.... .229... iii“. . V , . V .1 ... . ‘. ..K. 5‘ z, .. ..x... .1. .....v. ... . 5 HR: ..npi . sir“)-.. £4.30... fix uni»... ..."...nuynha . ... in ...: : ...... 1111...].(1 1:. .. ......l 2.1x ......s ....V 1!. . w". 2 km. .7 \A .4...) _ .2133... .v. 7;...» \u vll‘. . A. V... e «um:- ‘ . 3 x? Waugh 1. \l 1.3mm... {1. . . ... a . , . . .....i . .. ... .Iilv' \2 h ..‘Y‘X lt‘ I. 1'“ ......32»... ......l... . . . . 4 “a. V . . . h. .9 . . . ...:in :‘ . V ... . ...V .. . “wens? . 4.9 . .... .95.... , .... . r‘ .5; vv . r n.. .3 . : .... a! . I . 11"....3: “my .. . . rain... .— . . ‘ . 51: . . 1 . . ...: {1v , . 1 13.... h .2 II. “‘ 11V 3. . ....1..........v. $1371 4191.1: ... ...-... J . . .... 9 xv.. :1 ;;Al 5 9531* £591“! 93'1- .. ... . ‘36:... ct .2.) (....B‘q‘ a. $9 ...-.4: .. and ., J; . £9,433}. gi. ,3. « V I..- 2... .. 1 -..... .1; E . I7 . 1 THESIS 3 I!”UNI!H!I”!HINIUIIUIHIHHIIHIUllllfllllllllflfll 301421 6513 This is to certify that the dissertation entitled PEDOLOGY AND DATING OF COLLUVIAL DEPOSITS IN THE BLUE RIDGE MOUNTAINS, NORTH CAROLINA presented by Johan Liebens has been accepted towards fulfillment of the requirements for Ph . D . degree in Geography W/W Major pro Date January 22, 1996 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 LiaR‘ARY Michigan State University PLACE N RETURN BOX to move thie checked from your record. TO AVOID FINES return on or bdore dete due. DATE DUE DATE DUE DATE DUE - II I | MSU Ie An Aflimetive Action/Ewe! Opportunity Inetltution W ”3-9.1 PEDOLOGY AND DATING OF COLLUVIAL DEPOSITS IN THE BLUE RIDGE MOUNTAINS, NORTH CAROLINA. By Johan Liebens A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Geography 1 996 ABSTRACT PEDOLOGY AND DATING OF COLLUVIAL DEPOSITS IN THE BLUE RIDGE MOUNTAINS, NORTH CAROLINA. By Johan Liebens Colluvium and soils in areas of warm, humid climates and metamorphic bedrock have not been studied in detail. The study area for this research, the Little Tennessee River basin in the Southern Blue Ridge of North Carolina, exhibits these characteristics. This research examines debris flow deposits of this basin, and their associated soil, and places their evolution in a temporal framework. The results of this study provide characterization data and help identify any possible temporal synchroneity of past periods of increased landscape stability/instability in the region. The sedimentology and geomorphology of three sequences of multiple colluvial deposits were studied. Pedons on each deposit of each sequence were described and analyzed semi-quantitatively. A relative chronostratigraphy of the deposits was determined with established techniques based on weathering data, soil morphology and iron species ratios. The sediments of the debris flow deposits typically consist of massive, clast supported diamicts. Flow viscosities appear to have been widely variable and may have been related to the texture of the source material. Shear zones in the debris flows rendered the sediments heterogeneous and are reflected in the irregular depth trends of many pedological properties. Depth to maximum clay accumulation in the soils systematically occurs at greater depths at higher elevations. Clay mineral assemblages vary little within sites; between site differences may be linked to the age of the deposits (timespan of subaerial weathering). Consistent results from qualitative and statistical evaluation of the relative-age indicators of the deposits point to the effectiveness of the relative-age proxies used in this study. These proxies indicate that colluvial deposits at higher elevations are usually older. The apparent isochronous occurrence of debris flow activity at the three sites suggests that external factors such as climate variations influence the timing of debris flow activity in the region. Differences in age between sites suggest that intrinsic factors co-determine the instability of the basin. ACKNOWLEDGMENTS My foremost appreciation is due to my advisor, Dr. R]. Schaetzl, for his invaluable advice, support, encouragement, and skillful editing. I am also grateful for the interest and advice of the other members of my committee, Dr. D.G. Brown, Dr. G. Larson, Dr. D.P. Lusch, and Dr. M.A. Velbel. My sincere thanks go to Dr. F.W. Cambray for his assistance with fabric analysis. Field and laboratory equipment of the Department of Geography of Michigan State University was used extensively. Laboratory facilities were also provided by the Department of Crop and Soil Sciences of Michigan State University. This research could not have been carried out without the support of the landowners who allowed me to dig soil pits on their property. My sincere thanks are due to the Fortner, Bradley, J acobstein, Ramsey and Hanson families. Several graduate students of the Department of Geography helped with parts of this research. Dave Nolan was an invaluable help during the surveying of the sites and often put geography in perspective. Linda Barrett was the ideal company, and a good sounding board, during the writing stage of the project. Finally, three people were instrumental in the successful completion of this work and provided all the understanding and mental support one needs in an endeavor like this. My parents always have supported and encouraged my educational aspirations and did not fail to do so now. My wife, Hilde, inspired me by being the perfect example of dedication and perseverance. This research was supported by NSF grant SBR-9405198 made to R. Schaetzl and J. Liebens, and by GSA grant 5404-94 and a grant from Harvard Travelers Club Permanent Fund made to J. Liebens. iv TABLE OF CONTENTS List of tables ...................................................................................... xiii List of figures ..................................................................................... xiv Part I: Introduction .................................... . ............................................. 1 1 Introduction ..................................................................................... 2 1.1 Background ......................................................................... 2 1.2 Social and economic impact of debris flows .................................... 3 1.3 Research objectives ................................................................. 5 1.4 Relevance of the study ............................................................. 7 2 Literature review ............................................................................... 8 2.1 Terminology of mass flows .......... . ............................................. 8 2.2 Localizing factors of debris flows .............................................. 11 2.2.1 Precipitation amount .................................................. 1 1 2.2.2 Precipitation intensity ................................................. 12 2.2.3 Local factors ........................................................... 14 2.2.3.1 Precipitation distribution .................................. 14 2.2.3.2 Slope ......................................................... 14 2.2.3.3 Hillslope form and aspect ................................. 15 2.2.3.4 Geology ..................................................... 15 2.2.3.5 Regolith thickness and texture ............................ 16 2.2.3.6 Vegetation ................................................... 17 3 Studyarea .................................................................................... 19 3.1 Introduction ....................................................................... 19 3.2 Study areadescription ............................................................ 22 3.2.1 Climate .................................................................. 22 3.2.2 Geology ................................................................ 23 3.2.3 Geomorphology ....................................................... 27 3.2.4 Soils ..................................................................... 32 3.2.5 Colluvium .............................................................. 33 3.2.5.1 Geomorphology and sedimentology ..................... 33 3.2.5.2 Elevations ................................................... 35 3.2.5.3 Coarse fragments ........................................... 35 3.2.5.4 Ages ......................................................... 36 3.2.6 Vegetation .............................................................. 38 3.3 Site selection ...................................................................... 39 3.3.1 Reconnaissance survey ............................................... 39 3.3.1.1 Methods ..................................................... 39 3.3.1.1.1 General ........................................... 39 3.3.1.1.2 Specific methods ................................ 39 3.3.1.2 Results ....................................................... 41 3.3.2 Selection criteria ....................................................... 42 3.3.3 Site description and location ......................................... 44 3.3.3.1 Alarkasite ................................................... 44 3.3.3.2 Hidden Valley site .......................................... 48 vi 3.3.3.3 Bradley site ................................................. 52 Part II: Characterization of sediments and soils ............................................... 56 4 Introduction ................................................................................... 57 5 Sediments ..................................................................................... 59 5.1 Literature review .................................................................. 59 5.1.1 Sorting .................................................................. 59 5.1.2 Bed geometry .......................................................... 60 5.1.3 Stratification ............................................................ 60 5. 1 .4 Grading ................................................................. 60 5.1.5 Conglomerate framework ............................................ 61 5.1.6 Fabric shape ............................................................ 62 5.1.7 Fabric strength ......................................................... 64 5.2 Methods ............................................................................ 65 5 .2. 1 Data collection ......................................................... 65 5.2.2 Data analysis ........................................................... 68 5.3 Results ............................................................................. 70 5 .3 . 1 Sediment description .................................................. 70 5.3.1.1 General ...................................................... 70 5.3.1.2 Alarkasite ................................................... 72 5.3.1.3 Hidden Valley site .......................................... 82 5.3.1.4 Bradley site ................................................. 89 5.3.2 Fabric ................................................................... 92 5.3.2.1 A-axis ........................................................ 92 5.3.2.2 A-bplane .................................................... 97 vii 5.4 Discussion ......................................................................... 97 5.4.1 General ................................................................. 97 5.4.2 Alarkasite .............................................................. 99 5.4.3 Hidden Valley site .................................................... 100 5.4.4 Bradley site ............................................................ 101 5.4.5 Fabric .................................................................. 102 5.5 Conclusions ...................................................................... 105 6 Soils .......................................................................................... 108 6.1 Literature review ................................................................. 108 6.1.1 Morphology ........................................................... 108 6.1.2 Physical and chemical properties ................................... 109 6.1.2.1 Clay content ................................................ 109 6.1.2.2 Bulk density ............................................... 109 6.1.2.3 Basesaturation ............................................ 110 6.1.2.4 Free iron content .......................................... 110 6.1.3 Clay mineralogy and weathering ................................... 111 6.1.3.1 Biotite ....................................................... 111 6.1.3.2 Chlorite ..................................................... 111 6.1.3.3 VermiculiteandHIV ...................................... 111 6.1.3.4 Kaolinite .................................................... 112 6.1.3.5 Gibbsite .................................................... 113 6.1.3.6 Ironoxides ................................................. 114 6.1.4 Genesis ................................................................ 115 6.1.5 Classification .......................................................... 117 6.2 Methods ........................................................................... 118 6.2.1 Datacollection ........................................................ 118 viii 6.2.2 Data analysis .......................................................... 1 18 6.3 Results ............................................................................ 120 6.3.1 Soil morphology and classification ................................. 120 6.3.1.1 Alarkasite .................................................. 120 6.3.1.2 Hidden Valley site ......................................... 121 6.3.1.3 Bradley site ................................................ 122 6.3.2 Physical and chemical properties ................................... 122 6.3.2.1 Alarkasite .................................................. 122 6.3.2.2 Hidden Valley site ......................................... 126 6.3.2.3 Bradley site ................................................ 129 6.3.3 Clay mineralogy ...................................................... 133 6.3.3.1 General ..................................................... 133 6.3.3.2 Alarkasite .................................................. 135 6.3.3.3 Hidden Valley site ......................................... 135 6.3.3.4 Bradley site ................................................ 135 6.4 Discussion ........................................................................ 136 6.4.1 Saprolite ............................................................... 136 6.4.2 Lithologic discontinuities ............................................ 137 6.4.3 Clay content ........................................................... 141 6.4.4 Clay mineralogy ...................................................... 142 6.5 Conclusions ...................................................................... 144 Part III: Relative age dating .................................................................... 146 7 Introduction .................................................................................. 147 8 Clast weathering ............................................................................ 