1970). Population 4700 to 5000, of which 48% is native. Most live in scattered small villages. Population density 0.1/mi2 (1970) Population 14,000 to 15,000, of which 7% is native living chiefly in scattered small villages. Population density 1.1/mi? (1970). Local native villages Area is isolated from ground access and has no internal ground transpor- tation system except for a segment of Yukon River- Prudhoe Bay haul road. Air transportation be- tween some coastal com- munities and major cities and villages in the In- terior, mostly for pas- sengers and light haulage. Major freight by ship to coastal localities during ice-free season. No ground transportation system except in Fairbanks and environs and locally in the vicinity of Nome. Alcan Highway terminates in Fairbanks. Alaska Railroad operates south- ward from Fairbanks and serves Healy Creek coal field. River transporta- tion on Yukon and other major rivers. Air trans- port between most commun- ities, mainly for passen- gers and light haulage. Limited ground transporta- tion system mostly in the vicinity of Anchorage and to the north and east. Alaska Railroad serves Cook Inlet area and passes near the Matanuska and other coal fields. Ocean access to Anchorage and other locations. Air transportation between Anchorage and most local communities. Primarily a wildlife habitat. Tundra used mainly as rangeland for caribou. Some moose in major river valleys. Subsistence hunting and fishing. Coal development at Usibelli Mine. Rangeland for Dall sheep, bison, caribou, and moose. Occasional sport hunting. Rural settlement. Local farming in Matanuska Valley and on Kenai Pen- insula. Gas and oil production. Sport hunting, fishing, and rec- reation. Rural settlement. Some timbering locally. Urban development (industrial and residential). Heavy dependence by natives on fish, caribou, walrus, seals, and whales for subsistence. Some use of wild- life for commercial purposes (walrus ivory). Some subsistence hunting and fishing. Native culture and economy preserved in part in small local areas, such as Tyonek village on Cook Inlet. Some subsistence hunting and fishing in Cook Inlet area. Mixed subsistence and cash economy among Aleuts. Source: National Research Council (1980). @Beluga and Yentna fields are considered to be portions of the larger Susitna field. 48 88 Wolff et al., 1973) indicate that large-scale, commercial production of arctic coal was generally not economical at the time the studies were con- ducted. Consequently, the fields of central and southern Alaska, particu- larly those in the south near tidewater, appear more likely for near-term development. However, future operations in these fields will also incur large expenses for required transportation facilities. Ross (1981) estimated the cost of a coal port terminal with a 5-to-9 x 106 t/yr capacity at approx- imately $65 million, excluding land costs. Additional capital would be re- quired in order to provide transportation from the mine to the port. At present, the prospects for a significantly expanded market and in- creased coal use within Alaska do not appear favorable. Most analysts agree the major potential external markets for Alaskan coal are the west coast states of the conterminous U.S. and the Pacific Rim countries of Japan, Korea, and Taiwan. McFarland (1978) reported that coal shipped from Cook Inlet to a Washington-Oregon coastal site by ocean transport was marginally competetive in price ($1.31/10© Btu) with coal that could be supplied by rail from Wyoming ($1.24/10© Btu) (see also Anderson, 1978). However, most reports indicate that additional near-term needs for this market will be filled by western U.S. coal (DeVries, 1981; Edblom, 1981). Similarly, there is agreement that the far eastern market has the largest potential. Japan imports approximately 99% of its oil, 73% of its natural gas, and 77% of its coal (Edblom, 1981). Due to the rising costs and political uncertainties of imported petroleum, several existing generating plants in Japan either have been, or will be, converted to coal. Current plans also call for the construction of a number of new coal-fired plants by the early 1990s (DeVries, 1981; Kazickas, 1981; Murai, 1977) that will dramatically in- crease the importation of steam coal. Kazickas (1981) predicted that annual coal imports for electricity will reach 4.0 x 107 t by 1990, which must be added to an additional 1.0 to 1.3 x 107 t that is imported for other indus- trial uses. Other estimates suggest even larger imports (Wilson-Smith, 1981). DeVries (1981) reported that about 63% of Japan's anticipated coal requirements will not be needed until after 1987 and that only 9% of the pro- jected coal needs for the next 13 years had committed supplies. Conse- quently, that market is still available. Korea and Taiwan are also expected to require escalated steam coal imports in the near future. Kazickas (1981) estimated that within 10 years, Korea will annually import more than 1.0 x 107 t and Taiwan will import approximately 1.5 x 107 t each year; other esti- mates are higher. Historically, most Japanese coal imports have been supplied primarily by Australia and secondarily by Canada, the United States (from east coast ports), and South Africa, with lesser quantities from China and the Soviet Union (Ross, 1981). The presence of Alaska's large quantities of strippable, low-sulfur coal near tidewater, together with the state's geographic position relatively near the Far East, places Alaskan coal in a competitive and even favorable position to capture a significant portion of this future market. Once the market has been secured, and present indications are that it will be in the near future, expansion of the coal industry in Alaska should follow quickly. 89 Environmental Constraints. As in the case of socioeconomic con- straints, there are those environmental conditions that must be protected from adverse impacts due to increased coal production and those that will control or influence future developments. The general environmental condi- tions in Alaska's coal-bearing regions are summarized in Table 19 and dis- cussed in more detail in Section 2. The level of concern for environmental protection and conservation in Alaska was demonstrated during the extensive debate and deliberations over the d-2 issue. The vast areas of wilderness, scenic beauty, and wildlife habitat within Alaska warrant careful mine planning and operational design to minimize deleterious environmental effects. As mentioned earlier, future operations will fall under the jurisdiction of numerous federal and state regulations and legal standards, many of which are directed toward environ- mental protection. Principal among these for coal mining activities are the performance standards associated with Public Law 95-87 and the surface coal mining program to be adopted by the state. The permitting process will govern the three major phases of future operations: environmental assessment, mining, and reclamation. These issues must be addressed and evaluated in the mine plan and permit application, with appropriate measures detailed to indi- cate compliance with clean air and water quality standards (Sturdevant, 1981). Therefore, unless variances are granted or compliance standards modi- fied to reflect the regional differences in environmental conditions in Alaska, as recommended by the National Research Council (1980), future coal operations will be required to meet essentially the same criteria for envi- ronmental protection as operations elsewhere in the country. In some cases this may be much more difficult and expensive than at mines in the 48 conter- minous states. Certain environmental conditions will exert their influences on future mining operations. Geologic conditions will determine basic mining method- ologies, and overburden and soil properties will influence extraction and reclamation techniques. Climatic conditions will affect operations and may require adjustments to standard mining practices. Some of the effects of climate at the Usibelli Mine were discussed earlier in this section. The presence of permafrost and the physical and chemical properties of frozen surface materials may require innovative handling procedures to ensure physical stability of the material and to prevent degradation of surface- and ground-water quality through increased suspended sediment and dissolved solids. Experience with reclamation of surface-mined land in Alaska is essen- tially limited to that at the Usibelli Coal Mine. Although the practices there have apparently been successful, the diversity of environmental condi- tions in the state may preclude the direct application of this experience at future mines located elsewhere. Some of the information gained during re- vegetation efforts and investigations associated with arctic petroleum ex- ploration sites and the trans-Alaska pipeline may be applicable to future mine sites. It is also possible that the results of revegetation studies at selected western coal mines under broadly similar climatic and geologic con- ditions may prove useful for mines in south-central Alaska. Due to the re- ported success of reclamation efforts at the Usibelli Mine and those associ- ated with the construction of the pipeline, many individuals feel that re- vegetation and land reclamation should not be limiting factors in future coal Table 19 Environmental Conditions in Alaskan Coal-Bearing Regions Region Basin Field Climate Permafrost Soils Vegetation Hydrology South- Cook Broad Pass No data Discont inuous Well drained thin Lowland spruce-hardwood Central Inlet permafrost soils with dark forest surface layer Deep permafrost table. Yentna? Normal temperature All areas gen- Well drained Lowland and upland Beluga? range: summer 44° erally free of strongly acid Spruce and hardwood to 69°F, winter -4° permafrost soils forest. Some moist to 40°F; total pre- tundra. cipitation 29 in.; Ground water available snowfall 119 in. in most areas where Matanuska Normal temperature All areas gen- Well drained Bottomland spruce and cares rnoreune wenera ty range: summer 44° erally free of loamy or poplar forest - ‘ i °. a runoff occurs from May to 67°F, winter permafrost gravelly gray ° endl * to September. Many 6° to 42°F; total soils rj - ae i glacier-fed, sediment- precipitation 14 ; ; laden streams. in.; snowfall 69 in. Kenai Normal temperature All areas gen- Well drained Lowland and upland Kenai offshore range: summer 49° erally free of strongly acid spruce-hardwood to 59°F, winter permafrost soils. forest. Some moist 17° to 42°F; total tundra. precipitation 28 in.; snowfall 101 in. Other Bering River Normal temperature Generally free Poorly drained Coastal western hemlock Glaciers extensive and Chignik Herendeen Bay range: summer 43° to 58°F, winter 22° to 39°F; total precipitation 102 in.; snowfall 109 in. of permafrost Normal temperature range: summer 39° to 60°F, winter 20° to 51°F; total precipitation 127 in.; snowfall 59 in. Generally free of permafrost Normal temperature range: summer 34° to 54°F, winter 13° to 31°F; total precipitation 43 in. snowfall 98 in. soils in water- laid materials. Well drained sandy soils de- veloped in vol- canic materials. and Sitka spruce; alpine tundra and barren ground. High brush of willow, alder, birch and wide variety of low shrubs, grasses, herbs, ferns, and mosses. surrounding area of coal deposits. Heavy surface runoff from glacier melt- water. No information on ground water but bed- rock supplies believed to be very limited. No information on ground or surface water 06 Interior Nenana Other Arctic North Slope Other Healy Creek Lignite Creek Jarvis Creek Wood River Tatlanika Teklanika Eagle-Circle Arctic coastal plain fields: Meade River, Colville River, etc. Foothills fields Point Hope Normal temperature range: summer 35° to 66°F, winter -7° to 27°F; total precipitation 14 in.; snowfall 60 to 70 in. Normal temperature range: summer 37° to 73°F, winter -24° to 25°F; total precipitation 11 in.; snowfall 50 in. Normal temperature range: summer 34° to 64°F, winter -36° to -5°F; total precipitation 5 to 10 in.; snowfall 30 in. Normal temperature range: summer 34° to 49°F, winter -16° to 21°F; total precipitation 10 in.; snowfall 36 in. Generally under- lain by discon- tinuous perma- frost up to 100 ft thick Discontinuous permafrost Continuous perma- frost. Thickness ranges from 750 to 2000 ft. Continuous perma- frost. Greatest thickness mea- sures 1168 ft. Well drained brown soils to poorly drained soils with peaty surface layer. Shallow permafrost table. Poorly to well drained soils with peaty sur- face layer. Shallow perma- frost table. Poorly drained soils with peaty surface layer. Permafrost near surface. Poorly to well drained soils with shallow bedrock or permafrost. Upland spruce-hardwood forest; alpine tundra, shrublands, and barren ground. Upland spruce-hardwood forest; alpine tundra, shrublands, and barren ground. Predominantly moist to wet tundra (grasses, sedges, lichens, mosses, and low shrubs). Some alpine tundra. Alpine and moist tundra Ground-water supplies limited. Best reser- voirs are unfrozen al- luvial materials in major river valleys. Streams freeze over during winter. About 80 to 85% of runoff occurs from June through Sep- tember. Streams flowing north from Alaska Range fed by glacial melt- waters. Shallow lakes common along major river flats. Limited ground-water supplies because of permafrost. Streams freeze over during the winter. Approximately 90 to 95% of runoff occurs from June to mid-September. Shal- low thaw lakes abun- dant in coastal plain. SOURCE: National Research Council (1980). @Beluga and Yentna fields are considered to be portions of the larger Susitna field. 