150 ix 8. 1 Literature review ................................................................. 150 8.1.1 General ................................................................ 150 8.1.2 Weathering rind thickness .......................................... 150 8.1.3 Percent weathered clasts ............................................. 152 8.2 Methods ........................................................................... 152 8.3 Results and discussion .......................................................... 157 8.4 Conclusions ...................................................................... 156 9 Soil development ............................................................................ 158 9.1 Literature review ................................................................. 158 9.2 Methods ........................................................................... 160 9.2.1 Selection of relative-age indicators ................................. 160 9.2.2 Data collection and analysis ......................................... 161 9.3 Results and discussion .......................................................... 162 9.3.1 Rubification ........................................................... 162 9.3.2 Clay content ........................................................... 164 9.4 Conclusions ...................................................................... 166 10 Iron species .................................................................................. 168 10.1 Literature review ................................................................. 168 10.1.1 General ............................................................... 168 10.1.2 Fed and Feo .......................................................... 169 10.1.3 Fed/Pet, (Fed-Feo)/Fet and Feo/Fed .............................. 170 10.2 Methods ........................................................................... 171 10.2.1 Selection of relative-age indicators ................................ 171 10.2.2 Data collection and analysis ........................................ 173 10.3 Results and discussion .......................................................... 174 10.4 Conclusions ...................................................................... 177 11 Discriminant analyses ...................................................................... 178 11.1 Literature review ................................................................. 178 11.2 Methods ........................................................................... 179 11.2.1 Construction of relative-age groups ............................... 179 11.2.2 Data analysis ......................................................... 181 11.3 Results and discussion ................ i .......................................... 182 1 1.4 Estimation of absolute ages ..................................................... 184 l 1.5 Conclusions ...................................................................... 185 12 Conclusions ................................................................................. 186 Part IV: Conclusions ............................................................................ 188 Part V: Appendixes Appendix 1: Generalized description of selected pedons of reconnaissance survey ..... 193 Appendix 2: Longitudinal profiles, Alarka and Bradley sites .............................. 198 Appendix 3: Particle size fractionation data .................................................. 201 Appendix 4: Sorting and bulk density data ................................................... 210 Appendix 5: Stereograms for clast a-axis ..................................................... 214 Appendix 6: Stereograms for clast a-b plane ................................................. 221 Appendix 7: Clay mineralogy data ............................................................. 225 Appendix 8: Iron species data .................................................................. 229 Part VI: List of references List of references ................................................................................ 233 Table 1. Table 2. Table 3. Table 4. Table 5. Table 6. Table 7. Table 8. Table 9. Table 10. Table 11. Table 12. Table 13. Table 14. Table 15. Table 16. Table 17. Table 18. Table 19. Table 20. LIST OF TABLES Clast fabric data for debris flow-like deposits ................................... 65 Geomorphic position of soil pits .................................................. 66 Soil horizons sampled for clast characterization and fabric .................... 67 Pedon descriptions .................................................................. 73 Clast sizes, all sites ................................................................. 83 Clast sphericities, all sites......... ................................................. 84 Clast lithologies, all sites ........................................................... 85 Orientation of clast a-axes, total population ...................................... 93 Orientation of clast a-axes, subpopulations ...................................... 95 Orientation of clast a-b planes ..................................................... 98 Clay mineralogy by pedon, all sites ............................................. 136 Clast weathering data, all sites ................................................... 155 Cluster analysis of Clast weathering data ........................................ 157 Field description of selected pedons of reconnaissance survey .............. 195 Particle size analyses of selected pedons of reconnaissance survey ......... 197 Particle size data, all sites ......................................................... 203 Clayfree particle size data, all sites ............................................... 208 Sorting and bulk density data, all sites .......................................... 212 Clay mineralogy data, all sites .................................................... 227 Iron species data, all sites ......................................................... 231 xiii Figure 1. Figure 2. Figure 3. Figure 4. Figure 5. Figure 6. Figure 7. Figure 8. Figure 9. Figure 10. Figure 11. Figure 12. Figure 13. Figure 14. Figure 15. Figure 16. Figure 17. Figure 18. Figure 19. Figure 20. Figure 21. LIST OF FIGURES Physiographic provinces of the Southern Appalachians ...................... 20 Location of study area ............................................................. 21 Precipitation frequency diagram for Coweeta Hydrological Laboratory .... 24 Regional geology of southwest North Carolina ................................ 25 Relief map of the Blue Ridge and Piedmont physiographic provinces ...... 28 Location of study sites ............................................................ 45 Basin shapes, all sites ............................................................. 46 Contour map of Alarka site ....................................................... 47 Cross section, Alarka site ......................................................... 49 Contour map of Hidden Valley site .............................................. 50 Cross section, Hidden Valley site ................................................ 51 Cross section, Bradley site ............ i ........................................... 54 Contour map ofBradley site ....................................................... 55 Fabric shape and strength as a function of S 1/S2 and 82/S3 ................. 71 Percent coarse fragments as a function of depth, all sites ..................... 79 Percent fine gravel as a function of depth, Alarka site ......................... 80 Percent clay free sand as a function of depth, Alarka site ..................... 81 Roundness of clasts, all sites ..................................................... 86 Interpretive log column of lithofacies units, lowest deposit at Hidden Valley site ................................................................. 87 Interpretive log column of lithofacies 'units, highest deposit at Hidden Valley site ................................................................. 88 Interpretive log column of lithofacies units, Bradley site ..................... 90 xiv Figure 22. Figure 23. Figure 24. Figure 25. Figure 26. Figure 27. Figure 28. Figure 29. Figure 30. Figure 31. Figure 32. Figure 33. Figure 34. Figure 35. Figure 36. Figure 37. Figure 38. Figure 39. Figure 40. Figure 41. Figure 42. Figure 43. Figure 44. Figure 45. Figure 46. Figure 47. Figure 48. Logarithmic ratio plot of eigenvalues, selected sites ........................... 96 Logarithmic ratio plots, this study and other depositional environments... 106 Clay mass as a function of depth, Alarka site ................................. 123 Bulk density as a function of depth, Alarka site ............................... 124 Clayfree silt as a function of depth, selected pedons ......................... 125 Free iron content as a function of depth, Alarka site .......................... 127 Clay mass as a function of depth, Hidden Valley site ........................ 128 Bulk density as a function of depth, Hidden Valley site ..................... 130 Free iron content as a function of depth, Hidden Valley site ................ 131 Clay mass as a function of depth, Bradley site ................................ 132 Free iron content as a function of depth, Bradley site ........................ 134 Depth of maximum clay content, all sites ...................................... 141 Rubification indices, all sites .................................................... 163 Clay content, all sites ................... . .......................................... 165 Fed/Fet as a function of depth, selected pedons .............................. 175 Mean profile weighted Fed/Fet, all sites ....................................... 176 Discriminant scores, all sites .................................................... 183 Longitudinal profile, Alarka site ................................................ 200 Longitudinal profile, Bradley site ............................................... 201 Stereograms of long axis of clasts, lower deposits at Alarka site ........... 216 Stereograms of long axis of clasts, upper deposit at Alarka site ............ 217 Stereogram of long axis of clasts, Hidden Valley site ........................ 218 Stereograms of long axis of clasts, Bradley site .............................. 219 Stereograms of long axis of clasts, subpopl .................................. 220 Stereograms of long axis of clasts, subpop2 .................................. 221 Stereograms of poles of a—b planes of clasts, lower deposits at Alarka site 223 Stereograms of poles of a-b planes of clasts, upper deposit at Alarka site 224 XV Figure 49. Stereograms of poles of a-b plane of clasts, Bradley site .................... 225 xvi PART I: INTRODUCTION 1. INTRODUCTION 1.1 Background The geomorphology of the Southern Blue Ridge has yet to be studied in detail. Pen'glacial phenomena of the region have been studied to some extent (e. g., Michalek, 1968; Raymond, 1976; Mills, 1981; Daniels et al., 1987) but work on alluvial and colluvial deposits from zones that were not affected by periglacial conditions is still in its infancy. One prevalent geomorphological aspect of these other morpho-climatic zones are debris flow deposits, the most common and spectacular form of mass wasting in the region (Scott, 1972; Mills, 1987). According to Scott (1972) and Schneider (1973) these types of mass movements occur more frequently in the Southern Blue Ridge of North Carolina than in any other part of the Appalachians. Mills (1987) provided a number of reasons for their high incidence in the Southern Blue Ridge: i) high permeability of regolith derived from crystalline rocks relative to that of regolith derived from shales on the Appalachian plateau, ii) more intense rainfall, and iii) high susceptibility of crystalline rocks to saprolitization. In spite of their frequent occurrence, many questions about the origin and distinctive characteristics of debris flows in the Southern Blue Ridge, and humid-temperate regions in general, remain to be answered. The sedimentology and geomorphology of debris flows have been studied extensively on alluvial fans in arid and glacial environments, but little attention has been given to debris flow deposits in humid-temperate areas (Kochel and Johnson, 1984). This is also true of studies of soil development on debris flow deposits (Kesel and Spicer, 1985; Norfleet and Smith, 1989). Particularly lacking are studies of rates of soil formation on this type of colluvial deposit in humid- temperate areas (Kesel and Spicer, 1985). Another question that needs further examination deals with the origin(s) of debris flows. It has been suggested that debris flows in the Southern Blue Ridge are associated 3 with climate change, in particular post-glacial climate change (Mills, 1981; 1982a; 1983; Kochel and Johnson, 1984). However, it also has been recognized that intrinsic factors, unrelated to climate, may control the formation of debris flows (Mills, 1983; Wells and Harvey, 1987). Mills ( 1982a) voiced that an obvious approach to verify the influence of climatic controls on the formation of colluvial deposits would be to see if the deposits date to times of glacial climates. In addition to providing information on the effects of intrinsic and external controls, absolute ages could also serve to identify periods of increased landscape instability, during which debris flow incidence was higher. Absolute dates on debris flow deposits in the Southern Blue Ridge have only recently become available (Kochel and Johnson, 1984; Kochel, 1990), providing only preliminary data. The primary reason for the scarcity of absolute dates, according to Clark (1979), is that sites where datable, essentially-continuous sedimentary records have been preserved are scarce and widely separated. 1.2 ' onomici ac fdebris flows Mass movements occur in every state of the USA and in virtually every country in the world. More than 20 million mass movement deposits exist in the USA alone (Brabb, 1989). In the USA, annual economic losses from mass movements, subsidence and ground failures exceed those from all other natural hazards combined (National Research Council, 1985). In Japan, more than half of all deaths due to natural hazards are caused by slope failures (Schuster, 1978). Debris flows are among the most destructive mass movements. In Canada they formed 61% of known, damaging mass movements from 1855 to 1983 (Cruden et al., 1989). Thirty percent of the fatalities during heavy rainstorrns in Japan are caused by debris flows and about 90 lives are lost annually from this phenomenon (Takahashi, 1981). 4 Because of their viscosity, debris flows can travel large distances and often affect floodplains, which are usually centers of economic activity. Estimates of the annual cost of damage to private dwellings by mass movements range from $44 million (Brabb, 1984) to $400 million (Krohn and Slosson, 1976). Indirect economic losses to forestry, fisheries, agriculture and infrastructure are difficult to evaluate but may be larger than direct costs. Chassie and Goughnour (1976) believe that $100 million is a conservative estimate of the annual cost of mass movement damage to highways and roads in the USA. Total annual losses in the USA have been estimated to be as high as $1 billion (Schuster, 1978; Fleming and Taylor, 1980) to $1.5 billion (Jahns, 1978; Schuster and Fleming, 1986). Annual damage estimates fluctuate considerably because of the infrequent, irregular occurrence of large, catastrophic events. One such event, Hurricane Camille, triggered debris flows in western Virginia in 1969. More than 150 people perished due to this storm, which damaged and destroyed property worth $420 million (Williams and Guy, 1973) In countries other than the USA the death toll from debris flow events has been much higher. The principal reason for this is the lack of resources for adequate awareness campaigns and mitigation in Third World countries. The Andean countries are among the hardest hit. In 1962 and 1970 debris avalanches on Mt. Huascaran in Peru killed 5000 and 18000 people, respectively, and destroyed several towns (Schuster, 1978). In 1985, a volcanic debris flow, or lahar, completely destroyed the town of Armero, Colombia, and killed 20,000 people. In 1987 a landslide in Ecuador resulted in 1000 deaths and ruptured the trans-Ecuadorian oil pipeline, resulting in an approximate $1.5 billion loss. Despite a growing understanding of the factors and processes involved in slope failures, economic and social losses from them continue to increase. This is largely a consequence of land development that expands onto steeply sloping terrain, where mass movements occur frequently. 5 A considerable amount of experience with control of mass movements has been gathered in the last decades. Active and passive mitigation measures have reduced damage by mass movements but require financial efforts. Passive measures such as hazard mapping and warning systems are relatively inexpensive and have some merit. Active intervention usually involves engineering structures, such as retaining walls, dams, channel linings and deposition basins, and are more costly. The most economical and expeditious means of supporting unstable slopes is the reinforcement of potential source areas with rock "rip rap". Near Port Alice, in the Canadian Cordillera, debris flows at the beginning of the 1970's caused $700,000 damage (Cruden et al., 1989). Protective dikes built in the following years cost $250,000. In the same general area, debris flow defense measures worth $1.1 million were put in place after debris flows destroyed $2.5 million in property. In Howe Sound, Canada, debris flow mitigation structures costing $20 million were built after storm-induced debris flows caused $1 million in damage (Cruden et al., 1989). In Washington, avoidance and correction of damage due to landslides that occurred around the middle of the century behind the Grand Coulee dam cost $20 million (Schuster, 1978). 1.3 Objectives The rationale for this research centered on the lack of a thorough understanding of sedirnentological, pedological and age characteristics of debris flow deposits in the Southern Blue Ridge, in spite of the obvious importance of the phenomenon to the landscape evolution of the region. A better comprehension of these characteristics can possibly lead to insights in the timing and evolution of these prevalent colluvial deposits. Because this research is designed to help elucidate the geomorphic and pedogenic history of the Southern Blue Ridge as recorded in the Little Tennessee River basin in North 6 Carolina, the findings should help to identify periods of stability/instability in the region. To achieve this general objective, three specific goals were incorporated in this study. 1. The first specific goal was to locate sites in the Little Tennessee River basin that have a sequence of debris flow deposits. To optimize the applicability of methods used for the following goals, deposits that showed little signs of erosion, had soils with contrasting development and occurred in basins with a uniform lithology, were chosen. This goal was achieved by studying aerial photographs, soil surveys, geologic and topographic maps, and by field reconnaissance. 2. The second part of the study had a dual purpose: to characterize the sedimentologic facies of the debris flow deposits and to elucidate pedogenetic imprinting on the deposits. These goals were achieved by describing sedimentologic attributes, geomorphologic features, and soil properties on the debris flow deposits, and by laboratory analyses of sedimentary fabric, texture, bulk density and clay mineralogy. 3. The third specific goal was to determine the relative ages of the deposits in each sequence. This goal represented an initial step in the identification of periods of increased landscape stability/instability in the Southern Blue Ridge. It was accomplished using data on clast weathering, soil morphology and ratios of iron species. 1,4 Relevance One of the goals of this research is to provide insight into the facies characteristics of debris flow deposits of the Southern Blue Ridge. As mentioned above (section 1.1), few studies on the sedimentology of colluvial deposits are available. This study will, therefore, produce characterization data that were previously not widely accessible. Additionally, sedimentary facies characteristics, and fabric characteristics in particular, are related to flow regimes. A better understanding of these characteristics can, therefore, lead to a more complete knowledge of the processes involved in the movement of debris flows. This may, in turn, lead to better mitigation and abatement measures. This research will also shed light on how the soils associated with the debris flow deposits have evolved in this warm to temperate humid environment. It will provide data on soil-geomorphic relationships and spatial variability of soils on previously poorly documented colluvial deposits. Although the relative-age indicators used here have been tested extensively, they have not often been used in warm areas where annual precipitation exceeds 100 cm (Birkeland, 1990). Their application in the present research will, therefore, be an evaluation of the consistency of the method. The results of this research may help advance the construction of a Blue Ridge and Appalachian morpho-stratigraphic framework. Such an advance will further the understanding of future climatic and other environmental changes in the region (Clark, 1979). It can therefore be anticipated that the results, through their implications for climate change and carbon cycling by pedologic processes in regolith, may ultimately make a contribution to the Global Change effort. 2. LITERATURE REVIEW Wm Mass wasting is a general term for the downslope movement of regolith or rock material under the influence of gravity without the direct aid of water, air or ice. These media are, however, frequently involved in mass wasting by reducing the shear strength of slope materials. The term landslide is sometimes used as a synonym for rapid mass wasting events (Cruden et al., 1989) although it usually is employed for mass wasting with the exclusion of creep and frozen ground phenomena (Hutchinson, 1968; Selby, 1982). In a stricter sense, landslide is used for a group of mass movements wherein shear failure occurs along a specific surface or combination of surfaces (Schuster, 1978; Zéruba, 1982). Many criteria are available for classifying mass wasting events but in practice, the following criteria are used most often for distinguishing between the types of mass movement: (i) type of material, (ii) mechanism of movement, (iii) velocity of movement, (iv) geometry of failure surface and, (v) water content. In most classification systems, flowage is recognized as one of the mechanisms. Vames (1978) defined flow as a: "... movement within displaced mass such that the form taken by moving material or the apparent distribution of velocities and displacements resemble those of viscous fluids. Slip surfaces within moving material are usually not visible or are short lived. The boundary between moving mass and material in place may be a sharp surface of differential movement or a zone of distributed shear. Movement ranges from extremely rapid to extremely slow". Originally, the term debris avalanche was defined by Sharpe (1938) and was used to describe a rapid and relatively wet flowage. According to the definition, a debris avalanche has a long and relatively narrow track, occurs on a steep mountain slope or 9 hillside in a humid climate, and is almost invariably preceded by heavy rains. Vames (1958) used debris avalanche for a flow of relatively dry unconsolidated material and debris flow for a similar but wetter mass movement. In 1975 Vames did not distinguish between wet and dry flows anymore and used debris flow to describe all flows of coarse soil materials. In 1978, Vames used the term debris avalanche again and defined it as a very rapid to extremely rapid debris flow. In the classification system of Vames (1958; 1978), the term mudflow is a subcategory of debris flow with less than 50 percent gravel. Lowe (1979) divided debris flows into non-cohesive types in which dispersive pressure is the dominant sediment support mechanism, and cohesive types, to which he applies the term "mudflow", in which matrix strength and buoyancy serve as the dominant support mechanisms. Pomeroy (1982) defined debris flow as a rapid downward movement of largely saturated material. He distinguished between debris avalanches and debris flows on the basis of velocity, debris avalanches being faster (Pomeroy, 1982; 1985; 1986). The material comprising debris avalanches usually is more heterogeneous than that of debris flows. Debris flows commonly follow narrow preexisting drainage paths and largely consist, in contrast to mudflows, of coarse rock fragments. VanDine (1985) used debris torrent for the kind of debris flow that follows narrow preexisting drainage channels. He defines it as a: "... mass movement that involves water-charged, predominantly coarse grained inorganic and organic material flowing rapidly down a steep, confined preexisting channel". In the 19803 several rheologic sub—classifications of mass flows appeared in the literature (Pierson, 1984; Savage, 1984; Iverson, 1985; Pierson and Costa, 1987). Pierson and Costa (1987) defined mass flow as ..."the continuous, irreversible deformation of a geologic material that occurs in response to applied stress". The applied stress in most geomorphic situations is gravity. Their classification scheme can accommodate the above, more traditional, nomenclature and is based on the rheologic response of a sediment-water 10 mixture. This response is governed by velocity, a surrogate for shear, and sediment concentration. The earlier definitions of debris flow correspond, in this new classification, with a viscous slurry flow of water and fines that behaves as a plastic fluid and moves at speeds of between 10‘8 and 40 m S‘l. Their lower velocity limit (10'8 m 8‘1 or 0.3 m yr'l) is in contradiction with most previous nomenclatures that described debris flows as being rapid. Debris avalanches correspond, in the new classification, with inertial granular flows of water, air and fines that behave as plastic fluids and travel at speeds of 5 to 40 m 3'1. Innes (1983) revised an existing sub-classification of debris flows and divided them in catastrophic debris flows, lahars, hillslope flows and valley confined flows. The author recognized that catastrophic may not be an appropriate term since all types of debris flow are initiated by extreme events. The term "debris slide" has been employed to describe movements similar to debris flows. Although debris slides often develop into flows by liquefaction and internal deformation of the material (Selby, 1982; Clark, 1987; Ellen and Fleming, 1987; Mills et al., 1987), the term debris slide should be reserved for mass movements that are commonly rapid but have a relatively low water content and involve translational or planar movement (Pomeroy, 1984). Debris flow is sometimes confused with earth flow (Keefer and Johnson, 1983). A debris flow differs from an earth flow in several respects. In a debris flow, granular material admixed with water is mobilized into materials that move more rapidly than earth flows. Debris flows are not bounded by discrete shear surfaces, and most movement takes place by distributed internal shear. Debris flow material is coarser-grained than earth flow material (Vames, 1978). In addition, debris flow deposits generally form during a single episode of activity, so that, within a few minutes or hours, a typical debris flow mobilizes, flows through a channel, forms a deposit, and dries. These deposits are rarely remobilized (Keefer and Johnson, 1983). By analogy with the definition of Vames (1958; 1978), the term debris flow will be employed in this study for rapid and wet mass movements of coarse earthy materials in 11 which flowage is the most common form of movement. Some deposits in the study area may have resulted from debris avalanches sensu Vames (1958) or may have been initiated as debris slides. Since distinguishing the different kinds of deposits is difficult unless observations can be made during or immediately after the event, and because this distinction is not necessary for this research, all will be referred to as debris flow deposits. In the study area of this research the individual deposits have a lobate shape in plan and occur, together with other debris flow deposits, in elongated depositional areas. Therefore, the generic term "deposit" will be used throughout this study, rather than the often used term "fan". 2.2 alizin fact rs 2.2.1 Precipitation amount Excessive moisture is generally considered to be the main factor affecting the occurrence of debris flows (Wolman and Gerson, 1978; Innes, 1983; Pomeroy, 1984; Mills, 1987). The majority of authors attribute the moisture that produces debris flows to rainfall, but rapid snow melt (Morton and Campbell, 1974; Owens, 1974) and rapid drainage of lakes (Costa, 1984) have also been suggested as a trigger. High rainfall appears to have been the dominant triggering process in historical Appalachian debris flows (Kochel, 1987). In the southern Appalachians, synoptic weather patterns indicate that most debris flows can be linked to the incursion of warm, moist tropical air masses between May and November (Kochel, 1990). Other rains that produced debris flows in the region have resulted from convective storms along a front separating extratropical air masses from tropical air. The amounts of precipitation associated with flows, however, vary greatly. Hurricane Camille brought 710 mm of rain (in 8 hours) to certain parts of Virginia and 12 produced numerous flows (Williams and Guy, 1973; Kochel and Johnson, 1984), but as little as 100 mm has caused slopes to fail in Tennessee and Vermont (Bogucki, 1976; Ratte and Rhodes, 1977; Kochel, 1990). Neary et al. (1986) recorded a total of 164 mm (in 12 hours) near a flow in North Carolina. Eschner and Patric (1982) concluded that 125 mm of rainfall (in 24 hours) triggers debris flows at most Appalachian sites, in spite of the high porosity of these forested soils. They estimate a general return period of 100 years for this kind of storm. In some parts of the study area of the present research the return period for storms that size is 4 years (see section 3.2). 