16 92 development. However, it should be noted that precise procedures for large- scale reclamation under arctic and subarctic conditions are uncertain. Addi- tional discussion of revegetation and reclamation considerations and research results obtained under Alaskan conditions can be found in Alaska Rural Devel- opment Council (1977), Conwell (1976, 1977), Johnson (1981), Johnson et al. (1977), Johnson and Van Cleve (1976), Lawson et al. (1978), McKendrick (1980), McKendrick and Mitchell (1978), Mitchell (1978, 1979), Mitchell and McKendrick (1975), and Palazzo et al. (1980). Technological Constraints. As noted previously, Alaska has vast coal deposits; much of this, however, cannot be mined economically with existing technology. Disregarding the economic aspects for the moment, the geologic characteristics of these deposits also preclude their development with exist- ing technology. Historically, most U.S. coal came from underground mines located prin- cipally in the East and Midwest. In recent years, a large number of surface mines have been opened, especially in the western states, and currently more coal is produced from surface operations than from underground mines. Sur- face extraction offers the advantage of lower costs (where stripping ratios are small) and more complete removal of the coal deposit. Similarly, early production in Alaska was from underground mines, but the existing Usibelli Mine is a surface operation. During the past four decades, surface mining technology has advanced considerably, and the efficiency and capabilities of earth-moving equipment have increased accordingly. Draglines and power shovels with capacities ex- ceeding 75 m> are available. Large-payload haul trucks, scrapers, dozers, and associated equipment allow the removal and relocation of large quantities of material in a short time. Coal can be cleaned, sized, blended, and other- wise processed to meet the specifications of individual consumers, and im- proved transportation schemes (e.g., conveyor belt systems and slurry pipe- lines) are technically feasible and in limited operation. Coal in all of the major basins in Alaska, including the North Slope, possibly could be mined with present equipment and technology (see Kaiser Engineers, 1977). However, it is doubtful that existing knowledge and ex- perience is sufficient to ensure successful reclamation and to meet other en- vironmental performance standards applicable to a large stripping operation on the North Slope. Furthermore, modifications of current mining techniques would be required to handle large quantities of permafrost material and to develop practices best suited for this extreme environment. In the interior and south-central regions, some potential reclamation difficulties may arise, but it seems likely that existing technology would be sufficient to conduct successful surface mining operations through at least the initial revegeta- tion phase. Alaska also has a tremendous quantity of coal beneath offshore waters; the coal-bearing strata of the North Slope Basin extend northwestward beneath the Chukchi Sea, and large quantities of coal are also located beneath Cook Inlet. These deposits cannot be extracted with present methods. However, recent field experimentation with in-situ coal gasification indicates the potential feasibility of this emerging technology (McConkey et al., 1977a). 93 This procedure, if perfected, would allow these coal resources to be devel- oped in the future. This technology would also offer the added advantage of permitting the converted gas product to be transported economically from the remote Arctic through a pipeline system. In addition to in-situ conversion, surface coal gasification and liq- uefaction technology has advanced to the point that commercial operations may be feasible in the near future if a significant level of research and tech- nology refinement is sustained. These processes have a high potential for future use in Alaska for the production of clean export fuels as an alterna- tive or supplement to bulk coal marketing. 5.4.2 Coal Mining Forecasts The National Research Council (1980) stated that prospects were good that small-scale mining at many scattered locations in the state would in- crease in the future. These operations would supply coal for local use through intermittent mining during perhaps only a few weeks each year or every two or three years. It is unlikely that annual production from any single mine operated in this manner would exceed a few hundred metric tons. Due to the uncertainties involved, future large-scale coal develop- ments cannot be predicted. In 1977, several individuals with good knowledge of Alaskan coal resources and mining technology were asked to rank the 12 largest known fields in the state according to the likelihood of development and to estimate when development might occur. The responses of these indi- viduals were compiled and reported by McConkey et al. (1977a, pp. 91-97), and the following discussion is based upon this information. The Nenana field (classed as a basin by National Research Council [1980] and in this report) was estimated to be the site of expanded produc- tion in the immediate future, in response to increased demand for electricity in the Fairbanks area. The recoverability of this coal has been demonstrated by the Usibelli Mine, which has leases on proven reserves exceeding 1.1 x 108 t. Coals from the Nenana field were not considered to have an immediate chance for export because of the distance to tidewater. Expanded production for the existing mine-mouth plant could be limited due to air quality regula- tions for areas close to national parks. The area considered to have the best potential for development of a new, large-scale mine (4.5 x106 t/yr) was the Beluga-Chuitna area of the Susitna field (see Figure 23). Development was estimated to begin by 1980 to 1985. The major deterrents to development were considered to be the dense vegetative cover that inhibits exploration, complex geologic conditions, low coal rank, environmental considerations, social and political problems asso- ciated with development in the area, and cost of harbor construction. It was stated that a major mine-mouth power plant would probably be built if the mining project became a reality, with a possible second mine established for export production. The third field considered to have favorable potential for intermedi- ate-term development (1985-2000) was Jarvis Creek, situated at the eastern edge of the Nenana Basin. Although this field is small, with stripping 94 reserves estimated at less than 5.5 x 10/7 t, a local group was reported to be drilling and exploring the deposits at the time of the survey (May 1977) and the construction of a mine-mouth generating plant was under considera- tion. The electricity could be sold to Golden Valley Electric Association for use in the Fairbanks area and for operating pumps along the trans-Alaska pipeline. The Matanuska field (see Figure 23) was ranked fourth in the sequence of possible development areas, but no timing estimate was reported. It ap- pears that the opinion for potential mining activities in this field was based principally on the location (adjacent to the Alaska railroad), past production, and the higher rank of the remaining coals. Most comments sug- gested unfavorable prospects for significant production. It was noted that coking-quality coal could be produced with a washing facility and underground operation if economic conditions were favorable. The Herendeen Bay, Chignik, and Unga Island fields were ranked fifth, sixth, and seventh, respectively with possible developments occurring from 1990 to 1995. Data were insufficient to effectively estimate resources and developmental prospects. Comments indicated that, in addition to local use, limited export of coal could be possible, primarily on the basis of the lo- cation adjacent to tidewater and the presence of higher quality coals. The Northern Alaska deposits were ranked eighth, and little chance of development was expected before 1990 or 2000. Even though the resources are very large, the environmental and transportation difficulties, severe cli- mate, and lack of detailed geologic knowledge of the area were cited as major delays to large-scale operations. Potential development could involve trans- porting finely-ground coal in a slurry using oil as the fluid medium. The existing oil pipeline could be used to convey the coal-oil slurry. This pos- sibility would have an added advantage of extending the life of the petroleum fields and the costly pipeline and support facilities. The Bering River, Eagle, Kenai, and Broad Pass fields (see Figure 23) were ranked ninth through twelfth, respectively. The complex geologic con- ditions were cited as the principal deterrent to developing the Bering River field. The latter three fields were considered to have insignificant re- sources of lower quality coals. Although these estimated developments must be viewed as subjective opinions and "educated guesses," they were obtained from recognized authori- ties on Alaskan coal resources. The estimated progression of development seems tenable when all factors influencing the development of a large-capa- city mining operation are considered collectively. In general, conditions since these estimates were made generally support the projected activities. However, some recent events suggest a more optimistic future for the Alaskan coal industry than anticipated in 1977. 5.4.3 Recent Activities A major obstacle to developing an expanded coal industry in Alaska was removed in December 1980 when Congress passed Public Law 96-487 establishing a framework for resolving the confusion and uncertainties of the land owner- 95 ship, jurisdiction, and management alternatives surrounding the d-2 issue. This Act, by placing specific tracts of land within national conservation units, defined the remaining areas available for state and native selection and will ultimately establish ownership of the coal resources. Regardless of the final distribution of patented land, the federal government will retain ownership of the largest proportion of Alaskan coal resources. This is because the National Petroleum Reserve, Alaska (NPR-A) also incorporates a significant portion of the North Slope coal basin, which is estimated to contain more than 90% of the reserves in the state. Based upon limited information, Brody and DeVries (1981) estimated that about 75% of the North Slope coal reserves are owned by the federal government, with ownership of the remaining 25% split between the state and native corpora- tions. They also reported that the state controls most of the land in the Susitna basin except for the Capps area of the Beluga field, which is owned by the Cook Inlet Region, Inc., a native corporation. The state also owns about 75% of the Nenana basin, with the remainder controlled by the federal government (Brody and DeVries, 1981). Thus, most of the coal likely to be developed in the near future is controlled by the state and native corpora- tions. Four federal coal leases were in effect in Alaska as of January 1981. Two of these, covering approximately 770 ha in the Healy area, were scheduled to be transferred to the state within a few months. The remaining two leases were being cancelled due to delinquent payments (Brody and DeVries, 1981). Four preference-right lease applications (PRLA, an application for a lease that will be issued if applicant has discovered commercial quantities of coal) were also in effect in January 1981. One of these, covering 65 ha near Healy, was scheduled for transfer to the state. An “initial showing" was filed in December 1980 on a PRLA covering 1036 ha in the Jarvis Creek Field; a decision to grant a lease or reject the application was to be made in 1981. The final two PRLAs, both for sites near Point Lay, were to be conveyed to either the state or the Arctic Slope Regional Corporation during fiscal year 1981 (Brody and DeVries, 1981). A total of 52 active coal leases covering 41,680 ha of state land were in effect as of November 1980. As shown in Table 20, 63% of the leased acre- age was in the Beluga field, 31% was in the Nenana field near Healy, and the remainder was divided evenly between the Matanuska field and Upper Beluga Lake area. At the same time, there were six applications to convert pros- pecting permits to leases. These involved a total of 6414 ha in the Kenai field. In addition, 415 applications for prospecting permits had been filed during the preceding five years but were not reviewed due to a moratorium on the issuance of new state coal prospecting permits. These applications in- volved 805,173 ha, of which 90% were located in the Yentna-Susitna area (Brody and DeVries, 1981). Table 20 also shows that as of November 1980, the five major lease- holders controlled more than 90% of the total state leased acreage. Of the five major lessees at the time, Usibelli was producing coal in a commercial operation, two (Placer Amex Inc. and the BHW group) were reported to be ac— tively seeking coal markets, and Mobil was reportedly conducting geologic re- search and exploration on the lease resources. The intentions and activities of the AMAX subsidiary were unknown (Brody and DeVries, 1981). 96 Table 20 Coal Lease Areas on State Land as of November 1980, Tabulated by Location and Lessee (in acres)@ Location Mobilb — BHWC Bocd mre —sucm£ ~— Other Total Beluga 23,080 20,471 17,686 800 - 2,390 64,527 Healy —— a -- 12,820 15,832 3,520 32\5.172 Matanuska -- -- -- -- -- 3,170 .. 3,170 Upper Beluga Lake -- -- -- 3,080 -- -- 3,080 Kenai == = = — = 40 40 Total 12,080 20,471 17,686 16,700 15,832 9,120 102,989 Source: Brody and DeVries (1981). 