2.2.2 Precipitation intensity Rainfall intensity, in relation to infiltration rates of the soils and antecedent moisture, is also an important factor controlling the occurrence of debris flows (Pomeroy, 1984; Jacobson, et al. 1989; Kochel, 1990). Neary et al. (1986) postulate that, for the Blue Ridge, intensities of 90 to 100 mm hr] cause debris sliding. This intensity has a return period of less than 200 years in the study area of the present research. Nearly 40% of the maximum one hour rainfall intensities recorded by Neary and Swift (1987) in a debris flow generating storm in western North Carolina were in excess of 76 mm. In the study area of the present research, that intensity has a return period of about 50 years (see section 3.2). It also has been argued that the rainfall intensity required to trigger flows may be very small. In northwest England, Wells and Harvey (1987) found that intensities as low as 50 mm hrl could trigger debris flows while Kadomura et al. (1980) suggest that 9 mm hr'1 is sufficient to produce mudflows on volcanic cones. Campbell (1975) determined that a rainfall intensity as low as 6.4 mm hr'l can be a threshold condition to initiate slope failure in areas of southern California where the total seasonal antecedent rainfall had reached 254 mm. l3 Intensity-duration graphs show that the combined effect of intensity and duration, i. e. total precipitation, is more important than either of the two factors separately (Caine, 1980; Wieczorek, 1987; Kochel, 1990). According to Kochel's data for the Appalachians, a total of 100 mm constitutes a threshold of minimum rainfall necessary to trigger debris flows. Pomeroy (1984) disagreed with this kind of generalization and believes that antecedent moisture is the determinant factor concerning how much rainfall is required to initiate slope failure. Cannon and Ellen's (1985) findings in the San Francisco Bay region corroborate this idea. For storms exceeding a certain total rainfall threshold they found a strong correlation with periods of continuous high-intensity rainfall and the timing of debris flows. In the same region, Wieczorek (1987) noted that 7 to 30 day antecedent rainfall values for debris flow triggering storms were about twice those of storms that did not trigger flows. Storms with less than 280 mm of seasonal antecedent rainfall did not produce debris flows. In the southern Appalachians, on the other hand, mass flows can occur during intense rainfall events irrespective of the preceding moisture conditions (Clark, 1987). Regolith texture is also an important factor affecting the duration and intensity of rainfall required to trigger debris flows (Sidle and Swanston, 1982; Jacobson et al., 1989). Presumably, long-duration/low-intensity rainfall is needed to fill soil moisture reservoirs in fine -grained regolith, whereas short-duration/high-intensity rainfall is required to fill soil moisture reservoirs in coarse regolith where infiltration rates are high. 2.2.3 Local factors Average rainfall amounts and intensities cannot explain why debris flows occur on one slope and the adjacent slope remains stable. Thus, various reasons for local variations in slope failure have been identified. 14 2.2.3.1 Precipitation distribution Large differences in precipitation amounts on adjacent slopes can occur (Williams and Guy, 1973; Mills, 1987). The differences may be due to a highly cellular precipitation pattern or to the orientation (aspect) of the slopes with respect to the predominant direction of the rainfall. 2.2.3.2 Slope Slope angle considerably affects slope stability. Most failing slopes have an angle that is higher than the angle of repose (Scott, 1972; Koch, 1974; Bogucki, 1976; Escher and Patric, 1982; Pomeroy, 1984; Neary et al., 1986). Williams and Guy (1973) found slopes of less than 30° that failed and the authors explain their observation by the unusually high precipitation intensity causing the flows (710 mm in 8 hours). They also note, however, that flows usually originate on the steepest part of the slope and that average slope angle may therefore not be a good proxy for slope instability. After hurricane Camille, flow head scars could be observed on slopes as low as 15% but the average slope was 35% (Schneider, 1973). In northern Pennsylvania, in an area of sedimentary rocks, debris flows occurred on slopes between 10 and 25% (Pomeroy, 1982). In some areas susceptible to debris flow formation, flows were absent on slopes of more than 45%, probably because the regolith is very thin or absent on those very steep slopes (Pomeroy, 1984; Innes, 1983). Neary and Swift (1987), however, observed that the largest debris flows of their study region had initiated on slopes of more than 70%. 2.2.3.3 Hillslope form and aspect Although some have argued that hillslope form and aspect have little effect on the formation of debris flows (Schneider, 1973; Mills, 1987) cross-slope profile has been identified as an important factor. In some areas, the majority of debris flows occurred along previously existing depressions in hillslopes (Hack and Goodlet, 1960; Williams and Guy, 1973; Bogucki, 1976; Pomeroy, 1984; Clark, 1987; Kochel, 1987). Subsurface flow 15 convergence in such hollows made these sites the most probable locations for failure (Dietrich and Dom, 1984; Reneau et al., 1984). Peatross (1986) used 23 geomorphic variables in a discriminant analysis to identify sites that had been stable and unstable during Hurricane Camille. Variables most successful in producing the two groups included slope angle, slope aspect and hillslope form. Most debris flows caused by Hurricane Camille (Williams and Guy, 1973) and a majority of debris flows in northern and western Pennsylvania (Pomeroy, 1980; 1982; 1984; 1986) were found to occur on northward- or eastward-facing slopes. The regolith on these slopes was probably more moist (prior to the debris flow triggering storm) than regolith on other slopes due to a lowered incidence of direct sunshine. In areas with well developed geologic structures there was a stronger relationship between aspect and flow initiation than in structureless areas (Schneider, 1973; Bogucki, 1976). This influence of structure is an alternative for the presumed influence of moisture and lack of sunshine (Kochel, 1987). 2.2.3.4 Geology Structural and lithologic factors influence the location of debris flows. Lithology has an indirect influence because it is partially responsible for the characteristics of the overlying soil. Debris flows are more common on highly weatherable bedrock (Schneider, 1973; Bogucki, 1976; Pomeroy, 1980) and on cyclic sedimentary rocks with incompetent layers (Pomeroy, 1982). Slopes in northwestern Pennsylvania were particularly unstable when the susceptible incompetent layer occurred near the bottom of the slope (Pomeroy, 1982). Structure has a more direct influence on the origin of flows, especially when the dip is steep (Bogucki, 1976). When the dip angle is low, bedrock orientation does not appear to be a significant factor (Schneider, 1973). Most debris flows in the area studied by Pomeroy (1982) were associated with nearly vertical stress relief joints. The joints increase 16 the secondary permeability of the bedrock and together with other structures they influence the localization of debris flows by creating preferential paths for groundwater flow and seepage (Pomeroy, 1984). The influence of bedrock orientation is not equally important throughout the Southern Blue Ridge, but seems more significant in its western portion (Scott, 1972). In the west, the relatively low metamorphic grade of the rocks (see below) has left them more diverse in their susceptibility to weathering. As a consequence, layers weather differentially and offer more potential planes of failure than might a more homogeneous regolith. Schneider (1973) noted that more debris flows occurred on soils with lower stone content than soils with a greater stone content. The author interpreted this to mean that the stone content reflects bedrock characteristics which have an impact on flow localization, rather than the stone content within the soil affecting its potential to fail. The soils with higher stone content, in this case, generally were derived from rocks with higher quartz sandstone content. 2.2.3.5 Regolith thickness and texture A thin soil mantle is generally thought to favor the formation of flows, presumably because it will accelerate saturation of the regolith during heavy rainfall events (Scott, 1972; Escher and Patric, 1982; Neary et al., 1986; Neary and Swift, 1987). Pomeroy (1984) believed, however, that an extremely thin regolith mantle prevents debris flows from forming. In areas where the regolith was thin, slope failure took place at the colluvial- bedrock interface but where the regolith was thick (>5 m) failures happened within the colluvial cover (Pomeroy, 1982; 1984). Working on the Cumberland Plateau in northern Alabama, Pomeroy (1985) concluded that a supply of regolith downslope from the head area is needed for debris avalanches to occur because he observed an avalanche that terminated as soon as it encountered rock ledges. Others (Schneider, 1973; Mills, 1987) 17 have argued that the thickness of the regolith has little effect on the formation of debris flows. The effect of regolith texture is hard to determine because the original composition of the failed mass cannot be determined exactly. It is thought that at least a small amount of clay is needed to sustain the flow (Pierson, 1981) but a high amount of clay aggravates the saturation and loss of cohesion in the regolith (Neary and Swift, 1987). In the San Francisco Bay region regoliths with a broad range of textures are susceptible to debris flow formation but 98% of the flows have more than 8% clay (Ellen and Fleming, 1987). 2.2.3.6 Vegetation Ascertaining the effect of vegetation and land use on debris flow formation is difficult in the Blue Ridge because most steep slopes, where flows occur, are forested (Mills, 1987). Nevertheless, Scott (1972) reported that clear cutting in the Blue Ridge was not a major factor contributing to debris sliding, and Eschner and Patric (1982) reported that flows in forested areas of the eastern United States occurred irrespective of vegetative cover. Schneider (1973) observed fewer debris flows in forested areas than in cleared land but those under forest tended to be deeper. Daniels et al. (1987) stated that the importance of vegetation in protecting s10pes against mass wasting was demonstrated by the numerous reports of slope failure after logging in the southern Appalachians and reference Hursh (1941). Hursh did not mention, however, any influence of logging on mass movements. In the western United States most workers (O'Loughlin, 1974; Greswell et al., 1979; Swanston, 1979; Swanson et al., 1981) agree that road construction and logging can aggravate flow incidence. Also well known are the effects of the destruction of vegetation by fires on the generation of debris flows in southern California In British Columbia, control of timber harvesting methods and reforestation are employed as primary mitigation measures for debris flow hazards (Hungr et al., 1987). The inconsistent results from the 18 Appalachian hardwood forests can be explained by the rapid sprouting of the root system that is left in place after clearcutting (Neary et al., 1986; Escher and Patric, 1982). 19 3 STUDY AREA 3.1 Introduction The Blue Ridge Mountains run as a long band, in the center of the Appalachian system, from Pennsylvania to Georgia. The Blue Ridge is narrow at its northeastern end in Pennsylvania but widens southwestward into Tennessee and North Carolina, where it is a massive highland 120 km in width (Figure 1). The Blue Ridge Physiographic Province has two subdivisions, the Northern Blue Ridge and the Southern Blue Ridge. The Southern Blue Ridge extends from the Roanoke River in Virginia southwestward to Cartersville, Ga (Hack, 1982). The study area initially considered for this research is located in the Southern Blue Ridge. It coincides with the basin of the Little Tennessee River south of its confluence with the Tuckasegee River (Figure 2). The drainage basin of the Little Tennessee River lies almost completely within Macon County, North Carolina. Franklin, the main urban center of the study area and county seat of Macon County, is situated in the center of the basin. It provided a centralized base of operations for this study. Small parts of the basin extend northwards into Swain County, NC, and southward into Rabun County, GA. The Coweeta Hydrologic Laboratory of the US. Forest Service is located in the study area. The Laboratory is an extremely well studied, outdoor, forest ecology and hydrology research site. More than half a century of climatologic, hydrologic and vegetation data have been recorded at Coweeta. The Laboratory lies about 10 km south of Franklin just west of highway 441. Several of the following paragraphs draw heavily on research carried out at Coweeta. In section 3.2 the physical characteristics of the Little Tennessee River basin are placed in a regional context. The section describes the natural environment of the Southern Physiographic provinces of the 76° Southern Appalachians / 78° After Hack (1987) Figure 1 21 Location of study area “35966 , o Kno Ville ‘3 ,6 G o\\<\ @ . “06‘ @ $06 Ashville o 23 .- @ .\~ . ‘\ Franklin ~\ A a) ' C 3 l 2: a -' 8 c \ ‘8 c 2 :0 ° ./ . ‘\ ‘u ' North Carolina south Carohna Georgia 6‘ e o 4 "9,9 Gainesville (9 ' ZZZ: % Study area 010 20 3O 40 km Smoky Mountains National Park Figure 2 22 Blue Ridge while focusing on local features. The following section (3.3) explains how the specific sites for this research were selected, and provides descriptions of those sites. 3.2 tu are escri tion 3.2.1 Climate. The general climate of the Little Tennessee River basin is humid temperate. The average annual precipitation just southeast of the basin is the highest in the eastern United States (>200 cm annually) and keeps soils in protected areas moist year around. In the study area local precipitation and temperature vary as a function of elevation and aspect. Franklin, located at 650 m elevation, receives 130 cm of precipitation annually, 50 % of which falls in April through September. The average annual snowfall is 20 cm. Thunderstorms occur on about 46 days each year. The heaviest l-day rainfall recorded at Franklin was 15 cm. At higher elevations, just north of the study area, intense rainstorms associated with hurricanes have dumped as much as 30 cm in one event (Neary et al., 1986). Rainfall increases to the south of Franklin and/or at higher elevations. The average temperature for the year is 14 ° C; the summer average is 22° C, while in winter it is 4° C (Thomas, 1993). During 5 years in 10 the last spring frost occurs before May 4. The first frost in fall usually occurs earlier than October 11. At the Coweeta Hydrologic Laboratory, 10 km south of Franklin, precipitation ranges from 170 cm at 750 In elevation to 250 cm at elevations around 1500 m (Douglas and Swank, 1975). Annual evapotranspiration at lower elevations approximates 90 cm. The mean annual temperature for that area is 13° C. The climate of the Laboratory is classified, under Koppen's system, as marine humid temperate (Cfb) (Swift et al., 1988). The lower elevations of the Laboratory, and of the main valley of the study area, have a humid subtropical (Cfa) climate. Of the 133 storms that occur in the Laboratory annually, 23 50% exceed 0.5 cm and 6% exceed 5 cm. Storms exceeding 5 cm of rainfall deliver 34% of the total annual precipitation. Most of the storms have low intensities because rates surpass 1 cm hr'1 only 4% of the time (Swift et al., 1988). Seventy percent of the precipitation at the Laboratory falls with intensities of <2 mm hr'l. Nevertheless, storms with intensities that potentially could cause debris flow have relatively short return periods. Intensities as high as 7.5 cm hr'1 have return periods of only 50 years (Figure 3). 3.2.2 Geology The Blue Ridge Mountains coincide with the crest of an anticlinorium. This structural trend leads to the emergence of Middle and Upper Precambrian and Lower Cambrian rocks in the Southern Blue Ridge. The Little Tennessee River basin is crossed by the Hayesville fault which separates the western from the eastern Southern Blue Ridge (Figure 4) (Hatcher, 1988). The western part of the Southern Blue Ridge consists of sequences of metasedimentary rocks and schists. These rocks were derived from and deposited on the Precambrian Grenville basement of eastern North America. They date from the Upper Precambrian (Hadley and Nelson, 1971; Hatcher, 1988). In the Little Tennessee basin they belong to the Great Smoky group and consist of feldspatic metasandstone with interbeds of quartz-mica schist, phyllite and slate. Beds of conglomerate are also present. The Grenville basement is almost completely absent from the eastern Southern Blue Ridge. This region contains an assemblage of Upper Precambrian rocks which are in many ways different from those to the west of the Hayesville fault (Hadley and Nelson, 1971). Most of the mafic and ultramafic rocks of the Southern Blue Ridge are found in this eastern part. Lithologically the region is characterized by biotite and muscovite schist and gneiss. South of Franklin, in the only part of the study area mapped in detail, rocks of the Coweeta Group and Tallulah Falls Formation can be found (Hatcher, 1980). The Coweeta Group 24 Depth (mm) Precipitation frequency diagram 250 24 hr 12 hr 200 ~ 6 hr 150 — 3 hr ' / 100 — / / 60 min 50 fl //// 30 min / / 15 min // //*5 rrin /L// O / 2 5 10 25 50 100 Return period (years) After Swift et al. (1988) Figure 3 25 as: 5:22.. .22 $39 :2 29:: a 9.8.. ucoEommn catnEwooi g momflocm 2::me I mufitmzd I v 059; 9.8.. ocmE St: ecu can: I mxoou b35863 20323 E azoumzoasm ooooo . ,_ momflocm 339m a .842 .80...» ..w . n“ .893 ..\ \«J I I I .... A \x «122‘ renown \\\ X \KGKGKK EV. on ma cm 2 E m o . & /\/\I\/\I\/\/\ «: o..EVo _ n. ..\ w t» I)... x \ )‘sz \ . ‘N‘l $7 . /\/\/\/ r I \r\’\ \. . v s \ .\ \. . . /\/\/\ .. \ o .. \ ..... x x. \ .8032 em 3" L o c . _ \ boonm k no r 920.80 5 «£6.60 ztoz . . . >>m “.0 30.80 _m:o_mom boom» . . .DMomw 26 dates to the Upper Precambrian or Lower Cambrian. It comprises coarse biotite-garnet schist, metasandstone, quartz-feldspar gneiss, quartz-diorite gneiss and interlayers of pelitic schist. The Tallulah Falls Formation, which occupies a significant portion of the eastern Southern Blue Ridge, is of Upper Precambrian age and consists of coarse grained biotite gneiss and schist. The metamorphic grade of the rocks is lower in the western than in the eastern Blue Ridge (Hadley and Nelson, 1971; Hatcher, 1988). There is a rapid increase, in an easterly direction, from a greenschist facies to an upper amphibolite or granulite facies. Hadley and Nelson (1971) and Hatcher (1988) differ in their location of metamorphic zones and facies. According to Hadley and Nelson (1971) the staurolite and kyanite zones of the amphibolite facies are reached west of the Hayesville fault while Hatcher (1988) situates those zones east of the fault. Hadley and Nelson (1971) locate the sillimanite zone of maximum metamorphic grade in a NE - SW trending zone from Asheville through Franklin into Georgia, but Hatcher (1988) postulates that the sillimanite zone only occurs in NE Georgia. Toward the eastern limit of the Blue Ridge, marked by the Blue Ridge Escarpment and the Brevard fault zone, the metamorphic grade drops off strongly. The Hayesville fault is the major fault in the study area. It is a pre- to synmetamorphic thrust fault that was emplaced prior to the thermal peak which affected the rocks of the Southern Blue Ridge (Hatcher, 1988). It was deformed by later folds that imparted a very sinuous outcrop to it. The Brevard fault zone, to the SE of the study area, roughly follows the Blue Ridge escarpment in North Carolina. It is probably the youngest major fault zone in the region, and is believed to be Paleozoic and may have been active up to the Triassic Period. Other kinds of tectonism may have been active in the region as late as the middle Mesozoic (Hack, 1982). The abrupt difference in relief between the Southern Blue Ridge and the Piedmont cannot be the result of lithologic differences between the two regions because the topographic boundary between the two regions does not correspond to the boundary 27 between Piedmont rocks and Blue Ridge rocks (Hack, 1982). Therefore some kind of tectonism must have influenced the spectacular topographic difference. White (1950) thought that the escarpment was essentially a fault scarp. The Brevard fault zone can not have played a major role because the Southern Blue Ridge has high relief. It must owe its present topography to vertical uplift younger than the middle Mesozoic. Hack (1982) therefore believed that it probably belongs to the same erosional and tectonic block as the Valley and Ridge province. According to Battiau-Queney (1989) the Blue Ridge, Valley and Ridge and Piedmont all are allochtonous terranes above a major detachment, and the terranes were thrust westward by horizontal creep of a ductile crust over Grenville basement rocks. The Blue Ridge escarpment reaches its highest relief where discontinuities in the deeper crust interfere with surficial geological features such as the Brevard fault zone. However, as pointed out by Hack (1982) the escarpment cannot be completely the result of surficial tectonics because of their timing. The crustal creep process proposed by Battiau-Queney (1989) could explain the sustained uplift of the Appalachian highlands and the relatively stable position of the Blue Ridge escarpment. Hatcher's ( 1988) observations seem to indicate that he considers the Hayesville fault to be another boundary between allochtonous terranes, i.e. the western and eastern Southern Blue Ridge. 3.2.3 Geomorphology The Southern Blue Ridge is an area of steep relief (Figure 5) that is bounded on its southeast margin by a bold escarpment. The Blue Ridge Escarpment forms the boundary between the Blue Ridge and the much flatter Piedmont. Rivers on the escarpment drain southeastward to the Piedmont and the Coastal Plain. From the Georgia - South Carolina border northward, the crest of the escarpment cuts diagonally across the entire Blue Ridge anticlinorium. The northwest border of the Southern Blue Ridge province follows the outcrop of quartzite beds. 28 Relief map of the Blue Ridge and Piedmont physiographic provinces \~ .3 s‘r‘“ 5W8: v, In \ , ‘ 'Z' .s“\ :- K . ‘,\Z . \ \ \ \ ‘51-.“ .s“\\\\\‘~\ ' Z ..-\<\\‘ I Steep hills and mountains Moderate relief Low relief 1:] Very low relief After Hack (1987) Figure 5 29 . The Southern Blue Ridge province can be divided into several areas on the basis of topography. Although some of these areas correspond to distinctive geologic areas, some do not and rock control is responsible for many landscape features (Hack 1982). The largest part of the Southern Blue Ridge consists of the ranges of high mountains, many of which exceed 1500 m in altitude. Local relief in the highlands generally exceeds 900 m. Some rock control and structural influence on topography is evident. For example, the highest part of the Great Smoky Mountains is underlain by massive conglomerate and sandstone; here, deep trench-like valleys follow weak rock outcrop areas like carbonate belts. In one case differential weathering and erosion has led to the formation of a 150 km long trench that extends from Georgia into North Carolina. The trench culminates in the Nantahala Gorge just west of the study area of this dissertation. The Asheville basin may be partly controlled by the outcrop of weak rocks but also by intersecting shear zones. Along the northwest and southeast borders of the highlands the ranges tend to be roughly parallel to the northeast regional strike of the rocks. Many valleys have distinct linear trends and some parallel the strike of the rocks or the trends of the thrust faults. Other valleys clearly crosscut the structural grain, as do major basins such as the Little Tennessee River (Clark, 1989). Still others are controlled by joint systems at right angles to the dominant structures (Acker and Hatcher, 1970) One aspect of the geomorphology of the southern Blue Ridge that has received much attention are debris flows. They are considered to be the major transporters of unconsolidated material down the flanks of the mountains and into the channels (Scott, 1972). Colluvial fans in general and debris flows in particular have been attributed to periglacial conditions during the last glaciation (King, 1964; Michalek, 1968). This interpretation needs confirmation (Shafer, 1984; Mills, 1987) because it is probable that drier air associated with Arctic airrnasses may have made intense rainstorms somewhat less common during the pleniglacials than at present (Delcourt and Delcourt, 1984) and because 30 episodes of alluviation/colluviation in the Southern Appalachians seem to have been confined to late glacial/interglacial transitions (Delcourt, 1980). Appalachian debris flow deposits usually do not display fan morphology but are elongated and irregular, because they are often constricted between narrow basin interfluves (Kochel and Johnson, 1984). In the Blue Ridge of Virginia, fans that formed down-basin in expanded valley reaches are free from lateral confinement so that true fan forms are found (Kochel and Johnson, 1984). Variations in debris fan shape appear to be controlled by bedrock resistance, bedrock structure and grain size of the debris (Kochel, 1990). Deposits developed in areas of unifomrly resistant rocks tend to be smaller and have the most irregular shape because they form in confined, narrow valleys. Most Blue Ridge debris flow deposits are less than 20 m thick and occur on slopes between 4 and 10%. They have a hummocky lobate topography even if the overall shape of the deposit is fan-like (Kochel and Johnson, 1983; Kochel, 1990). Many recent deposits are bordered by boulder levees (Hack and Goodlett, 1960). Kochel and Johnson (1984) did not observe levees or terminal lobes on fans in Virginia after the passage of Hunicane Camille, but contend that much of the morphological evidence may have been destroyed by storm runoff and human intervention. Because of the predominance of steep slopes in a humid climate, saprolite rarely extends completely to the surface in the Southern Blue Ridge. Most slopes are covered with a thin layer of creep colluvium. Level landforms are usually the site of deposition of more catastrophic colluvial processes or of alluvial processes. Most studies dealing with pedology and geomorphology have encountered colluvial materials at the surface. Only near Linville Falls, 175 km northeast of the study area of this dissertation, has saprolite been reported to extend to the surface (Reed, 1964). In the Joyce Kilmer Memorial Forest, 60 km northwest of the study area, transect studies indicated that more than one third of the landscape was underlain by deep saprolite (Daniels et al., 1987). Evidence of local colluvial activity in the upper soil horizons was 3 1 almost always present. In the Blue Ridge Front, saprolite was only present on the ridges and footslopes but colluvium occurred on all parts of the landscape (Graham et al., 1990). According to the authors, the absence of saprolite on the steep backslopes is due to mass wasting and throughflow that removes water that would otherwise promote weathering. A mass balance study in the study area indicates that the short-term rate of saprolitization is about 38 m Ma'l (Velbel, 1985). In other regions of the Appalachian system, i. e. the Piedmont, saprolitization rates have been found to be much smaller (i 4 m Ma'l) (Pavich, 1986; Cleaves et al., 1970). However, the average erosion rate for the eastern US (40 m Ma'l), estimated by Hack (1970), is similar to the saprolitization rate of Velbel (1985). Taking changes in climate and soil C02 levels into account, Cleaves (1989) argued that saprolitization rates on the Maryland Piedmont must have been between 15 m Ma'1 and 48 m Ma'l since the Late Miocene. In the absence of erosion, this translates into the production of a saprolite between 80 and 255 m thick (Cleaves, 1989). Considering that the thickness of the present day saprolite is about an order of magnitude smaller, this suggests that, since the Late Miocene, significant erosion took place on Piedmont uplands. Pavich (1989) concluded that erosion and saprolitization on the Piedmont of Virginia were between 4 and 20 m Ma‘l. Pavich reached his conclusion after determining saprolitization rates on the basis of soil erosion, sediment flux and 10Be inventory studies. Although the results of Cleaves (1989) and Pavich (1989) cannot be extended to the study area of the present research without careful examination of underlying assumptions, they do indicate that the Southern Blue Ridge may have undergone considerable erosion during the Quaternary. 32 3.2.4 Soils The following section contains a general description of the soils in the study area. For a more thorough discussion of the soils on a regional scale and of soils on colluvium the reader is referred to section 5.2. The soils of the study area have formed predominantly in regolith weathered from metamorphic and metasedimentary rocks under hardwood (mainly oak-hickory) forest vegetation. Most of the regolith underwent transportation and the only landscape positions where truly residual, saprolitic soils can be found are thought to be the ridgetops and shoulder SIOpes (Thomas, in press). The dominant soil temperature regime is mesic; soils above 1460 m have a frigid temperature regime. The soil moisture regime is udic. Most soils, with the exception of three series on floodplains, are well to excessively drained. They usually are moderately deep to very deep, although shallow soils occur on steep slopes in the high mountains. Soils on floodplains and river terraces have a neutral soil reaction but all other soils are acid to extremely acid (3.520 %) valley floors of lst order streams but a majority occur in 2nd and 3rd order valleys. Their surfaces display an irregular topography and are characterized by the presence of large boulders that often are more than a meter in diameter. In some wide valleys, usually at elevations around or above 1000 m, extensive areas with a lobate topography occur. 