4] acre = 0.4047 ha. bMobil Oil Corporation, Denver. CBass/Hunt/Wilson group, represented by Starkey Wilson, Dallas. dBeluga Coal Company, owned by Placer Amex Inc., San Francisco. ©Meadowlark Farms, owned by AMAX, Cincinnati. fUsibelli Coal Mine, Inc., Healy, Alaska. Native corporations also own lands with good potential for coal devel- opment. Under the entitlements of the Alaska Native Claims Settlement Act and through land trades between the state and federal governments, the Cook Inlet Region, Inc. (CIRI) has obtained ownership of two of the six leases held by Placer Amex Inc. in the Capps deposit of the Beluga field. This par- ticular area of CIRI ownership covers 3335 ha of leased land. Placer Amex has an additional four leases on state land in the Chuitna deposit. Placer Amex Inc. and CIRI have cooperatively conducted a $3.8 million study funded by the U.S. Department of Energy to examine the feasibility of methanol pro- duction from Beluga coals (Brody and DeVries, 1981). At the time of this writing, the report detailing the results of this study is in the final stages of completion. A listing of all state and federal coal leases in Alaska, descriptive location information, and a discussion of relevant leasing policies are given by Brody and DeVries (1981). In December 1980, a 30,000-t test shipment of coal from the Usibelli Mine was shipped from Anchorage. The coal was transported to South Korea (Republic of Korea) by Sun Eel Shipping Company for test blending and burning by the South Korean cement industry (Anchorage Times, 1980; Knowlton, 1980a). A contract was signed in 1981 to supply 7.1 x 10® t of Usibelli coal to South Korea over a 10-year period at an average rate of approximately 800,000 t/yr (Rao, 1981). Initially, delivery was to begin in the late spring of 1982 (Eakins, 1981). However, due to difficulties encountered by Sun Eel in se- curing a buyer and obtaining approval for the transactions from the Republic of Korea government, the shipping date has been postponed until the summer of 97 1983 (Denton, 1982). The coal will be transported from the mine to Seward by the Alaska Railroad and loaded onto ships at that point for transport to Seoul. Construction of a modern coal-handling terminal and shiploading fa- cility is tentatively planned for Seward. Officials at the Usibelli Mine have indicated that their present an- nual production capacity is approximately 1.8 x 106 t, with a capability to expand production to more than 3.6 x 10© t (Eakins, 1981). If the mine con- tinues to produce about 650,000 t for local consumption and produces 800,000 t for export, the total of 1.45 x 10© t is less than existing capacity and considerably less than potential capability. South Korea is in the process of converting several cement and power plants from oil to coal. This conver- sion will require an estimated increase of 6.4 x 106 t in imported coal by 1985 (Knowlton, 1980a). It has been estimated that Sun Eel Shipping could be purchasing as much as 3.6 x 10® t of Alaskan coal annually by the mid-1980s (Anchorage Times, 1981). Because of the existing contract with Sun Eel, at least a portion of this coal would likely be supplied by the Usibelli Mine. Annual production at this mine by the mid- to late-1980s could thus increase to five or six times that of 1980. Although no mine is presently producing coal from the Beluga field, recent activities indicate that mining in that area could be under way in the near future. Two major leaseholders in the area, Placer Amex Inc. and the BHW group (see Table 20), have been actively pursuing a market for their re- serves for some time. The increased and projected needs of Japan to import steam coal were discussed by Patsch in 1975. He also indicated that a large- scale operation producing at least 4.5 x 10® t would be required to justify the large financial commitment by Placer Amex in developing the necessary mining and support facilities. It was reported in December 1979 that the Marubeni Corporation, a Japanese trading firm, was in the early stages of negotiation with the BHW group regarding development of their coal leases (Coal Outlook, 1979). Marubeni was reportedly interested in a potential pro- duction of 3.6 to 9.1 x 10° t annually. In October 1980, representatives of three government-backed Japanese firms met with Alaskan state officials to discuss mining Beluga coal (Knowlton, 1980b). The Electric Power Development Co., Coal Resources Development Co., and Tokyo Electric Power Co. indicated a desire to work with Placer Amex Inc. and the BHW group in examining the eco- nomic feasibility of mining and transporting Beluga coal to Japan. Knowlton (1980b) also reported that a preliminary study by the BHW group indicated that an operation producing 7 x 106 t/yr for 30 years would be economically feasible and would use less than 20% of their lease area. The level of Japanese interest in Beluga coal has remained high. It was reported in January 1981 that the Japanese government intended to fund a study to solve the potential problem of spontaneous combustion of the coal during shipment (Knowlton, 1981). Such a commitment suggests that Japan is intent on using Beluga coal in the near future. In August 1981, Coal Industry News reported that Diamond Shamrock Com- pany had formed a joint venture with the Chuitna Coal Company (which is owned by the Hunt Energy Corporation of the BHW group) to develop engineering, mar- keting, and mining plans for Beluga coal. Diamond Shamrock will manage the 50/50 venture through its subsidiary, the Diamond Alaska Coal Company. Re- cently, Bechtel Corporation was contracted to prepare an environmental impact 98 statement and to develop a mine plan for an anticipated production range of 9 to 23 x 106 t annually (Eakins, 1981). Placer Amex has also continued to develop plans for using and market- ing Beluga coal. Recently a 1200-t coal sample was shipped to Japan for testing and evaluation (Kirshenbaum, 1981). As noted previously, Placer Amex Inc. and Cook Inlet Region, Inc., recently sponsored a study to evaluate the feasibility of converting Beluga coal to methanol. Although the detailed re- sults of this study are not available at this time, indications are that the conversion operation is technologically feasible and that, if the appropriate market can be realized, a commercial operation would be economically viable. Placer Amex intends to continue its plans for methanol production and the ex- porting of steam coal (Kirshenbaum, 1981); one scenario involving both of these uses requires an annual production of approximately 9 x 106 ¢ starting in 1987 or 1988 (Eakins, 1981). Taiwan has also shown an interest in obtaining Beluga coal. In August 1981, a delegation of Taiwan businessmen representing Taiwan Power Company (Taipower) visited Alaska and made stops at the Usibelli Coal Mine and the Beluga coal field (Carpenter, 198la). During the 10-day visit, officials of Taipower, the only public electrical utility in Taiwan, met with representa- tives of both Placer Amex and the Diamond Shamrock-BHW joint venture. Thus, all three of the major coal-importing countries in the Far East (Japan, South Korea, and Taiwan) have expressed an active interest in purchasing coal from Alaska. The potential levels of production from the Beluga field would warrant construction of a separate port facility rather than transporting the coal around Cook Inlet to Seward for shipment. If both Placer Amex and Diamond Shamrock-BHW secure sufficient markets for their coal, it seems likely that the costs of constructing and operating the necessary transportation and loading facilities would be shared between the two organizations. This would have the obvious benefit of significantly decreasing the amount of initial capital investment required from each to export the coal. Mobil Oil Corporation also has a large tract of leased land in the north end of the Beluga field (see Table 20). Production is anticipated from this area by 1993 (Eakins, 1981). Most of the anticipated developments of Alaska coal resources have centered in the Nenana and Beluga fields. However, recent activities in the Bering River field indicate an interest in developing those resources as well. Although the coals in this field are higher-grade bituminous and some anthracite is present, the estimated resources are small in comparison to the Nenana and Beluga fields and the geologic characteristics are apparently very complex (Sanders, 1975b). The Hyundai Corporation of South Korea, in cooper- ation with Chugach Natives, Inc., has conducted one season of exploratory drilling in the area. Preliminary indications are that the geologic condi- tions are not as adverse as previously thought, and current plans are to ex- pand the drilling program during the summer of 1982 (Eakins, 1981). The characteristics of these coals appear to be well-suited for the intended pro- cessing into briquettes for home heating. Consequently, a lower level of production would justify an active operation than if the coal were to be used 99 to raise steam. Apparently, if the results of the exploration activities indicate that at least 900,000 t can be produced annually at a reasonable cost, development will proceed (Eakins, 1981). At present, the majority of the Bering River field is within the Chugach National Forest. However, Chugach Natives, Inc., has selected this area as a portion of the entitlements under the Alaska Native Claims Settle- ment Act. Thus, ownership of the coal is in question at this time; however, opinion seems to be that the native corporation will soon receive title to the area (Eakins, 1981). Although mining could be permitted on National Forest land under the multiple-use management scheme, it appears that the probability of production, as well as the timing of such an operation, would be enhanced under native ownership. In addition to these major developments, other coal exploration and drilling activities elsewhere in the state have been reported. During the summer of 1981, Canadian Superior Exploration conducted a drilling program in the vicinity of the "Little Tonzona Coal Occurrence" (Rao, 1981). This area is located north of the Alaska Range about 56 km southwest of Mt. McKinley National Park and 40 km northeast of Farewell (see McConkey et al., 1977b, p- 179). At the same time, drilling activity was reportedly under way in a portion of the Chignik field on the Alaska Peninsula (see Figure 23) (Rao, 1981). The results of these exploration activities are unknown at this time. 5.4.4 Summary Despite the vast coal resources in Alaska, coal production has histor- ically been relatively low, for a number of reasons. Present indications, however, are that this trend will be reversed in the near future. Most of Alaska's coal is located in the North Slope basin. The fed- eral government will retain ownership of the majority of these resources by virtue of the location of the National Petroleum Reserve in Alaska. Because of the remote location and lack of transportation facilities, harsh and un- usual environmental conditions, and prohibitive economic factors, it is ex- tremely unlikely that arctic coal will be used in the foreseeable future other than for possible small-scale, local consumption. If economic condi- tions change and coal extraction and/or synfuel technologies advance, it is possible that this tremendous resource could eventually be developed and the coal transported out of the region through the pipeline network already con- structed to transport petroleum from the North Slope. There is general agreement that near-term increases in coal production will occur primarily in expanded operations within the Nenana basin and in new operations in the western portion of the Cook Inlet basin. Other devel- opment may occur in smaller fields along the southern coast. Most of these lands are, or soon will be, owned by the state or native corporations. Many of the political and legal decisions and issues presently facing the state government can significantly affect the timing and degree of coal development in these areas. Decisions on the state surface mining program, taxation, royalty rates, environmental regulation, and other related topics could either enhance or retard expansion activities by the coal industry. 100 Efforts to secure an export market for Alaskan coal have been under way for several years. Changing internal political policies and increased energy requirements in far eastern countries, political uncertainties and escalated costs associated with world petroleum marketing, and other inter- national trade relationships and considerations have placed Alaska in a very favorable position to provide Pacific Rim countries with a steady, long-term supply of coal. The recent contract to supply Usibelli coal to South Korea, as well as continuing discussions with Japan and Taiwan regarding Cook Inlet coal, suggests that large quantities of coal may be mined and exported in the very near future. State officials recently indicated that these developments should occur by 1986, and that by the mid-1990s Alaska could be exporting more than 2.7 x 107 t of coal annually (Carpenter, 1981b). This estimate does not include production that might be required by potential coal conver- sion or synfuels operations in the state. A large-scale operation of this type could consume several million tons as feedstock each year. The significant intensification of geologic exploration and evaluation activities in many different coal fields adds credence to the optimistic pro- jections of expanded coal mining and production. Although a number of legal and political, socioeconomic, environmental, and technological impediments must be addressed and resolved, the next few years should be a time of rapid growth and expansion for the coal industry in Alaska. 