3.3.2 Selection criteria Of all the potential debris flow study sites that had been visited, the 10 best sites met the following criteria: 43 - potential to encounter a sequence of deposits in the same local area, - perceived likelihood that the deposits were produced by a debris flow like event, 0 absence of accelerated surficial erosion as evidenced by the microtopography and completeness of the soil profiles, 0 differences in degree of soil development on the deposits of a sequence, - uniformity of the local bedrock, ° accessibility of the site to a backhoe This selection procedure resulted in a list of potential sites that spanned the basin of the Little Tennessee River. All sites had a steplike topography with two or three lobate deposits. Some sites had an additional fan shaped deposit in the lowest position of the sequence. Although surficial erosion was probably not absent from any of the sites, it was judged to be acceptable because of the absence of extensive gullying or slumping and because of the completeness of soil profiles. The local lithology was uniform in all cases and usually was comprised of quartz-rich gneiss and metasandstone. Slopes at two of the sites were steep (>30 %) and obviously near the limit of accessibility. The 10 sites were then revisited and ranked, based on a reevaluation of the above six criteria. Subsequently, the landowners of the sites were contacted, starting from the top of the list. The three highest ranked sites for which permission was obtained became the subject of this research. To obtain an accurate representation of the topography of the three sites, they were surveyed with a total station. At each site, a reference point with X, Y, Z coordinates of (0,0,0) was established arbitrarily. The deposits and the lower part of the surrounding slopes were then surveyed on a 20 m grid. At sharp changes in slope a finer grid (approximately 10 m) was used. Because all deposits of a site were not completely visible from the original reference point the total station had to be moved to secondary reference points. To control for inaccuracies in instrument set up, an already surveyed grid node and 44 the previous instrument position were surveyed again from each new instrument position. The measurements were automatically stored in an electronic field book and subsequently downloaded to a PC for processing. The contour maps, cross sections and longitudinal profiles referred to in section 3.3.3 are based on this survey. 3.3.3 Site location and description 3.3.3.1 Alarka site The Alarka site is located in East Alarka, 5 km south of Bryson City (Figure 6). The deposits occupy the valley floor of the Little Alarka Creek, about 1.7 km NW of its confluence with Alarka Creek. At the study site the Little Alarka Creek is a small third order stream. Its drainage area above the site is 1.7 km2 and has a circular shape (Figure 7). The valley bottom is at "755 m amsl, more than 300 m lower than the tops of the surrounding mountains. The average slope of the valley walls is about 35 %. The lithology of the adjacent mountain sides is relatively uniform. The rocks belong to the Upper Precambrian Great Smoky Group (Hadley and Nelson, 1971) and comprise feldspathic metasandstone, gneiss and interbeds of quartz-mica schist. No soil survey for the area is available but general field observations on lower slope positions indicate that those soils developed in boulder rich colluvium of metamorphic rocks. The valley floor has a well pronounced hummocky topography. A number of lobate rrricrolandforms cover the valley floor over its entire width. They extend along the valley axis over a distance of approximately 500 m and probably reflect the deposits of individual debris flow events. The lobate topography is best expressed in the middle of the colluvial stretch. That central part, 275 m long and 150 m wide, is the focus of the study. To the eye it appears that there are three deposits. They can be observed between +2 and -4 m, 2 and 12 m and 9 and 18 m, respectively on Figure 8. The deposits are approximately 50 m wide at their base and slightly longer than 60 m. Along their long axis they have an average slope 45 Bryson City a M-\"“‘\ _____ /' Alarka site ‘— \~\~ \. ® )\ / l/ i, l‘) ( 74, ~ . ) ' ”r \. .Wd. ‘ ~ N «\4\ 9Q\-} \‘(\/ s. .i ' 0e. . A (R?- K~\‘\.’\ \ '3 D.) at m .1 ea / ‘31 u- " \ \ / ‘ . 0/ . ‘ A-_/ m\./\\ (3‘7 \ l ,/' __,_ ’fiVer ’/\\~ €10 / /'\,’/'d / , 3" Bradley fire (I - m _— k ' \ . :’\ - . - / ‘j < \.j >;_L‘--_’\_z—~/ \ a I I) ( (JP. I ' “~./~- / ./ '\ L J / / \.\ i Hidden Valley site ~ . \‘Q 0“./ ~"\.r’ L'A‘ detr ‘\‘) "‘~——-\ __j/ BELL-“ET, 'f) I, If S ,j /.J' k / L \ \ /\-~)| \. (\‘ \\ '/\. .’\ f' , K, -\_/ \.J L\ ./ i? l \ fl 1' mm“ 1 Figure 6 46 n 2:9“. 26 >35 Ex F md /O V 000 / (we uouerg elroI _/’§\—/// 9v. 9.3: 96 33> :82: mamcm Emmm mum >235 47 Emozm w 9:9“. 8:08 890 E doc—320 mNP oo— mk 0m mm 0 mm- Om- mm- COT mNT I I fi q ‘ E .5:ng mum 9.32. 6 Q9: 59:00 0m- mm Om mm 00F mm— UJ ‘eouersga 48 of about 16 %, with short reaches that are slightly steeper and which presumably represent the fronts of the different deposits (Appendix 2). The deposits have a relative height of about 3 m and, in cross section, present a convex shape (Figure 9). The valley has only one major drainage channel, the Little Alarka Creek, which flows at the side of the deposits. This imparts an asymmetric cross section to the lower part of the valley (Figure 9). The soils on the deposits are mostly Eutric Dystrochrepts (see section 6.3). 3.3.3.2 Hidden Valley site This site is located in the town of Bumingtown, 8 km NW from Franklin, in an area known locally as Hidden Valley (Figure 6). The valley is part of the headwaters of the Iotla Branch of Iotla Creek. It has a wide, flat floodplain and contains two unnamed, second order streams that drain 0.5 km2. The floodplain is at an elevation of 640 m amsl, the crests of the surrounding ridges are at approximately 900 m. The slopes have a gradient of between 24 % and 38 %. The rocks in Hidden valley are of Middle and Upper Precambrian age (Hadley and Nelson, 1971) and consist mainly of biotite schist and gneiss. The northernmost crests of the surrounding ridges contain migmatitic gneisses, in addition to the above two rock types. Some of the gneiss is very quartz-rich and approximates quartzite. In the saprolite, highly weathered, well-drained soils have developed. In most parts the soils are very deep but on the steepest gradients and on slope breaks only moderately deep soils exist. In undisturbed settings a 20-25 cm thick A horizon overlays a reddish Bt horizon. Below the Bt horizon the amount of coarse fragments increases with depth and gradually becomes a Cr horizon (Thomas, in press). Between the two streams of Hidden Valley there are three clearly separated topographic levels (Figure 10). The levels are separated by short steep slopes between 14 and 24 m, and 36 and 44 m. The lowest level coincides with the floodplain and is about 200 m wide (Figure 11). The middle bench is about 75 m wide and has a sharp contact with the floodplain. Its front is steep (i 40 %) and about 10 m high. The top of the second 49 m 939“. m new < B cosmoo. L2 m 939”. mom m E .oocmEQ < com 09 9.: o: ONF oop ow ow ov cm 0 L _ _ if _ > _ _ . _ bi _ _ . bi hi 1— » _ _ _ . . l b‘l \ N- \\ . ///l . H o l - N .\ / H // H w 96 more? 60:08 meo w 'UOllBAGE 50 E596 |:| or 9:9“. 5:03 meo ........ E 62865 omv oov omm oom omm com of 03 on o om- oo 7 . \m. \ - em. a \ t\ m C. [\9 . \ om E.co:m>o_m . on: 2w >m=m> 522: he amE .5250 w ‘eouersgq 51 0mm 2 9:3 E @8920 m ucw < B cosmoo. L2 or 9:9... wow o=m> :82: .820me meO u: 'uoireAeja 52 bench is relatively flat (10 %) and gradually slopes up toward the front of the highest level. The highest level has a relatively steep front (i 30 %) about 10 m high, but its top is nearly horizontal in some parts. It grades into the mountain side with a well defined break of slope. The sequence of the three topographic levels is 540 m long and covers a total difference in elevation of 80 m. The soils on the floodplain are mapped as very deep, well-drained Hurrric Hapludults (Thomas, in press). The soils that were excavated for this study are Typic Hapludults (see section 6.3). Pedologically the two highest levels are not distinguishable from the hillsides (Thomas, in press) and their soils can be categorized as Typic Hapludults. 3.3.3.3 Bradley Site The Bradley site is located in the Oak Grove community, 12 km NW of Franklin, just off state highway 28 (Figure 6). It is 7 km north of the Hidden Valley site and approximately halfway between the Alarka and Hidden Valley sites. The deposits occur in a small, unnamed, tributary valley of Bradley Creek. Further north other tributaries of Bradley Creek have produced various small alluvial/colluvial fans. The unnamed valley studied holds two second order streams that drain 0.2 km2 (Figure 7). The basin has a relative relief of 350 m, between 650 m and 1000 m amsl, and slopes with an average gradient of 36 %. This site has a more elongated shape than the basins of the two other sites (Figure 7). The basin is crossed by the Hayesville fault which separates the Great Smoky Group from Middle Precambrian rocks (Hadley and Nelson, 1971). Gneisses, usually rich in quartz, dominate the lithology, but quartzites, metasandstone and migmatites are also present. The soils on the slopes are very similar to those at Hidden Valley, being well-drained and moderately to very deep. Their profiles usually are characterized by reddish horizons and by coarse fragments that, with depth, gradually grade into a Cr horizon. 53 This site is also similar to Hidden Valley in that two creeks border the colluvial deposits. In their middle reaches the creeks are deeply incised in separate V-shaped ravines. Near their confluence with Bradley Creek the ravines unite and become one wide valley with a relatively smooth valley floor. The creeks occupy the two sides of the valley floor, rather than an intermediate position, and are barely incised at all at this point (Figure 12). Distinct deposits or topographic levels are not as readily identifiable here as at the two other sites. Only slight topographic undulations, l to 2 m in height, hint at the possible presence of different deposits (Appendix 2). The colluvial material on the valley floor has a drawn out shape and is 330 m long and 150 m wide. At the confluence with Bradley Creek the deposit is no longer confined by valley walls and becomes fan shaped (Figure 13). The slope of the valley floor increases from about 12 % to more than 20 % on the fan-like front. As in Hidden Valley, the soils of the deposits are mapped as well-drained Humic Hapludults (Thomas, in press). The present research shows that they are Typic and Umbric Dystrochrepts. 54 NF 9:9“. E 605320 ow om m can < *0 cozmoo. .2 m: 959“. mom or mw irrv 0N rvii mm Irri om llll mm 9% >935 60.6me $9.0 w 'UOlIBAGB 55 9 9:9". mNP of. mm Om mm Emwzw 8:08 890 E 6958.0 0 mm- Om- mm- 007 mNT OmT mmT ooN- mNN- lulj Jr xxx... E cozm>w_m - J V i/ Egg V MM ozm >235 do amE 59:00 .00 ..- ur ‘eouersicj PART II: CHARACTERIZATION OF SEDIMENTS AND SOILS 56 57 4 INTRODUCTION The sedimentology of debris flow deposits has been studied extensively (Shultz, 1984; Rust and Koster, 1984; Wells, 1984; Wells and Harvey, 1987; Walton and Palmer, 1988; Ono, 1990; Whipple and Dunne, 1992). Many studies of debris flows have been carried out in the Appalachians (Williams and Guy, 1973; Mills, 1983; 1988; Kochel and Johnson, 1984; Pomeroy, 1986; Kochel, 1987; 1990; Whittecar and Ryter, 1992). The knowledge acquired in these studies will be applied in this part to examine if deposits at the three sites may have formed by debris flow activity. The characterization of the sediments will also provide a better comprehension of the nature of the parent material in which the soils have developed. Relatively little information is available about the strength and variability of the fabric of debris flow deposits (Lawson,l979; Mills, 1984; Innes, 1984; Mills, 1988). This part will, therefore, not only use existing information but may also provide new data on the sedimentology, i.e. fabric, of debris flows. This kind of data may be important to better understand the origin, and possibly the mitigation, of debris flows. Few studies have been conducted on the soils of the Southern Appalachians (Norfleet and Smith, 1989). Recent papers include those by Daniels et al. (1987a; b), Norfleet and Smith (1989), Graham et al. (1989a; b; 1990), Graham and Buol (1990) and Feldman et al. (1991a; b). Although most of these studies involved some aspect of colluvial soils, none of them explicitly dealt with the pedogenetic imprinting on debris flow deposits. Chapter 6 will generate characterization data for soils that have developed on debris flow deposits and will relate them to the origin of the parent material. In later chapters these characterization data will be compared with relative-age estimates for the deposits. To maintain consistency throughout the text, "coarse fragments" will be used to designate particles >2 nun. It is equivalent to the sedimentological term " gravel". In accordance with Soil Survey Division Staff (1993) guidelines, "gravel" was reserved in 58 this research for equiaxial particles of between 2 and 75 mm. "Fine gravel" refers to particles of between 2 and 8 mm. 59 5 SEDIMENT S 5.1 Literamre review 5.1.1 Sorting By definition, debris flows consist of a mixture of clasts and a matrix material. The clasts present in debris flows typically are angular to sub-angular (Williams and Guy, 1973; Nemec and Steel, 1984; Wells and Harvey, 1987). Most studies describe debris flow deposits as a diamict in which both the matrix and the clasts are poorly sorted (Innes, 1983; Kochel and Johnson, 1984; Rust and Koster, 1984; Wells, 1984; Pomeroy, 1986; Wells and Harvey, 1987; Walton and Palmer, 1988; Kochel, 1990; Ono, 1990; Whipple and Dunne, 1992). Wells and Harvey (1987) studied viscous and fluid flows but did not find a difference in sorting or mean particle size of the matrix or the clasts. Average sorting coefficients reported in the literature range from 3.6 to 12.3. Maximum clast size can range to more than 8 m (Walton and Palmer, 1988; Shultz, 1984). No textural variation was found in the downslope direction by Williams and Guy (1973), probably because the deposits were too short (<1 km) for sorting to occur. Other Appalachian debris flow deposits also show very little change in their sediments from proximal to distal reaches although they are generally thicker and coarser at the apex (Pomeroy, 1986; Kochel, 1990) The lack of change can be explained because many debris flows retain their competence throughout their travel (Kochel, 1990). Mills (1988) studied various sedimentological parameters of colluvium in the Valley and Ridge province of Virginia but did not distinguish between debris flow deposits and other types of colluvium. His general comments suggested that the colluvium he studied was poorly sorted. The clasts in the colluvium were more angular (2.65 on the visual 60 comparison chart of Krumbein, 1941) and changed less in angularity with distance from the source than pebbles that had traveled a comparable distance in nearby creeks. 5.1.2 Bed geometry Beds in debris flow deposits are usually described as tabular or sheet-like (Bull, 1972; Rust and Koster, 1984; Shultz, 1984; Wells, 1984). Nemec and Steel (1984) noted that beds showed limited basal erosion but often had a highly lenticular overall geometry. Other authors also recorded that beds of debris flow origin have non-erosional, though sometimes clear, boundaries (Gloppen and Steel, 1981; Costa ,1984; Nemec and Steel, 1984; Wells, 1984; Velbel, 1987; Kochel, 1990). 5.1.3 Stratification Beds in debris flow deposits usually show no obvious stratification (Nemec and Steel, 1984; Rust and Koster, 1984; Smith, 1986; Wells and Harvey, 1987; Mills, 1988; Kochel, 1990). Some debris flow deposits have a crude layering resulting from an upward- coarsening sequence (Wells and Harvey, 1987). A succession of deposits may exhibit pronounced bedded due to distinct bed boundaries (Nemec and Steel, 1984). A tendency for channeling and some faint stratification were associated with dilute flows by Nemec and Steel (1984) but Wells and Harvey (1987) did not find stratification in fluid flow deposits. 5.1.4 Grading Grading in debris flow deposits has been found to be absent (Nemec and Steel, 1984; Shultz, 1984; Smith, 1986; Walton and Palmer, 1988), inverse (Williams and Guy, 1973; Rust and Koster, 1984; Kochel and Johnson, 1984; Walton and Palmer, 1988; 61 Kochel, 1990; Ono, 1990), normal (Nemec and Steel, 1984; Shultz, 1984; Walton and Palmer, 1988) and reverse-to—normal (Rust and Koster, 1984; Smith, 1986). Nemec and Steel (1984) called upon turbulence to explain normal grading but it is usually thought to be the result of differential settling of the clasts in a low yield strength material (Scott, 1986; Smith, 1986) that moves as a pseudoplastic flow (Shultz, 1984). Inverse grading has been interpreted as being the result of dispersive pressure (Nemec and Steel, 1984; Walton and Palmer, 1988). The dispersive pressure develops from shearing, usually near the bottom of the flow and is restricted to deposits with low original clay content (Smith, 1986). The grading of the basal few cm of a deposit can change downslope from ungraded to inversely graded (Nemec and Steel, 1984). This can be related to a change in the flow regime, i.e. development of shear, downstream (Shultz, 1984). Bipartite layered beds with the coarser one near the bottom of the deposit may originate from the settling of coarse material during the deposition of a fully turbulent flow. A stratified sandy capping with an erosional base or a stratified Clast-rich cap with a sharp lower contact can represent interflow fluvial activity (Nemec and Steel, 1984). 5.1.5 Conglomerate framework Rust and Koster (1984) stated that matrix supported diarrricts in which the matrix consists of unstratified muddy sand, mainly form by debris flow deposition. Matrix supported diarrricts of debris flow origin were also observed by Gloppen and Steel (1981), Shultz (1984), Watson and Palmer (1984), Wells (1984), Whipple and Dunne (1992) and Whittecar and Ryter (1992). Clast supported facies are often produced by fluvial processes (Miall, 1977; Rust and Koster, 1984; Smith, 1986; Kochel, 1990) but have been observed in debris flows as well (Gloppen and Steel, 1981; Nemec and Steel, 1984; Shultz, 1984; Wells, 1984; Walton and Palmer, 1988). Wells and Harvey (1987) distinguished two types of debris flow, viscous and fluid. In the viscous debris flow deposits they noted a 62 predominance of matrix-supported clasts, except in pressure ridge areas. The pressure ridges probably represent pulses or surges commonly observed in debris flows. The fluid flow deposits were characterized by a mud-rich matrix, supporting the clasts. Fan deposits studied by Kochel (1990) in central Virginia were matrix supported while others from western Virginia were clast supported. The author argues that the difference depends upon the amount of water that was involved in the flows that produced the fans. Flows in central Virginia were associated with less runoff because they formed in basins that were smaller and of a lower order than-those in western Virginia (Kochel, 1990). Consequently, they might have been flows with low amounts of water. 5.1.6 Fabric shape Debris flow deposits often present an isotropic fabric (Lawson, 1979; Innes, 1983; Nemec and Steel, 1984; Wells, 1984; Wells and Harvey, 1984). The disorganized Clast fabric may reflect short travel distance but often suggests plug flow or weakly sheared, high viscosity flow (Nemec and Steel, 1984) or, according to Lindsay (1968) and Owens (1973), turbulent flow. When there is a preferred clast orientation, the clast long (a-) axis is often subhorizontal and parallel to the flow direction (Nemec and Steel, 1984; Wells, 1984; Mills, 1988; Kochel, 1990). This fabric may originate from strongly sheared laminar flow, most likely due to clast interactions and dispersive pressure (Lindsay, 1968; Nemec and Steel, 1984). Traverse alignment of clasts can be expected to develop from more fluidal flows (Mills, 1984; Nemec and Steel, 1984; Smith, 1986; Kohlbeck et al., 1994) but Wells and Harvey (1987) did not find any preferred clast fabric in fluid flows. The latter authors did observe, however, that the long axes of clasts were typically oriented perpendicular to the flow direction in streamflow deposits. Shultz (1984) found a-axis orientations to be 63 clustered in two modes, parallel and perpendicular to the flow direction. It is not clear whether the samples came from viscous, fluid or both types of debris flow deposits. The viscosity of the flow also influences the dip of the a-axis. More fluid debris flows show subhorizontal orientation of clasts (Rust and Koster, 1984) whereas more viscous flows tend to have larger clasts in predominantly vertical orientations (Bull, 1963; Lawson, 1979; Nemec and Steel, 1984; Rust and Koster, 1984). The vertical orientation can only develop in matrix supported units that are viscous enough to support the upright position of the clasts (Bull, 1963; Reineck and Singh, 1986) or in clast supported units (Lawson, 1979). An alternative, but less likely, explanation offered by Lawson (1979) is that the vertical positions develop prior to flow and are preserved within parts of the flow that do not undergo shear. Most authors found a weak imbrication in debris flows, but found various directions of imbrication. Nemec and Steel (1984) state that clast upflow imbrication of the a- or b—axis is common in debris flow deposits. It is not clear whether they are referring to subaerial or subaqueous flows. Lahar deposits, especially matrix supported facies, also showed an upslope dip (Mills, 1984). The debris slope deposits studied by Williams and Guy (1973) and Kochel and Johnson (1984) exhibited poorly developed imbrication of the coarse clasts in the downslope direction. Yeend (1969) recorded an apparent downslope preferred orientation in two debris flow lobes. In one of these a secondary mode dipping upslope was identified. In small glacial debris flows, slightly more than half of the sampled sites had an upslope dip, and consequently, less than half had a downslope dip (Lawson, 1979). All of these sites were characterized by a very gentle dip of the a-axis of the clasts. In measurements of clast orientations in debris flow levees, Innes (1983) noted a preferred upslope orientation of the clasts that was intermediate between the flow direction and the local slope of the levee surface. The author speculated that the orientation was the result of an initial orientation parallel to the flow and a subsequent settling parallel to the direction of maximum slope. Mills (1988) noted that in the colluvium he studied, preferred 64 dip angles tended to be lower than the slope angle. However, the fabrics generally had an upslope imbrication with respect to the surface. In the debris flow pressure ridges which Williams and Harvey (1987) observed, clasts were oriented with their a-b planes dipping away from lobate boundaries and into the deposit. 5.1.7 Fabric strength Study of the orientation of clasts in different types of debris flows is critical to the evaluation of the classification of debris flows (Schultz, 1984). Nevertheless, very few studies on the fabric strength of debris flow deposits have been carried out and quantitative studies are almost completely lacking. Lawson (1979) studied debris flows that formed at the temrinus of a glacier. Visual observation of contoured Schmidt nets led him to conclude that in general the strength of the fabric in these deposits is weak but that the strength increases with increasing water content. According to Lawson (1979), shear, and thus clast alignment, occur only in a thin basal zone when the flow is relatively dry but as water content of the flow increases the shear zone thickness and clast alignment increase. Mills (1984) studied lahar facies in Mount St. Helens debris flows and examined clast supported and matrix supported units separately. The first and third eigenvalue was larger for the matrix supported units than for the clast supported units (Table 1). This means that although the matrix supported units have a somewhat stronger tendency to cluster around the mean direction, the fabric strength of the two facies is comparable. Their strength is also comparable to that of the small, glacial debris flows studied by Lawson (1979). In creep colluvium the a-axis of clasts was also parallel to the slope and imbricated upslope (Mills, 1983). In that environment the fabric strength was directly related to the gradient of the slope. Elevation, aspect, and soil texture appeared to have little effect on fabric strength (Mills, 1983). The strength of the fabric varied between different local 65 environments (Mills, 1988). The conclusions of Mills (1983) and Innes (1984) imply that fabric is not only a reflection of the genesis of the sediments, but also of local slope gradient and direction. Table 1: Clast fabric data for debris flow-like deposits. Author Deposit Deviation of mean a—axis S 1* 83* H from direction of flow Mills St. Helens lahar 50° 0.564 0.140 ll ( 1984) (all samples) Mills St. Helens lahar 65° 0.555 0.133 ll (1984) (clast supported) Mills St. Helens lahar 37° 0.572 0.147 (1984) (matrix supported) Lawson Glacial debris flow 28° 0.567 0.128 (1979) * S1=first eigenvalue, S3=third eigenvalue, see section 5.2.2 for explanation 5.2 Methods 5.2.1 .Data collection As mentioned in section 3.3 (site selection), a number of possible deposits were identified at each of the three sites. Because different microsites on colluvial deposits often possess different pedologic characteristics (Daniels et al., 1987; Graham et al., 1990) an attempt was made to open two backhoe pits on each deposit, one on the apex and one on the foot. To create optimal conditions to study lateral variations in sedimentary facies the pits were dug with their longest side (3 to 5 m) perpendicular to the central axis of the deposits. Pits were excavated to a depth of approximately 2 m. 66 The soil pits, and the pedons described in them, were assigned a three letter code (Table 2). The first letter stands for the site (A - Alarka, H - Hidden Valley, B - Bradley), the second letter stands for the position of the deposit in the sequence of each site (L - lower, M — middle (if more than 2 deposits), U - upper), and the third letter represents the position of the soil pit on the deposit (f - foot, a - apex). Because no soil pit could be opened on the apex of the lowest deposit in Alarka, a pit on the foot of another low deposit was opened. To distinguish between these two localities, a fourth digit ("1 " or "2") was added to the code of the pedons on the lower Alarka deposits. Table 2: Geomorphic position of soil pits. Site/Soil pit Deposit Slope Slope Aspect Depth of soil number position % pit, cm Alarka ALfl 1 foot 15 W 187 ALf2 2 foot ’ 16 w 193 ‘ AUf 3 foot 16 W 195 AUa 3 apex 16 W 213 Hidden Valley HLf 1 foot 5 SE 200 HLa 1 apex 11 SE 187 HMf 2 foot 15 S 223 HMa 2 apex 15 S 248 HUf 3 foot 13 S 227 HUa 3 apex 23 S 211 Bradley BLf 1 foot 12 W 125 BLa 1 apex 16 SW 187 BUa 2 apex 19 W 208 67 At each location sediments were described according to guidelines reported in Miall (1977; 1978; 1985), Schultz (1984) and Eyles et al. (1983). Properties recorded include nature of bounding surfaces, internal geometry, external geometry, and scale of beds. With regard to internal geometry, special attention was paid to grading, sorting and framework support. To characterize clast properties, 50 clasts (a-axis 2 2 cm) were selected at random and their size (length of a-axis), sphericity, roundness, lithology and relative weathering determined (Table 3). Table 3: Soil horizons sam led for clast characterization and fabric. Pedon Horizons sampled for clast Horizons sampled for clast characterization * fabric * ALf 1 all all ALf2 all all AUf all all AUa all all HLf 2Cg none u HMf Btl, Bt2 none HUf 2Bt2 2Bt2, 3Bt3 3Bt3 none HUa all none BLf 2Cg2 none BLa 2Bw2 4Bw4 4Bw4 none BUa 3Bw3 5Bw5 5Bw5 none * For corresponding sedimentary units, see Figures 19 and 20 Sphericity and roundness of the clasts were estimated by visual comparison with a chart from Powers (1982). For 30 other elongated clasts the azimuth and plunge of the long (a-) 68 axis and the dip and strike of the a-b plane were determined (Table 3). This field procedure consisted of carefully excavating clasts with an estimated a-b axial ratio greater than two, and measuring the dip and strike of the a-b plane. The clasts were then removed, an aluminum rod was inserted in the position of the long axis and the orientation of the rod determined with a Brunton compass. 