101 6 CONCLUSIONS AND RECOMMENDATIONS 6.1 CONCLUSIONS This report has presented an overview of the coal resources of Alaska; past and present coal mining operations; some of the environmental, socio- economic, political, and technological considerations related to developing those resources; and recent activities directed toward expanding future coal production in the state. The following general conclusions may be drawn: e Alaska has a tremendous amount of coal, possibly as much as that found in the 48 conterminous states. Total re- source estimates exceeding 4.5 x 10!2 t have been re- ported. Most of the resources are concentrated in three large geologic basins: the North Slope Basin and associ- ated strata extending northwestward beneath the Chukchi Sea, the Nenana Basin of the interior, and the Cook Inlet Basin in south-central Alaska. Smaller coal fields and other coal occurrences of unknown extent are scattered throughout the state. Perhaps 90% or more of the total resource base is found on the North Slope. Although coals ranked from lignite through anthracite are known to occur within the state, most deposits are subbituminous to bituminous with relatively high moisture and ash con- tent, moderate heat value, and very low sulfur content. Large quantities of such coal are well suited for surface mining operations. e Environmental conditions in Alaska are significantly dif- ferent from those found in the conterminous U.S. and, due to the great geographic diversity, broad variations exist within the state as well. Therefore, many mining and reclamation techniques developed in the "lower 48" are not directly applicable to Alaskan operations. Innova- tive planning and operational modifications will un- doubtedly be required to ensure protection or restoration of wildlife and fish habitats, tundra and taiga ecosys- tems, hydrologic balance (both quantity and quality), and other critical factors of the arctic and subarctic envi- ronment. Furthermore, each of the three major coal ba- sins has its own set of environmental characteristics. This will require individualized considerations for envi- ronmental protection and restoration during mining and may prohibit the direct transfer of technologies and operational procedures between basins. e Aside from the basic geologic setting of the coal-bearing strata, the climate and the presence of permafrost are probably the most significant environmental factors that will influence future coal developments. These factors may require specialized material-handling techniques, special scheduling of certain activities, and perhaps the development of new mining technologies and equipment 102 design. The net effect of these harsh and unique condi- tions is a decrease in overall mining efficiency and an increase in both operating and capital costs of coal pro- duction as compared to operations in the conterminous 48 states. These factors also will partially determine the environmental effects of mining and success of reclama- tion through their combined influence, as reflected in the short growing season, diminished rate of plant nutri- ent formation and cycling, slow rate of geochemical and soil-forming processes, distinct hydrologic regime, slope stability and soil engineering problems due to freeze- thaw activity and permafrost conditions, and related unique site characteristics. Most coal fields are located in remote areas of very low population density and no established surface transporta- tion system. Roads or railways will be needed in order to open a mining operation in these areas and to deliver coal or coal products to the consumer. This will add significantly to the costs of such an operation. Careful planning of future coal mining will be required to minimize adverse effects on the local population. Most coal-bearing areas of the state have a significant population of natives with at least a partial subsistence economy, so care must be taken to ensure minimal environ- mental disruption of extensive areas of wildlife habitat, migration corridors, and aquatic conditions. Further- more, these natives maintain a unique social and cultural environment that could be dramatically altered by coal development and its attendant local and regional socio- economic changes. Planning will also be required to prevent "boom town" conditions in rural areas and small villages if significant social changes and population shifts occur as a result of coal development. The passage of the Alaska National Interest Lands Conser- vation Act of 1980 effectively resolved the d-2 issue and removed a major obstacle to future coal development ac-— tivities. This act will allow resolution of the confu- sion and uncertainties of land ownership and management alternatives and will permit the remainder of state and native land entitlements to be fulfilled. This will allow coal in the selected areas to be mined in a planned and orderly manner. Much of Alaska's coal cannot be extracted with existing technology. This is particularly true of those deposits beneath the Chukchi Sea and Cook Inlet. It is also ques- tionable whether existing technology is adequate to ef- ficiently surface mine and properly reclaim extensive areas of arctic tundra. Additional technological ad- vancements and demonstrated capabilities will be required before a large-scale stripping operation could logically be anticipated for the North Slope. 103 e Essentially the only surface coal mining and land recla- mation experience in Alaska is that gained at the Usi- belli Mine in the Nenana basin of the interior. It is possible that some revegetation results obtained during restoration activities following construction of the trans-Alaska pipeline could be transferred to coal min- ing, but this has not yet been demonstrated. It may also be possible to apply techniques developed in certain western states to operations in the southern portion of Alaska, although this also remains questionable. Limited research results and experience at the Usibelli Mine in- dicate that revegetation should be possible. However, it must be stressed that successful reclamation to a desired post-mining land use entails much more than simply re- establishing an adequate vegetative cover. Land reclama- tion may be defined as the establishment of an environ- mentally compatible, functioning ecosystem with economic or ecological utility that may or may not be the same as that before mining. With this in mind, reclamation can be viewed as a complex process composed of a series of actions and resulting reactions, each with characteristic rates. Successful land reclamation can be accomplished within a reasonable period only when restoration actions are planned and feasible and resulting reactions are pre- dictable. Many years may be required after initial res- toration efforts to accurately evaluate the degree of reclamation success. e The successful operation of the Usibelli mine for nearly 40 years demonstrates that coal can be surface mined under Alaskan conditions and that such an operation can be profitable. There are several circumstances that make mining here more expensive than analogous operations in the 48 conterminous states; however, this is to be ex- pected. When other facets of the Alaskan economy are ex- amined, costs of goods and services are significantly greater than in the “lower 48," primarily due to higher shipping and transportation charges. Future mining oper- ations will require positive marketing advantages to com- pensate for the increased costs of producing coal. e Due to numerous environmental, socioeconomic, and techno- logical constraints, it is extremely unlikely that a large-scale, commercial surface mining operation will be developed on the North Slope in the foreseeable future. Increased near-term production will come primarily from expanded operations in the Nenana basin and from new op- erations in south-central Alaska where developmental con- straints are less limiting. Recent activities suggest that possible production could occur in the Beluga, Yentna, Jarvis Creek, and Bering River fields and that production could increase at the Usibelli Mine operations in the Lignite Creek field. 104 e Present indications are that Alaska is on the verge of a rapid and pronounced increase in coal production. Japan, South Korea, and Taiwan have all shown an active interest in importing large quantities of Alaskan coal in the near future. These countries are also investing developmental funds and participating in geologic explor- ation and evaluation activities within the state. Le current negotiations are successful, Alaska could be ex- porting several tens of millions of tons of coal annu- ally to these countries by the 1990s. Planning and de- sign of new mines to supply this coal could begin almost immediately. e Scientific data and an understanding of fundamental natu- ral relationships and processes in Alaska are limited and inadequate. In evaluating the applicability of the Sur- face Mining Control and Reclamation Act to Alaskan condi- tions, the National Research Council (1980, p. xxiii) concluded that "...the scientific base for much of Alas- ka, particularly the permafrost area, (is) inadequate to comply with the permitting requirements of the Act and inadequate for a predictive understanding of the response of Alaska's complex natural environments to mining and relamation." A concentrated effort is required to col- lect and evaluate the necessary basic data and informa- tion so that development of the state's coal resources can proceed in a timely and environmentally sound manner. 6.2 RECOMMENDATIONS Due to the relative paucity of scientific data and detailed research results for Alaska, together with the potentially complex, long-term effects of large-scale surface mining on the physical, biological, and socioeconomic environment, an almost endless list of research needs could be presented. A sampling of recommendations for specific research topics and other informa- tion relating to pertinent research activities for the state can be found in Alaska Institute of Water Resources (1980); Alonso and Rust (1976); Carlson and Butler (1973); Harlan (1974); Lawson et al. (1978); National Research Council (1980, 1981); Polar Research Board (1975, 1977, 1981); Rao and Beist- line (1981); Resource Development Council for Alaska, Inc. (1981); Scott (1976); Washburn (1980); Wolff et al. (1973); Zemansky et al. (1976); the un- published proceedings of the “Arctic Environmental Workshop" held in German- town, Maryland, in February 1978 under the sponsorship of the U.S. Department of Energy; the numerous publications resulting from the Tundra Biome Studies of the International Biological Program; the "Current Research Profile for Alaska," which is updated and published annually by the Arctic Environmental Information and Data Center in Anchorage; and the many reports published by the Cold Regions Research and Engineering Laboratory of the U.S. Army Corps of Engineers. The following abbreviated list of recommendations covers only those broad areas of research needs related to anticipated surface coal mining and land reclamation that appear to be of highest immediate priority. It should 105 be stressed that these recommendations are made to address real information gaps and research needs perceived by the authors during the course of this investigation. Therefore these research efforts are not suggested merely to satisfy scientific curiosity or to increase the body of general scientific knowledge. Rather, they suggest applied work that, if carefully planned, conducted, and reported, should benefit the State of Alaska, the coal indus-— try, and environmental concerns. Because these recommendations are broad in scope, no presentation is made of specific hypotheses, objectives, or method- ologies. Obviously, future research projects will require careful coordina- tion, planning, and design; existing information should be used whenever pos- sible to prevent redundant and unnecessary data collection. Ideally, future research would be coordinated and evaluated by a single organization to avoid duplication and to ensure the relevance and timeliness of the research effort. e A long-term, interdisciplinary program should be estab- lished immediately to collect, analyze, interpret, and report baseline scientific data on fundamental environ- mental conditions and relationships within the principal coal-bearing regions of Alaska. Long-term data are re- quired to define temporal variations and process rates characteristic of individual geographic areas. It is recommended that this program be developed around the research-watershed or drainage-basin concept for major physicochemical investigations, with latitude to use a more restricted or expanded study area as necessitated by specific research objectives. A program of this type should be under way long before expanded mining opera- tions begin, so that the resulting information can be used in preparing sound mining and reclamation plans and in evaluating the success of reclamation efforts follow- ing mining. Without an adequate knowledge of pre-mining conditions, it will be impossible to objectively evaluate the probable effects of mining and reclamation techniques on specific environmental conditions and to determine the rate at which the pre-mining condition is approached. It is further recommended that this program should concen- trate initially on unmined areas in the Nenana basin and in the Beluga area where future mining operations are ex- pected to develop first. e It is particularly important to obtain the essential data with which to estimate the long-term effects of surface mining on the arctic tundra ecosystem. Although some studies have been conducted to define various aspects of this unique environment, our understanding is far from adequate to accurately predict the effects of large-scale disturbances and to ensure adequate site restoration. Although coal extraction is not anticipated in this part of the state for many years, the very slow rate of natu- ral processes will require a lengthy observation and study period to quantify equilibrium dynamics and to evaluate future system responses to, and rate of recovery from, disturbances. Within the same context, the effects 106 of underground mining and coal conversion activities, particularly as these activities might alter the perma- frost thermal regime, should also be investigated and evaluated. An effective environmental-data storage and retrieval system should be established and continually updated to facilitate information transfer and availability. Al- though some computerized bibliographies are presently available for certain regions of the state, there is no centralized source of, or methodology for, obtaining available data and analytical results. It is recommended that the University of Alaska, perhaps through the Arctic Environmental Information and Data Center, assume the lead role in this effort. It is also recommended that a similar storage and retrieval system be developed to con- tain current information on all federal, state, and in- dustrial organizations and agencies involved in the de- velopment of Alaskan natural resources, as well as the function and activities of each, key individuals in- volved, and other pertinent information. It is recommended that small-scale surface-mining opera- tions within small representative areas be conducted on an experimental basis with detailed monitoring and evalu- tion in order to test and develop methodologies for min- ing and reclamation under the unique environmental condi- tions of the North Slope. If small operations are de- veloped to supply local native villages with coal, as some individuals have suggested, these could be used for testing, demonstration, and evaluation purposes. Experi- ence in surface mining and reclamation under arctic con- ditions is essentially nonexistent. Operations at the Usibelli Mine should be studied to de- termine if mining and reclamation techniques used there could be applied at other Alaskan locations. Even though this is the only large-scale surface coal mine in the state, very little detailed scientific and engineering information is readily available on material-handling methods, permafrost problems and solutions, quantitative reclamation results, and effects of the operation upon specific components of the environment. A detailed study and monitoring program could provide needed information to estimate effects of future mining and reclamation ac- tivities and to indicate possible alternative techniques to improve site restoration at the same or lower cost. This information could also be used to formulate and/or evaluate future environmental regulations and prevent un- necessary, costly, and excessively stringent performance standards while simultaneously preserving environmental quality. 107 e A geologic evaluation program should be initiated to col- lect and analyze the necessary data in order to accurate- ly characterize and delineate Alaskan coal resources. Current geologic information is inadequate to accurately define the total coal resource base for Alaska. Although the location and general characteristics of coal occur- rences are known, data are insufficient to determine the three-dimensional characteristics of the stratigraphic units. e An expanded effort should be undertaken to investigate and demonstrate revegetation techniques. Because of the time required to evaluate revegetative success under field conditions, research should begin as soon as pos- sible. Although some revegetation research results are available and other studies are in progress, much more information on plant species, overburden characteristics, soil-plant relationships, fertilizer requirements, sur- face treatments, growth rates, management requirements, and numerous other factors are required in order to de- velop Alaskan revegetation technology to the point that the desired cover is produced with the best species in the shortest time for the least cost. Invest igations within such a program should also be directed toward the establishment of vegetation with high agronomic potential under the given climatic and soil conditions. e Suitable agricultural land is in short supply in Alaska, and the costs of clearing and preparing additional areas for crop production or pasture are extremely high. The initially large yields of hay from revegetation areas at the Usibelli Mine suggest that agricultural production may be a viable post-mining land use. Therefore, it is recommended that a study be undertaken to evaluate the feasibility of, and develop a methodology for, reclaiming mined lands for agricultural uses. This investigation should consider those factors noted in the previous rec- ommendation, as well as special material-handling and re- placement techniques, agronomic management practices that would improve the productivity of the minesoil, and the economics of such a reclamation program. If this ap- proach appears practicable, the study should concentrate on devising appropriate reclamation methods applicable to conditions within the Nenana basin, as well as to those in the Beluga and Yentna fields, where the growing season is longer and the climate less severe. e The effects of large-scale surface mining on Alaskan wildlife are unknown and need to be investigated. This issue is of particular significance because many aspects of the diverse wildlife populations and habitats have no counterpart elsewhere in the U.S. and because a large proportion of the Alaskan populace engages in commercial, recreational, or subsistence uses of wildlife. 108 The effects of surface mining and reclamation upon the hydrologic conditions of an area and region should be determined. The hydrologic regime in arctic and subarc- tic regions with extensive permafrost is significantly different and more poorly understood than hydrologic response and processes in other environments. Due to the complex changes in the quantity and quality of both sur- face and ground water that could result from surface min- ing, both pre-mining conditions and potential effects upon the hydrologic system should be thoroughly investi- gated and quantified. 109 REFERENCES Alaska Division of Economic Enterprise, The Alaska Economy Year-End Perform- ance Report 1979, Alaska Department of Commerce and Economic Development, Juneau (1979). Alaska Division of Economic Enterprise, The Alaska Statistical Review 1980, Alaska Department of Commerce and Economic Development, Juneau (1980). Alaska Division of Geological and Geophysical Surveys, Mining Laws Applicable in Alaska, Alaska Department of Natural Resources, Division of Geological and Geophysical Surveys Information Circular 14, Anchorage (1979). Alaska Division of Lands, Regulations and Statutes Pertaining to Mining Rights on Alaska Lands as Contained in the Alaska Statutes and the Alaska Ad- ministrative Code, Alaska Department of Natural Resources, Juneau (1980). Alaska Institute of Water Resources, Alaska's Water Resources Problems and Research Needs, Five-Year Research Plan, University of Alaska, Fairbanks (1980). Alaska Rural Development Council, A Revegetation Guide for Alaska, Rural De- velopment Council Publication No. 2, Cooperative Extension Service, Univer- sity of Alaska, Fairbanks (1977). Alonso, W., and E. Rust, The Evolving Pattern of Village Alaska, Federal- State Land Use Planning Commission for Alaska Report No. 17, Anchorage (1976). Anchorage Times, Ship Pushes through Ice, Coal Loaded for Korea, Anchorage, p. A-l (Dec. 22, 1980). Anchorage Times, Korea to Buy Alaska Coal if New Loading Docks Built, Anchor- age (Feb. 20, 1981). Anderson, D., Feasibility Study of Mining Alaska Coal and Transportation b Slurry to the West Coast, U.S. Bureau of Mines Open File Report No. 17(1)-78, Juneau (1978). Andert, P., Alaska's Largest Dragline Begins Operation, The Frontiersman, Palmer, Alaska, p. 12 (Nov. 22, 1378) Averitt, P., Coal Reserves of the United States - A Progress Report, Januar 1, 1960, -S8. Geological Survey Bulletin 1136, Washington, D.C. (1961). Averitt, P., Stripping-Coal Resources of the United States - January 1, 1970, U.S. Geological Survey Bulletin 1322, Washington, D.C. (1970). Averitt, P., Coal, in Brobst, D.A., and W.P. Pratt, eds., United States Min- eral Resources, U.S. Geological Survey Professional Paper 820, Washington, D.C., pp. 133-142 (1973). Averitt, P., Coal Resources of the United States, January 1, 1974, U.S. Geo- logical Survey Bulletin 1412, Washington, D.C. (1975). 110 Averitt, P., and L. Lopez, Bibliography and Index of the U.S. Geological Sur- vey Publications Relating to Coal, 1882-1970, U.S. Geological Survey Bulletin 1377, Washington, D.C. (1972). Bailey, R.G., Ecoregions of the United States (Map), U.S. Department of Agri- culture, Forest Service, Ogden, Utah, scale 1:7;500,000 (1976). Bailey, R.G., Description of the Ecoregions of the United States, U.S. De- partment of Agriculture, Forest Service, Miscellaneous Publication No. 1391, Ogden, Utah (1980). Balding, G.O., Water Availability, Quality, and Use in Alaska, U.S. Geolog- ical Survey Open-File Report 76-513, Anchorage (1976). Barnes, F.F., A Review of the Geology and Coal Resources of the Bering River Coal Field, Alaska, U.S. Geological Survey Circular 146, Washington, D.C. (1951). Barnes, F.F., Coal Fields of the United States - Sheet 2, Alaska (Map), U.S. Geological Survey, Washington, D.C., scale 1:5,000,000 (1961). Barnes, F.F., Variation in Rank of Tertiary Coals in the Cook Inlet Basin, Alaska, U.S. Geological Survey Professional Paper 450-C, Washington, D.C. (1962). Barnes, F.F., Geology and Coal Resources of the Beluga-Yentna Region, Alaska, U.S. Geological Survey Bulletin 1202-C, Washington, D.C. (1966). Barnes, F.F., Coal Resources of Alaska, U.S. Geological Survey Bulletin 1242-B, Washington, D.C. (1967a). Barnes, F.F., Coal Resources of the Cape Lisburne-Colville River Region, Ss Alaska, U.S. Geological Survey Bulletin 1242-E, Washington, D.C. (1967b). Barnes, F.F., and E.H. Cobb, Geology and Coal Resources of the Homer District, Kenai Coal Field, Alaska, U.S. Geological Survey Bulletin 1058-F, Washington, D.C. (1959). Barnes, F.F., and T.G. Payne, The Wishbone Hill District, Matanuska Coal Field, Alaska, U.S. Geological Survey Bulletin 1016, Washington, D.C. (1956). Barnes, F.F., and D. Sokol, Geology and Coal Resources of the Little Susitna District, Matanuska Coal Field, Alaska, U.S. Geological Survey Bulletin 1058-D, Washington, D.C. (1959). Beikman, H.M., compiler, Geologic Map of Alaska, U.S. Geological Survey, scale 1:2,500,000, 2 sheets, Washington, D.C. (1980). Bottge, R., Changing Economics of Alaskan Coals, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal ‘75, Proceedings of Conference, Univer- sity of Alaska, MIRL Report 37, Fairbanks, pp. 139-149 (Oct. 15-17, 1975). Bottge, R. ed., Alaska's Mineral Potential, 1978, U.S. Bureau of Mines, Alaska Field Operations Center, Juneau (1978). 111 Brelsford, H.G., Governmental Monitoring of Major Projects, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Foundation, University of Colorado, Boulder, pp. 10-1 through 10-21 (1978). Brody, S., Alaska's Surface Coal Mining Program, Alaska State Legislature, House Reseach Agency Report 80-8, Juneau (198la). Brody, S., Issues Analyst, Alaska State Legislature, House Research Agency, Juneau, Personal Communication (1981b). Brody, S.E., and A.H. DeVries, Coal Policy Paper: Leasing and Taxation, Alaska State Legislature, House Research Agency Report 80-4, Juneau (1981). Brooks, A.H., Mineral Resources of Alaska, U.S. Geological Survey Bulletin 394, Washington, D.C. (1909). Brooks, A.H., and G.C. Martin, The Coal Resources of Alaska, in the Coal Re- sources of the World, 12th International Geological Congress, Canada, 2:541- 552 (1913). Brown, J., and N.A. Grave, Physical and Thermal Disturbance and Protection of Permafrost, U.S. Army Corps of Engineers, Cold Regions Research and Engineer- ing Laboratory Special Report 79-5, Hanover, N.H. (1979). Buch, J.W., T.A. Hendricks, and A.L. Toenges, Coal, in Mineral Position of the United States, U.S. 80th Congress, lst Session, Senate Committee on Public Lands, Inventory of Natural Resources, Washington, D.C. pp. 231-235 (1947). Callahan, J.E., Geology and Coal Resources of T. 6 S., R. 51 W., (resurveyed) Umiat Principal Meridian, in the Cape Beaufort Coal Field, Northwest Alaska, U.S. Geological Survey Open-File Report 496, Anchorage (1971). Callahan, J.E., Coal Investigations in Western Arctic Alaska, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal '75, Proceedings of Conference, University of Alaska, MIRL Report 37, Fairbanks, pp. 48-53 (Oct. 15-17, 1975). Carlson, R.F., and J. Butler, eds., Alaska Water Resources Research Needs for the 70's, Institute of Water Resources Report No. IWR-39, University of Alaska, Fairbanks (1973). Carpenter, D., Alaska Coal Interests Taiwan Delegation, Anchorage Times, Anchorage (Aug. 18, 198la). Carpenter, D., Coal Cheered at State "Pep Rally", Anchorage Times, Anchorage (March 27, 1981b). Clark, P.R., Transportation Economics of Coal Resources of Northern Slope Coal Fields, Alaska, University of Alaska, Mineral Industry Research Labora- tory Report No. 31, Fairbanks (1973). Coal Industry News, Diamond Shamrock in Alaska, Coal Industry News, 5(20), New York (Aug. 24, 1981). 112 Coal Outlook, Marubeni Studying Alaskan Coal, Pasha Publications, Washington, D.C. (Dec. 10, 1979). Cobb, E.H., Reports of the Alaska Division of Geological and Geophysical Sur- veys_ and Predecessor Agencies, 1913-1973, Indexed by Quadrangle, U.S. Geolog- ical Survey Open-File Report 74-209, Anchorage (1974a). Cobb, E.H., Selected U.S. Bureau of Mines Reports on Alaska Indexed by Quad- rangle, U.S. Geological Survey Open-File Report 74-52, Anchorage (1974b). Cobb, E.H., Synopsis of the Mineral Resources and Geolo of Alaska, U.S. Geological Survey Bulletin 1307, Washington, D.C. (1974c). Conwell, C.N., Alaskan Coals May Prove to be a Big Plus in Future Export Pic- ture, Mining Engineering, 24(10 282-84 (Oct. 19735, Conwell, C.N., Reclaiming Mined Land in Alaska, Transactions of the Society of Mining Engineers, 260:81-84 (March 1976). Conwell, C.N., Land Reclamation is an Integral Part of the Only Operating Coal Mine in Alaska, The Coal Miner, 2(3):21-27 (Sept. 1977). Cooper, H.M., et al., Analyses of Mine, Tipple, and Delivered Samples, in Analyses of Alaska Coals, U.S. Bureau of Mines Technical Paper 682, Washing- ton, D.C., pp. 19-69 (1946). Denton, S.W., Chief Engineer, Usibelli Coal Mine, Inc., Usibelli, Alaska, Personal Communication (1982). DeStefano, P.R., and T.H. Foss, Federal Lands in Alaska, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Foundation, University of Colorado, Boulder, pp. 6-1 through 6-48 (1978). DeVries, A.H., Coal Policy Paper: Markets for Alaskan Coal, Alaska State Legislature, House Research Agency Report 80-3, Juneau (1981). Eakins, R.H., Director, Alaska Division of Economic Enterprise and Head, Coal Task Force, Juneau, Personal Communication, (1981). Edblom, G., Markets for Bulk Coal, available from Resource Development Coun- cil for Alaska, Inc., Box 516, Anchorage (1981). ; Engleman, P., et al., Transportation and Development of Alaska Natural Re- sources, Federal-State Land Use Planning Commission for Alaska, Study No. 32, Anchorage (1978). Epps, A.C., Alaska's National Interest Lands: A Congressional Dilemma, Jour- nal of Soil and Water Conservation, 33(1):6-10 (1978). Federal-State Land Use Planning Commission for Alaska, Resources of Alaska: A Regional Summary, Anchorage (1974). Federal-State Land Use Planning Commission for Alaska, The D-2 Book, Lands of National Interest in Alaska, Anchorage (May 1977). 113 Ferrians, O.J. Jr., R. Kachadoorian, and G.W. Greene, Permafrost and Related Engineering Problems in Alaska, U.S. Geological Survey Professional Paper 678, Washington, D.C. (1969). Feulner, A.J., J.M. Childers, and V.W. Norman, Water Resources of Alaska, U.S. Geological Survey Open-File Report 1971, Anchorage (1971). Gates, G.O., Coal Fields of Alaska, in Analyses of Alaska Coals, U.S. Bureau of Mines Technical Paper 682, Washington, D.C., pp. 1-9 (1946). Geer, M., State-by-State Reports on Coal West of the Mississippi, Including Canada ~ Alaska, Coal Age, 78(4):75-77 (Mid-April, 1973). Grantz, A., G. Plafker, and R. Kachadoorian, Alaska's Good Friday Earthquake, March 27, 1964: A Preliminary Geologic Evaluation, U.S. Geological Survey Circular 491, Washington, D.C. (1964). Harlan, R.L., Hydrogeological Considerations in Northern Pipeline Development, Inland Waters Directorate, Environment Canada, Task Force on Northern Oil De- velopment Report No. 74-26, Ottawa (1974). Harrison, S.D., Alaska Population Overview, Alaska Department of Labor, Juneau (1979). Hartman, C.W., and P.R. Johnson, Environmental Atlas of Alaska (2nd edition), Institute of Water Resources, University of Alaska, Fairbanks (1978). Hopkins, D.M., Lignite Deposits near Broad Pass Station, Alaska, U.S. Geolog- ical Survey Bulletin 963-E, Washington, D.C., pp. 187-191 (1951). Hopkins, D.M., et al., Permafrost and Ground Water in Alaska, U.S. Geological Survey Professional Paper 264-F, Washington, D.C. (1955). Hunt, C.B., Physiography of the United States, W.H. Freeman and Company, San Francisco (1967). Iseri, K.T., and W.B. Langbein, Large Rivers of the United States, U.S. Geo- logical Survey Circular 686, Washington, D.C. (1974). Johnson, L., Revegetation and Selected Terrain Disturbances Along the Trans— Alaska Pipeline, 1975-1978, U.S. Army Corps of Engineers, Cold Regions Re- search and Engineering Laboratory Report 81-12, Hanover, N.H. (1981). Johnson, L., W. Quinn, and J. Brown, Revegetation and Erosion Control Obser- vations Along the Trans-Alaska Pipeline, 1975 Summer Construction Season, U.S. Army Corps of Engineers, Cold Regions Research and Engineering Labora- tory Special Report 77-8, Hanover, N.H. (1977). Johnson, L., and K. VanCleve, Revegetation in Arctic and Subarctic North America, A Literature Review, U.S. Army Corps of Engineers, Cold Regions Research and Engineering Laboratory Report 76-15, Hanover, N.H. (1976). Kaiser Engineers, Inc., Technical and Economic Feasibility Surface Mining Coal Deposits North Slope of Alaska, U.S. Bureau of Mines Open File Report 153-77, Juneau (1977). 114 Kazickas, J., World Coal Trade: Policies, Politics, and Procedures, available from Resource Development Council for Alaska, Inc., Box 516, Anchorage (1981). Kennedy, G., Director, Alaska Legislative Information Office, Washington, D.C., Personal Communication, (1981). Kirshenbaum, N.W., Manager of Mineral Projects, Placer Amex Inc., San Fran- cisco, Personal Communication (1981). Knowlton, J., Coal Purchaser Intends to Open Anchorage Office, Anchorage Times, Anchorage, p. A-l (December 26, 198da)- Knowlton, J., Japanese to Develop State Coal, Anchorage Times, Anchorage, p. A-1 (Nov. 25, 1980b). Knowlton, J., Japan Proposes $200 Million Coal Study, Anchorage Times, Anchorage (Jan. 9, 1981). Lachenbruch, A.H., Permafrost, in Fairbridge, R.W., ed., The Encyclopedia of Geomorphology, Reinhold Book Corp., New York (1968). Lawson, D.E., et al., Tundra Disturbance and Recovery Following the 1949 Ex- ploratory Drilling, Fish Creek, Northern Alaska, U.S. Army Corps of Engineers, Cold Regions Research and Engineering Laboratory Report 78-28, Hanover, N.H. (1978). Lowenfels, J., Laws and Regulations Directly Affecting Coal Exploration and Surface Coal Mining and Reclamation in Alaska, available from Resource Devel- opment Council for Alaska, Inc., Box 516, Anchorage (1981). Maloney, R.P., Reconnaissance of the Beluga River Coalfield, Alaska, U.S. Bureau of Mines Report of Investigations 5430, Washington, D.C. (1958). Martin, G.C., and J.E. Callahan, Preliminary Report on the Coal Resources of the National Petroleum Reserve in Alaska, U.S. Geological Survey Open-File Report 78-1033, Anchorage (1978). McConkey, W., et al., Alaska Regional Energy Resources Planning Project - Phase 1, Vol. I: Findings and Analysis, Alaska Division of Energy and Power Development, Anchorage tio7 7a). McConkey, W., et al., Alaska Regional Energy Resources Planning Project - Phase 1, Vol. II: Inventory of Oil, Gas, Coal, Hydroelectric, and Uranium Re- sources, Alaska Division of Energy and Power Development, Anchorage (1977b). McKendrick, J.D., Mine Reclamation in Portions of West Germany, Union of Soviet Socialist Republics and Alaska, Agroborealis, 12(1): 11-14 (Jan. 1980). McKendrick, J.D., and W.W. Mitchell, Fertilizing and Seeding Oil-Damaged Arc- tic Tundra to Effect Vegetation Recovery Prudhoe Bay, Alaska, Arctic, 3103): 296-304 (Sept. 1978). McFarland, C.E., The Alaskan Coal Industry; A Status Report, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Founda- tion, University of Colorado, Boulder, pp. 9-1 through 9-28 (1978). 115 McGee, D.L., Coal Reserve Study, Chuitna-Beluga-Capps Area, Alaska, Alaska Division of Geological and Geophysical Surveys Open-File Report 30, Anchorage (1972). McGee, D.L., Gasification Prospects and Application in Cook Inlet, Alaska, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal '75, Proceedings of Conference, University of Alaska, MIRL Report 37, Fairbanks, pp. 227-238 (Oct. 15-17, 1975). McGee, D.L., and K.S. Emmel, Alaska Coal Resources, Alaska Division of Geo- logical and Geophysical Surveys, Unnumbered Internal Report, Anchorage (1979). McGee, D.L., and K.M. O'Connor, Cook Inlet Basin Subsurface Coal Reserve Study, Alaska Division of Geological and Geophysical Surveys Open-File Report 74, Anchorage (1975a). McGee, D.L., and K.M. O'Connor, Mineral Resources of Alaska and the Impact of Federal Land Policies on Their Availability - Coal, Alaska Division of Geo- logical and Geophysical Surveys Open-File Report 51, Anchorage (1975b). Meacham, T.E., The State of Alaska as Landowner, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Foundation, Univer- sity of Colorado, Boulder, pp. 3-1 through 3-15 (1978). Mitchell, W.W., Development of Plant Materials for Revegetation in Alaska, in Kenny, S.T., ed., Proceedings of High Altitude Revegetation Workshop No. 3, Colorado Water Resources Research Institute Information Series No. 28, Color- ado State University, Fort Collins, pp. 101-115 (1978). Mitchell, W.W., Three Varieties of Native Alaskan Grasses for Revegetation Purposes, University of Alaska, Agricultural Experiment Station Circular No. 32, Palmer (1979). Mitchell, W.W., and J.D. McKendrick, Responses of Arctic, Boreal, and Alpine Biotypes in Reciprocal Transplants, in Brown, J., ed., Ecological Investiga- tions of the Tundra Biome in the Prudhoe Bay Region, Alaska, Biological Papers of the University of Alaska, Special Report No. 2, Fairbanks, pp. 92-111 (1975). Murai, S., Coal-Fired Power Generation in Japan - Present Status and Outlook, presented at Alaska Coal Workshop, Juneau (Sept. 23, 1977). Naske, C.M., and D.M. Triplehorn, The Federal Government and Alaska's Coal, The Northern Engineer, 12(3):11-23 (1980). National Research Council, Surface Coal Mining in Alaska: An Investigation of the Surface Mining Control and Reclamation Act of 1977 in Relation to Alaskan Conditions, prepared by Committee on Alaskan Coal Mining and Reclamation, National Academy Press, Washington, D.C. (1980). National Research Council, Surface Mining: Soil, Coal, and Society, prepared by Committee on Soil as a Resource in Relation to Surface Mining for Coal, National Academy Press, Washington, D.C. (1981). 116 Nelson, K., ed., Alaska Petroleum and Industrial Directory, Petroleum Infor- mation Corp., Anchorage (1978). Norman, J.K., and S.W. Silver, Alaska's D-2 Lands, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Foundation, Univer- sity of Colorado, Boulder, pp. 5-1 through 5-71 (1978). Olsen, M., et al., Beluga Coal Field Development: Social Effects and Manage- ment Alternatives, Pacific Northwest Laboratory Report PNL-RAP-29, Richland, Wash. (1979). Palazzo, A.J., S.D. Rindge, and D.A. Gaskin, Revegetation at Two Construction Sites in New Hampshire and Alaska, U.S. Army Corps of Engineers, Cold Regions Research and Engineering Laboratory Report 80-3, Hanover, N.H. (1980). Patsch, B.J.G., Exploration and Development of the Beluga Coal Field, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal '75, Proceedings of Con- ference, University of Alaska, MIRL Report 37, Fairbanks, pp. 72-83 (Oct. 15-17, 1975). Patsch, B.J.G., Beluga Field Potential - The Resource, presented at Coal Workshop, Alaska Coal and the Pacific, Juneau (Sept. 22-23, 1977). Péwé, T.L., Quaternary Geology of Alaska, U.S. Geological Survey Professional Paper 835, Washington, D.C. C1375). Polar Research Board, Opportunities for Permafrost-Related Research Associ- ated With the Trans-Alaska Pipeline System, prepared by Committee on Perma- frost, National Academy of Sciences, Washington, D.C. (1975). Polar Research Board, An Evaluation of Arctic Programs Supported by the Na- tional Science Foundation, prepared by Committee to Evaluate NSF Arctic Pro- grams, National Academy of Sciences, Washington, D.C. (1977). Polar Research Board, Polar Research Board Annual Report 1980 and Future Plans, National Academy Press, Washington, D.C. (1981). Rao, P.D., Characterization of Alaska's Coals, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal '75, Proceedings of Conference, University of Alaska, MIRL Report 37, Fairbanks, pp. 33-47 (Oct. 15-17, 1975). Rao, P.D., Petrographic, Mineralogical and Chemical Characterization of Cer- tain Arctic Alaskan Coals from the Cape Beaufort Region, University of Alaska, School of Mineral Industry Report 44, Fairbanks (1980). Rao, P.D., Mineral Industry Research Laboratory, University of Alaska, Fair- banks, Personal Communication, (1981). Rao, P.D., and E.H. Beistline, Coal Research Needs, available from Resource Development Council for Alaska, Inc., Box 516, Anchorage (1981). Rao, P.D., and E.N. Wolff, Characterization and Evaluation of Washability of Alaskan Coals, Final Technical Report for Phase I, U.S. Department of Energy Report DOE/ET/13350-Tl, Pittsburgh Mining Technology Center, Pittsburgh, Penn. (1979). 117 Rao, P.D., and E.N. Wolff, Characterization and Evaluation of Washability of Alaskan Coals, Final Technical Report for Phase I, U.S. Department of Energy Report DOE/ET/13350-T2, Pittsburgh Mining Technology Center, Pittsburgh, Penn. (1980). Resource Development Council for Alaska, Inc., Alaska Coal Marketing Confer- ence Proceedings, Box 516, Anchorage (1981). Ritter, M.K., Alaska Our Last Great First Chance, Natural Resources Defense Council Newsletter, 8(1):1-14 (March/April 1979). Ross, I.S., Anatomy of a Superport: A Look at Alaska's Needs, Proceedings of Alaska Coal Marketing Conference, Resource Development Council for Alaska, Inc., Anchorage, pp. 11-27 (Jan. 23, 1981). Rutledge, G., D. Lane, and G. Edblom, Alaska Regional Energy Resources Plan- ning Project - Phase 2, Vol. I: Beluga Coal District Analysis, Alaska Divi- sion of Energy and Power Development, Anchorage (1980). Sanders, R.B., Coal Resources of Alaska, in Rao, P.D., and E.N. Wolff (eds.), anon foe te. * . Focus on Alaska's Coal '75, Proceedings of Conference, University of Alaska, MIRL Report 37, Fairbanks, pp. 21-32 (Oct. 15-17, 1975a). Sanders, R.B., Geology and Coal Resources of the Bering River Field, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal '75, Proceedings of Con- ference, University of Alaska, MIRL Report 37, Fairbanks, pp. 54-58 (Oct. 15-17, 1975b). Sanford, R.S., and H.C. Pierce, Explorations of Coal Deposits of the Point Barrow and Wainwright Areas, Northern Alaska, U.S. Bureau of Mines Report of Investigations 3934, Washington, D.C. (1946). Scott, J.J., Research and Development Priorities Surface Mining Reclamation in Alaska, in Scott, J.J., Executive Summary Research and Development Prior- ities Surface Mining Reclamation, Vol. II, Department of Mining, Petroleum and Geological Engineering, University of Missouri, Rolla, pp. 60-75 (1976). Selkregg, L.L., Alaska Regional Profiles, Prepared for Joint Federal-State Land Use Planning Commission for Alaska and Office of the Governor, Arctic Environmental Information and Data Center, Anchorage, Six Volumes (1974-76). Shively, J.T., Alaska Native Corporations and Native Lands, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Foundation, University of Colorado, Boulder, pp. 4-1 through 4-33 (1978). Slaughter, C.W., T.G. Freeman, and G.L. Audsley, Cooperation in Water Re- sources Programs: Alaska's Example, Water Resources Bulletin, 10(4):802-812 (1974). Smith, D.W., and T. Tilsworth, eds., Environmental Standards for Northern Regions, Proceedings of Symposium, University of Alaska, Institute of Water Resources Report No. 62, Fairbanks (1975). 118 Stearns, S.R., Selected Aspects of Geology and Physiography of the Cold Re- gions, U.S. Army Material Command, Cold Regions Research and Engineering Lab- oratory Monograph I-Al, Hanover, N.H. (1965). Stearns, S.R., Permafrost (Perennially Frozen Ground), U.S. Army Material Command, Cold Regions Research and Engineering Laboratory Monograph I-A2, Hanover, N.H. (1966). Stenmark, R.J., An Introduction to Alaska, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Foundation, University of Colorado, Boulder, pp. 2-1 through 2-77 (1978). Sturdevant, D., Environmental Constraints to Coal Development, available from Resource Development Council for Alaska, Inc., Box 516, Anchorage (1981). Swanson, V.E., et al., Collection, Chemical Analysis, and Evaluation of Coal Samples in 1975, U.S. Geological Survey Open-File Report 76-468, Washington, D.C. (1976). Swinzow, G.K., Certain Aspects of Engineering Geology in Permafrost, Engineer- ing Geology, 3:177-215 (1969). Tailleur, I.L., Low-volatile Bituminous Coal of Mississippian Age on the Lis- burne Peninsula, Northwestern Alaska, in Geological Survey Research 1965, U.S. Geological Survey Professional Paper 525-B, pp. 34-38 (1965). Tailleur, I.L., and W.P. Brosge, Coal Resources of Northern Alaska May Be Na- tion's Largest, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal 75, Proceedings of Conference, University of Alaska MIRL Report 37, Fair- banks, pp. 219-226 (Oct. 15-17, 1975). Thornbury, W.D., Regional Geomorphology of the United States, John Wiley and Sons, Inc., New York (1965). Tillinghast, J.K., Developments in Alaska Environmental Law, in Proceedings of Mineral Development Institute, Rocky Mountain Mineral Law Foundation, University of Colorado, Boulder, pp. 11-1 through 11-29 (1978). Toenges, A.L., and T.R. Jolley, Investigation of Coal Deposits for Local Use in_the Arctic Regions of Alaska and Proposed Mine Development, U.S. Bureau of Mines Report of Investigations 4150, Washington, D.C. (1947). Toenges, A.L., and T.R. Jolley, Investigation of Coal Deposits in South Cen- tral Alaska and the Kenai Peninsula, U.S. Bureau of Mines Report of Investi- gations 4520, Washington, D.C. (1949). U.S. Bureau of the Census, Statistical Abstract of the United States: 1980, U.S. Department of Commerce, Washington, D.C. (1980). U.S. Geological Survey, Coal Resource Classification System of the U.S. Bureau of Mines and U.S. Geological Survey, U.S. Geological Survey Bulletin 1450-B, Washington, D.C. (1976). -S. Geological Survey, Water Resources Data for Alaska, Water Year 1977, U.S U.S. Geological Survey Water-Data Report AK-77-l1, Washington, D.C. (1978a). 119 U.S. Geological Survey, Water Resources Data for Illinois, Water Year 1977, U.S. Geological Survey Water-Data Report IL-77-l, Washington, D.C. (1978b). Usibelli, J., Mining Constraints and Operations at Usibelli Coal Mine, in Rao, P.D., and E.N. Wolff, eds., Focus on Alaska's Coal '75, Proceedings of Conference, University of Alaska, MIRL Report 37, Fairbanks, pp. 93-101 (Oct. 15-17, 1975). Usibelli Coal Mine, Inc., booklet produced by Usibelli Coal Mine, Inc., Usi- belli, Alaska (no date). Wahrhaftig, C., Coal Deposits of the Costello Creek Basin, Alaska, U.S. Geo- logical Survey Open-File Report, Anchorage (1944). Wahrhaftig, C., Geology and Coal Deposits of the Western Part of the Nenana Coal Field, Alaska, U.S. Geological Survey Bulletin 963-E, Washington, D.C., pp. 169-186 (1951) Wahrhaftig, C., Physiographic Divisions of Alaska, U.S. Geological Survey Professional Paper 482, Washington, D.C. (1965). Wahrhaftig, C., Coal Reserves of the Healy Creek and Lignite Creek Coal Ba- sins, Nenana Coal Field, Alaska, U.S. Geological Survey Open-File Report, Anchorage (1973). Wahrhaftig, C., et al., The Coal-Bearing Group in the Nenana Coal Field, Alaska, U.S. Geological Survey Bulletin 1274-D, Washington, D.C. (1969). Wahrhaftig, C., and J.H. Birman, Stripping-Coal Deposits on Lower Lignite Creek, Nenana Coal Field, Alaska, U.S. Geological Survey Circular 310, Wash- ington, D.C. (1954). Wahrhaftig, C., and C.A. Hickcox, Geology and Coal Deposits, Jarvis Creek Coal Field, Alaska, U.S. Geological Survey Bulletin 989-G, Washington, D.C. (1955). Wahrhaftig, C., C.A. Hickcox, and J. Freedman, Coal Deposits on Healy and Lignite Creeks, Nenana Coal Field, Alaska, U.S. Geological Survey Bulletin 963-E, Washington, D.C., pp. 141-165 (1951). Walker, F.K., Bibliography and Index of U.S. Geological Survey Publications Relating to Coal, 1971-1975, U.S. Geological Survey Circular 742, Washington, D.C. (1976). Warfield, R.S., Bituminous Coal Deposits of the Matanuska Coal Field, Alaska: Central and Western Parts, Wishbone District, U.S. Bureau of Mines Report of Investigations 5950, Washington, D.C. (1962). Warfield, R.S., Investigation of a Subbituminous Coal Deposit Suitable for Opencut Mining, Beluga River Coalfield, Alaska, U.S. Bureau of Mines Report of Investigations 6238, Washington, D.C. (1963). Warfield, R.S., Resume of Information on Alaskan Bituminous Coals with Par- ticular Emphasis on Coking Characteristics, U.S. Bureau of Mines Open-File Report 11-67, Juneau (1967). 120 Washburn, A.L., Focus on Polar Research, Science, 209(4457):643-652 (8 Aug. 1980). Williams, J.R., Ground Water in the Permafrost Regions of Alaska, U.S. Geo- logical Survey Professional Paper 696, Washington, D.C. (1970). Williams, T.K., State Taxation, in Proceedings of Alaska Mineral Development Institute, Rocky Mountain Mineral Law Foundation, University of Colorado, Boulder, pp. 12-1 through 12-19 (1978). Wilson-Smith, N.G., Far East Markets, available from Resource Development Council for Alaska, Inc., Box 516, Anchorage (1981). Wolff, E.N., et al., Optimum Transportation Systems to Serve the Mineral In- dustry North of the Yukon Basin in Alaska, University of Alaska, Mineral In- dustry Research Laboratory Report No. 29, Fairbanks (1973). Zemansky, G.M., T. Tilsworth, and D.J. Cook, Alaska Mining and Water Quality, Institute of Water Resources Report No. IWR-74, University of Alaska, Fair- banks (1976). Zenone, C., and G.S. Anderson, Summary Appraisals of the Nation's Ground- Water Resources - Alaska, U.S. Geological Survey Professional Paper 813-P, Washington, D.C. (1978). 121 APPENDIX COAL RESOURCE CLASSIFICATION SYSTEM OF THE U.S. BUREAU OF MINES AND U.S. GEOLOGICAL SURVEY Coal Resource Classification System of the U.S. Bureau of Mines and U.S. Geological Survey MINERAL RESOURCE CLASSIFICATION SYSTEMS OF THE U.S. BUREAU OF MINES AND U.S. GEOLOGICAL SURVEY GEOLOGICAL SURVEY BULLETIN 1450-B A report published jointly by the U.S. Bureau of Mines and U.S. Geological Survey Definitions of coal resource classification terms used by the U.S. Burcau of Mines and U.S. Geological Survey — eee UNITED STATES GOVERNMENT PRINTING OFFICE, WASHINGTON : 1976 UNITED STATES DEPARTMENT OF THE INTERIOR THOMAS S. KLEPPE, Secretary GEOLOGICAL SURVEY Vv. E. McKelvey, Director Library of Congress catalog-card No. 76-600026 For sale by the Superintendent of Documents, U.S. Government Printing Office Washington, D.C. 20402 Stock Number 024-001-02814-0 €21 FOREWORD In order to use mineral resource terms with precision and com- mon understanding and to compare resource data effectively, a joint U.S. Bureau of Mines and U.S. Geological Survey work group developed a standardized, definitive, broadly applicable classification system to derive uniform, coordinated resource esti- mates. The principles of the system are given in Chapter A of this series (Bulletin 1450-A). This chapter presents the classification system for coal resources. Future chapters will present classifica- tion terms for other specific commodities. Thames V, follie, OU. €. Via (<a Thomas V. Falkie Director, Bureau of Mines V.E. McKelvey Director, Geological Survey Ill CONTENTS Foreword Introduction -_- Classification system Glossary of coal classification terms Criteria for coal resource/reserve identification ILLUSTRATION FicureE 1. Classification of coal resources ------------------------- TABLE TABLE 1. Coal resource/reserve criteria --.....-.----------------- Page Page B2 Page 97T SSIFICATION SYSTEMS OF ‘THE YD U.S. GEOLO! . SURVEY MINERAL RESOURCE CI U.S. BUREAU OF MINES » COAL RESOURCE CLASSIFICATION SYSTEM OF THE U.S. BUREAU OF MINES AND U.S. GEOLOGICAL SURVEY INTRODUCTION This method of classification is in conformity with the provi- sions of the Joint Geological Survey-Bureau of Mines Resource Classification Agreement of November 21, 1973, covering all min- eral resources and will be used in future resource/reserve studies on coal conducted by agencies of the Department of the Interior. All resource and reserve estimates will be dated. Within this system the term “coal resource” designates the esti- mated quantity of coal in the ground in such form that economic extraction is currently or potentially feasible. The “coal reserve” is that part of the resource for which rank, quality, and quantity have been reasonably determined and which is deemed to be min- able at a profit under existing market conditions. CLASSIFICATION SYSTEM This system employs a concept by which coal beds are classified in terms of their degree of geologic identification and economic and technologie feasibility of recovery. In the following conceptual diagram (fig. 1) showing the relationship of the various factors involved, coal resources are located on the horizontal scale, in- creasingly to the left, according to their degree of geological assurance of existence, and on the vertical scale, increasingly up- ward, according to their degree of economic and technologic feasi- bility of recovery. The following general definitions of coal resource categories are amplified by the criteria for resource identification, which follow the Glossary. The criteria may be revised to reflect changing con- ditions without affecting the definitions. GLOSSARY OF COAL CLASSIFICATION TERMS Resources.—Concentrations of coal in such forms that economic extraction is currently or may become feasible. Bl B2 MINERAI, RESOURCE CLASSIFICATION SYSTEMS COAL RESOURCES As of January 1, i (billion short tons) IDENTIFIED UNDISCOVERED Demonstrated Interred HYPOTHETICAL SPECULATIVE (in known districts) iva Monsured Indicated districts) Reserves! ECONOMIC @ () [increasing degree of > 4 _t n 4 << — increasing degree of geologse assurance 4 * | suBEcONOMIC "Recovery factor = Total Remaining Resources Cumulative Product ion *Includes ___ billion tons of Reserve Base Total Original Resources coal that are not currently minable. (see recovery factor). Includes __ billion tons of Reserve Base that are not currently minable (see recovery factor). Average production, (5 yr. period) million short tons Production, (most recent year) —_ million short tons FIGuRE 1.—Classification of coal resources. Identified Resources.—Specific bodies of coal whose location, rank, quality, and quantity are known from geologic evidence supported by engineering measurements. Undiscovered Resources.—Unspecified bodies of coal surmised to exist on the basis of broad geologic knowledge and theory. Reserve Base.—That portion of the Identified Coal Resource from which Reserves are calculated. Reserve.—That portion of the Identified Coal Resource that can be economically mined at the time of determination. The re- serve is derived by applying a Recovery Factor to that com- ponent of the Identified Coal Resource designated as the Reserve Base. Recovery Factor.—The percentage of total tons of coal estimated to be recoverable from a given area in relation to the total tonnage estimated to be in the Reserve Base in the ground. economug feasibuity. scl COAL RESOURCE CLASSIFICATION SYSTEM B3 Identified Subeconomic Resources.—The part of coal resources that occurs in Demonstrated and Inferred Resources that is not now minable economically. Hypothetical Resources.—Undiscovered Coal Resources in beds that may reasonably be expected to exist in known mining districts under known geologic conditions. In general, Hypo- thetical Resources are in broad areas of coal fields where points of observation are absent and evidence is from distant outcrops, drill holes, or wells. Exploration that confirms their existence and reveals quantity and quality will permit their reclassification as a Reserve or Identified Subeconomic Resource. Speculative Resources.— Undiscovered coal in beds that may occur either in known types of deposits in a favorable geologic set- ting where no discoveries have been made, or in deposits that remain to be recognized. Exploration that confirms their existence and reveals quantity and quality will permit their reclassification as Reserves or Identified Subeconomic Resources. The following definitions are applicable to both the Reserve and Identified Subeconomic Resource components. Measured.—Coal for which estimates of the rank, quality, and quantity have been computed, within a margin of error of less than 20 percent, from sample analyses and measurements from closely spaced and geologically well-known sample sites. Indicated.—Coal for which estimates of the rank, quality, and quantity have been computed partly from sample analyses and measurements and partly from reasonable geologic pro- jections. Demonstrated.—A collective term for the sum of coal in both Measured and Indicated Resources and Reserves. Inferred.—Coal in unexplored extensions of Demonstrated Re- sources for which estimates of the quality and size are based on geologic evidence and projection. Rank.—The classification of coals relative to other coals, according to their degree of metamorphism, or progressive alteration, in the natural series from lignite to anthracite (Classification of Coal by Rank, 1938, American Society for Testing Materi- als, ASTM Designation D-388-88, p. 77-84). Quality or Grade.—Refers to individual measurements such as heat value, fixed carbon, moisture, ash, sulfur, phosphorus, major, minor, and trace elements, coking properties, petro- logic properties, and particular organic constituents. The indi- B4 MINERAL RESOURCE CLASSIFICATION SYSTEMS vidual quality. elements may be aggregated in various ways to classify coal for such special purposes as metallurgical, gas, petrochemical, and blending usages. CRITERIA FOR COAL RESOURCE/RESERVE IDENTIFICATION Estimates of the different classes of coal resources and reserves are arbitrarily based upon three criteria: (1) thickness, rank, and quality of the coal bed, (2) depth of the coal bed, and (3) the proximity of the coal resource data upon which the estimate was based. Depth and thickness are criteria because they control eco- nomic and technologic feasibility of recovery. The criteria for each class are described below and summarized in table 1 and will be used in preparing all Department of the Interior coal resource/ reserve estimates from January 1, 1975, until further revised. These criteria apply only to those coal bodies that are or will be economically extractable by underground, surface, and/or in situ methods. Coal thinner than 14 inches (35 cm) (anthracite and bituminous) and 30 inches (75 cm) (subbituminous and lignite) and all coal deeper than 6,000 feet (1,800 m) is excluded. These thinner and deeper coals will be considered at a later date. Coal containing more than 33 percent ash is excluded from resource and reserve estimates. Identified Resources.—Include beds of bituminous coal and an- thracite 14 inches (35 cm) or more thick and beds of sub- bituminous coal and lignite 30 inches (75 cm) or more thick that occur at depths to 6,000 feet (1,800 m), and whose exist- ence and quantity have been delineated within specified de- grees of geologic assurance as measured, indicated, or infer- red. Include also thinner and/or deeper beds that presently are being mined or for which there is evidence that they could be mined commercially. Undiscovered Resources.—Include beds of bituminous coal and anthracite 14 inches (35 cm) or more thick and beds of sub- bituminous coal and lignite 30 inches (75 cm) or more thick that are presumed to occur in unmapped and unexplored areas to depths of 6,000 feet (1,800 m). Remaining Resources.—Includes the sum of the Identified and Undiscovered Resources as of the date of the estimate. Cumulative Production.—Includes the sum of all production prior to the date of the estimate. 921 COAL RESOURCE CLASSIFICATION SYSTEM B5 TABLE 1.—Coal resource/reserve criteria Depth, Feet (Metres) TWEentimetteas Total Resources and Undiscovered Re- sources. Anthracite and =6,000 (1,800) =14 (35) bituminous coal. Subbituminous =6,000 (1,800) =30 (75) coal and lignite. Identified Resources." Anthracite and =6,000 (1,800) =14 (35) bituminous coal. Subbituminous =6,000 (1,800) ==30 (75) coal and lignite. Reserve Base.’ Anthracite and =1,000 (300) =28 (70) bituminous coal. Subbituminous =1,000 (300) =60 (150) coal. Lignit® scscoeececo= =120 (40) =60 (150) Reserves. Criteria same as Reserve Base but with Recovery Factor applied. Subeconomic Resources? Anthracite and 0-1,000 (300) 14 (35)-28 (70) bituminous 1,000 (300)-6,000 (1,800) =14 (35) coal. Subbituminous 0-1,000 (300) 30 (75)-60 (150) coal, 1.000 (300)-6,000 (1,800) =30 (75) ignite =-a=seeeanees 0-120 (40) 30 (75)-60 (150) 120 (40)-6,000 (1,800) =30 (75) indicated, and Inferred according to the 1Identified Resources are classified as Measured, di if geologic assurance as described in the text. “Ethe Reserve Base includes ‘some beds. that are thinner and/or deeper than the general criteria permit, but that presently are being mined or are judged to be minable com- mercially at this time. 2Also includes currently nonrecoverable part of Reserve Base. Total Original Resources.—Includes the sum of the Remaining Resources and Cumulative Production as of the date of the estimate. Reserve Base.—Includes beds of bituminous coal and anthracite 28 inches (70 cm) or more thick and beds of subbituminous- coal 60 inches (150 cm) or more thick that occur at depths to 1,000 feet (300 m). Includes also thinner and/or deeper beds that presently are being mined or for which there is evidence that they could be mined commercially at this time. Includes B6 MINERAL RESOURCE CLASSIFICATION SYSTEMS beds of lignite 60 inches (150 cm) or more thick which can be surface mined—generally those that occur at depths no greater than 120 feet (40 m). Reserve.—Includes that portion of the Reserve Base that can be mined at the time of classification (See Recovery Factor). Recovery Factor.—On a national basis the estimated Recovery Factor for the total Reserve Base is 50 percent. More precise recovery factors can be computed by determining the total coal in place and the total coal recoverable in any specific locale. Subeconomic Resources.—Include all Identified Resources that do not fall into the Reserve category. Include in this category beds of bituminous coal and anthracite 14 inches (35 cm) to 28 inches (70 em) thick and beds of subbituminous coal 30 inches (75 cm) to 60 inches (150 cm) thick that occur at depths to 1,000 feet (300 m). Include also beds of bituminous coal and anthracite 14 inches (35 cm) or more thick and beds of subbituminous coal 30 inches (75 cm) or more thick that occur at depths between 1,000 (300 m) and 6,000 feet (1,800 m). Include lignite beds 30 inches (75 cm) or more thick that cannot be surface mined—generally those that occur at depths greater than 120 feet (40 m), and lignite beds 30 inches (75 cm) to 60 inches (150 cm) thick that can be sur- face mined. Include the currently nonrecoverable portion of the Reserve Base. The following criteria are applicable to both the Reserve and Subeconomic Resources components: Measured.—Resources are computed from dimensions revealed in outcrops, trenches, mine workings, and drill holes. The points of observation and measurement are so closely spaced and the thickness and extent of coals are so well defined that the ton- nage is judged to be accurate within 20 percent of true ton- nage. Although the spacing of the points of observation neces- sary to demonstrate continuity of the coal differs from region to region according to the character of the coal beds, the points of observation are no greater than 14 mile (0.8 km) apart. Measured coal is projected to extend as a !,-mile (0.4- km) wide belt from the outcrop or points of observation or measurement. Indicated.—Resources are computed partly from specified meas- urements and partly from projection of visible data for a reasonable distance on the basis of geologic evidence. The L71 COAL RESOURCE CLASSIFICATION SYSTEM BT points of observation are !» (0.8 km) to 14 miles (2.4 km) apart. Indicated coal is projected to extend as a '/4-mile (0.8- km) wide belt that lies more than '4 mile (0.4 km) from the outcrop or points of observation or measurement. Inferred.—Quantitative estimates are based largely on broad knowledge of the geologic character of the bed or region and where few measurements of bed thickness are available. The estimates are based primarily on an assumed continuation from Demonstrated coal for which there is geologic evidence. The points of observation are 1!2 (2.4 km) to 6 miles (9.6 km) apart. Inferred coal is projected to extend as a 2'4-mile (3.6-km) wide belt that lies more than *, mile (1.2 km) from the outcrop or points of observation or measurement. Hypothetical Resources.—Quantitative estimates are based on a broad knowledge of the geologic character of a coal bed or region. Measurements of coal thickness are more than 6 miles (9.6 km) apart. The assumption of continuity of a coal bed is supported only by geologic evidence. Speculative Resources.—Quantitative estimates are based on geo- logic assumptions that undiscovered coal may occur in known types of deposits or in favorable geologic settings. 821 129 Alaskan Coal: Resources and Developmental Constraints. D.E. Edgar, L.J. Onesti, and G.M. Kaszynski. Argonne National Laboratory Report ANL/LRP-18. This report is one in a series being produced by the Land Reclamation Pro- gram. This program is a joint effort of the Energy and Environmental Sys- tems Division and the Division of Environmental Impact Studies at Argonne National Laboratory, Argonne, Illinois 60439. Sponsor: U.S. Department of Energy, Office of Energy Research. Program Summary: The Land Reclamation Program is addressing the need for coordinated ap- plied and basic research into the physical and ecological problems of land reclamation, and is advancing the development of cost-effective techniques for reclaiming land mined for coal. This program is con- ducting integrated research and development projects focused on near- and long-term reclamation problems in all major U.S. coal resource regions, and is evaluating and disseminating the results of related studies conducted at other research institutions. These activities in- volve close cooperation with the mining industry. Regional and site- specific reclamation problems are being addressed at research demonstra- tion sites throughout the country, and through laboratory and greenhouse experiments. Program Director: Ralph P. Carter Deputy Director, Biological Sciences: Ray R. Hinchman Deputy Director, Physical Sciences: Stanley D. Zellmer Principal Investigator for the projects discussed in this report: Dorland E. Edgar Publication Date: August 1982 Key Words: Alaskan Coal Arctic Ecosystems Coal Mining Natural Resource Management Reclamation Subarctic Ecosystems Surface Coal Mining 130 Distribution for ANL/LRP-18 Internal: R. P. Carter R. R. Hinchman S. D. Zellmer E. J. Croke A. B. Krisciunas ANL Contract Copy D. E. Edgar K. S. Macal ANL Libraries W. J. Hallett C. A. Malefyt (115) ANL Patent Department W. Harrison J. J. Roberts TIS Files (6) R. E. Rowland External: U.S. Department of Energy Technical Information Center, for distribution per UC-88 (143) Manager, U.S. Department of Energy Chicago Operations Office President, Argonne Universities Association Energy and Environmental Systems Division Review Committee: E. E. Angino, University of Kansas H. J. Barnett, Washington University R. L. Clodius, National Assn. of State Universities and Land Grant Colleges, Washington, D.C. . Egan, Environmental Research and Technology, Inc., Lexington, Mass. - Mullins, Indiana University Poundstone, Pittsburgh - Stukel, University of Illinois, Urbana J. J. Wortman, North Carolina State University Arctic Environmental Information and Data Center, Anchorage D. V. Bennett, Alaska Council on Science and Technology, Juneau J. W. Blumer, Mobil Oil Corporation, Denver S. Brody, House Research Agency, Juneau J. Brown, U.S. Army Cold Regions Research and Engineering Laboratory, Hanover, N.H. K. Brown, Division of Minerals and Energy Management, Anchorage R. F. Carlson, University of Alaska, Fairbanks R. Dawes, U.S. Office of Surface Mining, Denver S. W. Denton, Usibelli Coal Mine, Inc., Usibelli, Alaska A. J. Dixon, U.S. Senate, Washington, D.C. R. H. Eakins, Alaska Dept. of Commerce and Economic Development, Juneau P. Emery, U.S. Geological Survey/Water Resources Div., Anchorage J. E. Ferrell, U.S. Bureau of Mines, Washington, D.C. R. Franklin, U.S. Dept. of Energy, Washington, D.C. D. Gerkin, Mining and Reclamation Council, Washington, D.C. J. H. Gibbons, Office of Technology Assessment, Washington, D.C. S. A. Glen, Arch Minerals Corp., St. Louis, Mo. A. F. Grandt, Peabody Coal Co., St. Louis, Mo. J. C. Greene, Committee on Science and Technology, U.S. House of Representa- tives, Washington, D.C. J. S. Hammond, Governor of Alaska, Juneau D. H. Hunter, Utah International, Inc., San Francisco D. E. Kash, University of Oklahoma J. W. Katz, Alaska Dept. of Natural Resources, Juneau G. Kennedy, Alaska Legislative Information Office, Washington, D.C. usaaw UuUZap RErPOS RP AOMrMHAOKUESs avo R. 131 Kirshenbaum, Placer Amex Inc., San Francisco Mattson, U.S. Bureau of Mines, Juneau . McKendrick, University of Alaska, Palmer Mertz, Assistant Attorney General, Fairbanks Miller, U.S. Geological Survey, dacharase CA OUE . Mims, Illinois Dept. of Energy and Natural Resources, Chicago W. Mueller, Alaska Dept. of Environmental Conservation, Juneau . Mullan, National Coal Association, Washington, D.C. H. Murkowski, U.S. Senate, Washington, D.C. J. Onesti, Indiana University, Bloomington (5) . S. Osburn, U.S. Dept. of Energy, Washington, D.C. H. Percy, U.S. Senate, Washington, D.C. O. Perry, U.S. Office of Surface Mining, Pittsburgh Pernela, Alaska Dept. of Commerce and Economic Development, Anchorage T. Plass, U.S. Forest Service, Princeton, W.Va. Price, U.S. House of Representatives, Washington, D.C. Ralston, AMAX Coal Co., Indianapolis D. Rao, University of Alaska, Fairbanks R. Rastorfer, Chicago State University D. Reger, Division of Geological and Geophysical Surveys, Anchorage Resource Development Council for Alaska, Inc., Anchorage R. R. aa COmMmMAmMEnmAHaasy B. Sanders, Diamond Shamrock Corp., Anchorage Schaff, Division of Geological and Geophysical Surveys, Anchorage Schmidt, U.S. Bureau of Land Management, Anchorage (2) W. Sedwick, Division of Land and Water Management, Anchorage Simon, U.S. House of Representatives, Washington, D.C. H. Simonds, Freedman, Levy, Kroll & Simonds, Washington, D.C. Sims, Alaska Dept. of Commerce and Economic Development, Fairbanks Stevens, U.S. Senate, Washington, D.C. Stoops, Division of Research and Development, Anchorage Thomas, U.S. Office of Surface Mining, Mills, Wyo. . White, U.S. Geological Survey, Menlo Park, Calif. Wobber, U.S. Dept. of Energy, Washington, D.C. Yocum, AMAX Coal Co., Indianapolis Yould, Alaska Power Authority, Anchorage Young, U.S. House of Representatives, Washington, D.C.