5.2.2 Data analysis The coefficient of sorting for the matrix (grains < 2 mm) of the sediments was calculated with the formula So=\/Q3/Q1 in which Q1 and Q3 are the first and third quartile of the cumulative particle size distribution. The quartiles were graphically determined from graphs of the particle size data from soil horizons. Sorting is therefore reported for soil horizons rather than sedimentary units (Appendix 2). Facies codes are based on Miall (1977; 1978), Eyles et al. (1983) and Shultz (1984). The codes consist of two to three letters. The system of Eyles et al. (1983) was specifically designed for diamicts and uses a "D" as the first letter of the code. The second letter, "m" or "c", identifies either matrix support or clast support. The third letter refers to internal structure and separates massive unstructured organization (m) from stratified (s) and graded (g) units. Shultz (1984) used "g" for normally graded units and added "i" to represent reverse grading. The system of Miall (1977; 1978) was designed for use in fluvial environments. His code consists of two parts. The first part is a capital letter G, S or F, which represent diarrrict, sand and fines respectively. The second part describes the most characteristic internal feature of the lithofacies. The only code of Miall used in this study is F1, which stands for fines (sand, silt, clay) that are laminated. The orientations of the a-axis and the poles of the a-b planes were plotted on Schmidt nets with an equal area, lower hemisphere projection. To allow for comparisons between samples, the local slope was also plotted (Appendix 4, 5). Clasts that were 69 oriented traversely to the axis of the deposit (and presumably traversely to the direction of flow) were defined as those that had a long axis azimuth that was within any of two 45° sectors centered on the orthogonal on the axis of the deposit. Although quantitative methods to describe particle orientations have existed for a long time, relatively little effort has been invested in statistical analysis of fabric. For this study, the fabric of the diamict layers was analyzed with the eigenvalue method developed by Mark (1973) and Woodcock (1977). Successful applications of this method have been achieved for solifluction lobes (Nelson, 1985), till (Lawson, 1979; May et al., 1980), creep (Mills, 1983b) and debris flows (Mills, 1984). Woodcock (1977) contended that the use of eigenvalue plots may reveal fabric shape variation, possibly with genetic significance. Mills (1984) suggested that Clast fabric may be a useful property for the characterization of debris flow sedimentology, and that its measurement should be included in future studies. Lawson (1979) and Nelson (1985), however, cautioned that additional information beyond that of clast fabric is needed to positively identify the origin of a deposit. The eigenvalue method is based on the principle that eigenvectors represent the common variance in a data set and that all directional data can be completely described with respect to three orthogonal axes. Since eigenvectors are perpendicular to each other, three eigenvectors will completely describe the common variance in a directional data set. The first eigenvector may be considered as the direction around which the data are most concentrated. Conversely, the third eigenvector is associated with the direction that least coincides with the directional communality of the data. It is orthogonal to the best plane through the data. The second eigenvector is normal to eigenvectors 1 and 3. The first eigenvalue is a measure for the common variance in the data. If eigenvalue 1 (81) is high, the data are highly clustered. The third eigenvalue (S3) describes the remaining variance that can not be described by eigenvalues 1 (S l) and 2 (S2) and thus is a measure of scatter. The eigenvalues reported in this study are standardized eigenvalues, 70 also called significance values, that range from 0 to 1. More important than the magnitude of the eigenvalues, however, are their ratios which represent fabric strength (Figure 14). The ratio of S 1/S2 represents the goodness of fit of the data distribution to a uniaxial bundle while S2/S3 represents how well the data conform to a uniaxial girdle. The ratio K (K=(S1/82)/(S2/S3)) represents the fabric shape. Distributions with K=1 have equal girdle and cluster tendencies. Low values for K (K<1) define girdles, higher values define clusters (Figure 14). The deviation of the azimuths of the mean a-axis from the direction of the slope of the surface is reported in absolute and relative values. The absolute deviation is the mathematical subtraction of the azimuth of the mean a-axis from the azimuth of the slope. The supplement of the absolute deviation is provided as the relative deviation when the absolute deviation is more than 90°. The relative deviation is provided because the absolute deviation turned out to be large in most cases and large angles are harder to visualize than small angles. Moreover, there is no need for the absolute deviation because the direction of the orientations is shown by the plunge. The deviation from the general orientation rather than from the direction of the slope also furnishes additional information. 5.3 Results 5.3.1 Sediment description 5.3.1.1 General The boundaries between sedimentary units, as determined in the field, were rarely clearly distinguishable. This was at least in part due to the pedogenic alteration of the deposits, but may also reflect gradual boundaries between the units. Virtually all sedimentary units were classified as diamicts. They appeared to have a tabular geometry that blankets the deposits. This finding may be an artifact of the small size of the soil pits which did not allow the lateral extension of the sediments to be verified. No natural Fabric shape and strength as a function of 81/82 and 82/83 7 __ K \i/ ,/ 1] if?) /’ // \, \__// , (V/ .\\J/ \\ / /// U1 l \ /‘ 7‘\ s. \ \\ 4.. \ ‘ I \ A ,I I US“ \ X/ /> \ , ’ \ / 5 3 z \_ ( :é) .5405 |n(82/83) Figure 14 72 exposures were available to substantiate the tabular geometry. The bottom of the deposits was not encountered in any of the pits. Therefore, the sediments must have a thickness of 2 2 m in most cases. Interpretative sketches of lithofacies units are provided for the Hidden Valley and Bradley sites but not for the Alarka site, because no sedimentary boundaries were observed at the latter. 5.3.1.2 Alarka site. The sediments at the Alarka site are very poorly sorted (Appendix 4). Some contain more than 50 % coarse fragments (Table 4). In all four Alarka soil pits, the amount of coarse fragments decreases near the surface (Figure 15). This is, at least in part, due to human influence. The area near the Alarka site has been farmed extensively and large rocks near the surface have been systematically removed by the farmer. They were used to build a rock wall around one of the fields. It is virtually impossible to verify directly if the trend is also a natural one because most debris flow deposits in the area have been farmed, since they constitute the most level part of the landscape. However, in ALfl and ALf2 the fine gravel and sand fractions follow the same general trend as the coarse fragments (Figures 16, 17). Although increasing amounts of sand with depth may be due to decreasing weathering intensity with depth, the accordance of the trends of the three different fractions suggests that the trend is also in part natural because fine gravel and sand were not removed by the farmer. Consequently, the sediments at these two locations are normally graded. Their lithofacies code is therefore coded Dmg. Pedons AUf and AUa are located on a higher deposit. The fine gravel and sand fractions of these two pedons do not have a uniform trend with depth (Figure 16, 17). This indicates that the trend of the coarse fragments at AUf and AUa is mainly artificial. The lithofacies of the sediments at these two soil pits is therefore coded Dmm. 73 low + 00 + .m mm .m ... + cum 2me b0> new .0 o. E .3 8 x0 Sm ~50. +mo..wm. Nomm 9m + .00 + .m ow 880.8. .m. c + ..0 mm 030.... b0> ..0 .0 u 3 .w 88 .E .3 8 x0 9w 53.» Mam—-3 .Umm .0>a o... 000. .8 c + 8 cm 6.8.5 ..U .0 .. a a as .E .E a a 3m a»: 3% EN .0>8 m. . ...+..0 N. 03a... 3 .w new .E .E E 3 .w v? ~50. mmbm m Sam. . ... + ..0 w 0305 3 .0 w .E .3 cw Qm a»... 09m. N< .03mw o c+~.0 v 0.80.... m .m w .E .E .m0 Qm 56. 2.0 .< 64¢ 9m + 8 + .m mm 88.0.02 .8 ... + no we 030.... b0> ..0 .0 .. [MW .0 .3 a go Ev arm... +57%. NUmm Sm £3. .E .3 + 00 + 0» Om 0.08.0.8. 2 E58... .00 c + .6 9. 6.9.5 36> ..0 ... .a .6 9 m a .E .3 a a 2m 0;: 02-8. 6% 9m + 00 + .m Om 008.00; .m. c + no on 0.3.0:. ..0 .0 u _ .m n8 .E .E c x0 Sm «>08 mo.-0m .mm 0.0 + 00 + kw mN . ... + ..0 Q... 0.905 a .w £3. .E .E 2 3 ... b0> $m ES... omém Mm .0>0...w v ... + ..0 w 0305 3 .w sam .. .3 . m? m»... $0. m4. .26 m ... + ..0 m 030.8 3 .0 m .E .E . QM ~50. 0.6 < Cu? «m. mango.“ 3:258”. EOE 03908 5:30.... 8.00 ..0232 Eu 0080 8:03.200 438.com 25:0 .033 05.03% 0.53.0... 3.03. Goon. canto... 0....ok. 3.8.3.5000 .8003. .v 030... 74 M36 m>md 9.0 0.6 050... .0 .. .0 000 .0 .E .0. 0.0 m»... +m.m-ma. Um $wa Om .0 c + EH 6380 m .6 ea .6 .E a 6% ES. 320. 00 0.0 + 00 + 0» Om .0. c + film 0.00.5 m ..6 000 .0 .E 00 b0> Maw 58. .078. NUWN cum + 00 + 0» On .0. ... + n0 m 0.00.5 m .0 000 .0 .E 00 b0> Em m>n... 87% .09... .0>0mw ON 0% .0. ... + ..0 m 030...”. ..0 .0 u 3 .m 000 .E .E ..m . 9m MVW... am-wm .m .0>Ew o. 0.00.... 3 .w w .E .E . v? Mtrm... mm-.. 3m 3% m 0.00.... 3 .0 rm .E 18 . Sm firms To < 03¢. ...0'0 .5 .E 0.0 + 0w ov 0. $5.80 .0. c + ..6 o. 6.020 3 .0 Pa. .6 .E a 06> 6% Es. +3.8. 8mm 0.00 .E .E 0.0 + 00 ow 0. 05:00 .0 c + ..6 o. 656:. 3 .0 ea .6 .E a .0613 we ES... 8.00. 5% 0.0+ 00 + .w ow 000 .0. c + ..0 0. 0.3.0.3.. .0 .0 ..> 3 .m 000 .E .E 0. 3 00 30> ...m mwms £7... N3m 00 + 0m m. . c + ..0 m 0.00... 3 .w 000 .E .E ..w wk m>ms 2...? .3m. 0.0 + .w om .0. ... + ..6 m 636:. 09 3 .0 ea .5 .3 a 3.. 0.33 06-x m tam. . c + ..0 m 030.... 3 .0 w .E .E ..w .R mi: 3.0 < 5< 00 000300. 00:030.. 00.08 0290:. 5:00.. 00.00 :85: 80 00.000 8:80.050 -3800”. 0:030 .033 0020.5.“ 058.0... 00.02 500A. 3.00.00... 0&0... 8.2.00. 000.09.00.80 0000A. .v 030... 75 w\m M>m$ 0.00.00 .0 .0 .6 08.656300 .60 0.0 0.6. 000-000 000 Q0 yrs. .300» m 0.005 .. .. .0 3 .0 000.000.00.50 .00 $0 dams mmNKw. 00. $0 MWm... :50 .0 .E .0 a .0 000.226....m .6 0.0 0.6. 00.00. 000 030.0... .3 .0 00%30550 .00 wxm firm... no.6»: UN 0.005 3 .w 000 .E .3 .0 0% ”gm... cm.-. .. 0 g 0.00.5 ..0 .0 .. 3 .0 000 .E .E 00 3 .0 0.0 Mtg.» :73 Um. 00. 0:0 080009. 0.0000 .6...V .6 a .0 000 .E .E 6 0.0 0.000 00-00 00 00% 6.0000 .6 ... .e a .0 ......M .0 .Ie 6 0.0 0.00 00-00. :0 Mm 0000.09. .. .E 0. mink. C000. 6.0000 .6 ... é a .0 000 .6 .e .6 m0 0... 0.00 00-00 $0 0. .E .3 00 0.0.0000. [0.005 3 .0 000 .0 .E . N...” ~50... mm-.. < 0‘... 0.00.5 30> 000.000.6300 .0 0000.060...E $00-00 . Om 0.0 y; m... 00. 8 + .0 00 :50 .0 .0 .0 a .0 000.226050 .0 6.6 ..0 ES. 00...... 000 $0 m> 0.0 .2000 0 6.0000 .0 .E .E a .6 ..00 .a .5 .0 0.0 0.6. o. .00 00 onm 0.00.5 .0 .. ..> 3 .0 000 .0 .E . 0% ~50. No-03 Um g 6.0000 .6 .0 .0 3 .0 000 .E .E . 0.0. ”.000 00-00 0m I. 0.005 3 .0 w .E .0 . ...m ~56. and < ....“ 0.0 00.5.00. 0.005000. 00.05 0.00008 b00550 00.00 ..00002 80 30000 000000.000“. 40803. 0.00000 0030... 00:00:05 0.5000... 00.0.). 50.00 000.00... 0.0.000. 8.0.000. 000.09.000.00 00000. ... 0.00 ... 76 a 50> 000.05.03.00 .00 0.0 MW... +w0~0m~ .U a 3.0.. m. U 0.0. a»... 0000.000 9.0.. .0>0.m m 0.005 .0 .0 .. 3 .0 000 .E .3 .0 m. .0... Mo. 08-02 0E0 040.0% .00.» m 0.005 .0 .0 .. 3 .m 000 .... .... .0 .0... Mo. £70.. 00m. w NE .3 0. $5.00 .90.» m 0.005 0 .0 000 .0 .3 .0 w... M>0d 00-00 0mm. w .... .3 0. 0.5.0.. .0>0.m o. 0.005 3 .m 000 .... .3 . . w... m>0 00-? 07". .0»me m 0.005 3 .0 w .... .3 .00 v.0 ~50 mvém 0< % 0.005 .0 .. ..> 3 .0 w .E .... .0 0.0 m>0.~ w~-0. m< 0.005 3 .0 J .E .0 .0 v.0 M>0N 070 < 0.)... a 90.. m0 0. 0.0 0.08. 0000.000 9.0.. 0.005 .0 .0 .. 000 .... .3 0 m. 0... m>0.N +0800. 05 0000 0.005 .0 .0 .. . .0 000 .E .3 0 0... M>0.~ 007.... Um. 008 0.005 .0 .0 .0 . .0 000 .... .3 0 0.0 m>0.N v. 70.. Gm. 0% 000 0.005 .0 .0 .. 3 .w .0 0. ... .... 0. 3 0 0... M>0N 0..-?” ..m 0.005 3 .0 m .... .3 . 0... M>0 00-0. m< 0.005 0 .w w .... 0. . .3 . v... M>0 070 < .E 00 00.300. 0.00%... 00.0... 0.00.0... b00550 .0.00 ..00:=.>. E0 00.000 000000.000“. 408.00% 0:030 .033 0.3050 0.0.0.0... 00.0.2 anon. 00008... 0.0.0.0. A0500. 000......0000 0000... ”v 0.00 ... 77 8 t0 0.0 0.0.. .4. 0. ... + ..0 00 0.005 b0> .. .E .0 000.05.00.00 ..0 ..0 0.0 ~50. +0N.-0a NmUN 0.0 + 8 + 0.0 N 0.00... 02. a .0 30.22250 .0. 0.0 0.6. 00-00 .00 80 + 00 + 0.0 N 0.005 0 .0 000 .... o. .0 .3 . 0.0 ”$00.0 00-NN 3m. m .... .3 0.0 0. 0.5.0.. 000 + 8 + 0.0 N 0.00.... a .0 .... 0:0 0.. .... . 0.0 ”.00.. 00-0 < 0.... “001.08 ...0I0 ..0 + 0.0 N 6.... ..0 .0 .0 .E 0.. .E ...-h» .0 0.0 m>0N +. .N-.oN .m0 000 ..0 + a... N a... a .0 .E o: ..0 2 a . 0.0 5.0.. .00-00. 03% 0.0 arm ..0 + 0.0 N 0.000. .. .. 0 a .0 000 .... .3 . 0.0 0.00 00.00. 03.... 0.0 ”:00 00 + 0.0 N 0.005 .. .. .0 3 .0 000 .... .... .0 0.0 ...;0 ON.-N.- N3m [.03. 0.00.0 . a . ..0. .E .3 .m 0.0 00.0 0:0 .30 w .0 .3 0. 0.5.0.. ..0 + 0.0 N 0.005 0 .0 000 .... .3 .00 0.0 500 ..0-0. <0. ..0 + 0.0 N 0.00... b0> 0 .0 0 .... .... .0. 0.0 ~50 0.-.. < 0D... 0000.. .0>0.0 0 0.005 ..0 .0 .0 000 .... .3 0 0.0 «>0N +0NN-00. 050 mm“... 0 .0000. 0 8 + 0.0 0.. 0.000. ..U .0 .0 a .0 ..0. .5 .E .0 0.0 0.00.0 00.-.... 9% 000m 0 8 + 0.0 00 0.00.... ._U .0 .. ... .0 ..0. .E .... .0 0.0 0.00... 0. .-0. a... 000% 8 + 0.0 0 I 0.00.... ..0 .. .... ... .0 000 .e .3 0 0.0 «>00 0...... .... 0.00.... 09 3 .0 0 .5 .3 .U 0.0 0.00.0 0...... <0 w a .... o. 0.....00 0.00.... _ .0 00m .0 .... .0 0.0 0.00 0.0 ... 5: 0.. 00.0.00. 0.00:.005 .08... 0.00.0... 0.00500 .200 ..00032 ...0 00.0.5 0000.00.80 $0800.... 0:050 .033 0.30.50 0.0.0.0... .063. 0.000. conic... 0.00... 6.0.50. 0:00....0000 .5000 .0 0.00 ... 78 .mofizmHSm ,moEnooHoU ._o>8wu:w .mmacnc £88820 ”8:252“ 888 5m 8:36”: 458$ 5E6 3:3 8588 0390:5088: mo fiasco“. .5"— doc—583 .EEBEHE .o:cnae_m€ 880$ 856$ uEEoE?oco: we aim 8m ..ASNEHE doEEooHo .309“: x38 :0 :3: 8:5 3:58: oEEofiaxono: ho 5:525 Sum .20? Egon—o 5:080? EEC 288 we wiv— Sm 48:89:“: 8:56”: 85“: 5:283 9:803 288 :0 3050536 5m SEWEHE .:oEE8Ho .Bownu .33 b9“? x280 Ho 2qu? 6:5 288 no E85: 8m .EEmoEfl £885": 499$”? .fiooEmum 5E6 280$ Dawson 832 we £92wom9 :om .omstmnnu ..asumanw 820.8 .338": x38 we :3: 8:5 8:55: 832 ..o 30:855.: hon dug—mtg”: .383 523:: nsmunmm 52:8:me 5:283 EEC 338.8 ‘8 085. EL .888 bo>uo> 688"”: .oELfi .EicoEnE ”9mm: “E803 BBQ—Em we aim :om .EEUQEHE .fioBNB A88 mo :3: 8:5 238.8 “0 092m Sm .bquhxonxo S=o>flwv¢w ..Awmmcnc ,bo::d:onco $53088 Sioflo .82 Egon—o .82 3:8”: :82 >20 high—8 :82 3:8 om38n_8 :82"— .82 3:3 “Emu—fl 65:8: Sm “950:8 m: 2: m:o_§>obn< s. j + : R e; .5 9m .3 9 Mirna _ + 8 + :m mm 288 :3 .o .3 z b? 0? m>ws $84.2 93$ 28: a .0 ea no 25 E .3 _ in m»: 31:: 2:? g E o_ _o>a:w 3 285 B 6 2mm .5 .E :m Ev MVmS A: 78 935 m a: m _o>8w 9. “firm 3 .w mm .5 .E _ 0? firm 8.? NBmN la 03$: 3 6 :3 “E “B _ 3v arms all: :5 w .E .E 2 wEEm 36% m 28E m .m :8 .E .E _ Qm mi: 2.0 :< 5m mwac 3 2m 3 + 8 + :m m: 285 II :3 .E .3 c xo Em «*3. +5: 2 35¢ 3% use: a .0 ea : .E 2 a _ em mi :72 3% mg 8 a I 8 + :m mm 035.: b? B 6 :3 u .3 c b?» 3: firms 3.? mama g on _ 8 + a mm 2%.: a .0 em .5 Ba : .3 : :2 9m x»? 91m 3: $88 _ 03+ 8 + :m mm 2me 3 .w 3% .E :8 a .3 c D? Em ~50» ~32 m w .E .3 aw“: m 03:: a a 25 ea .5 .E _ S 53 2-: :< 3: § 85.3: wiofiwat ESE 0390:: 5:8: :28 =om:=2 8 8:80 8:036:00 583% 28:0 533 Eaugw 23on 362 58m 882.: 2:05 95:03 22388: :ouom “v 2an mp 2:9“. : Sm 685% 8:: fl 3m 6538 :81 Sm ”98an 8:5 . omN >985 we ‘uzdeq 7 a 5: “2:ng 8.5. 5.1 ficoEmfi :33 :1 5.83 8:: >m=m> :82... wfl< 6.38mi. cowo, 5< 6:0Em6 8:: _ N5< 65:8 :80 h :._< 55:8 8.5. a 0mm ...... com on.) C I/ , :. / o . omw .. A» , ... .- 2 x. . ./. \ »O , 00—. J. » . , .. .6 .25., on I O< : fid .\ ‘ ‘ [a O 90— cm 0 Myra—6‘ mozw __m .58: 6 885: m mm mEmEmm: 888 E8me 8O c0506.: EE QV mhzcm to “amok—wt .. S 9:9”. EE m - v ”95:3 3:? EE v - N 63:3 :39" _ L\‘\w .1 t! M, wHoom .oow .oom .oom Hof » I womp .m/x, .omF / ..omF M _, .xhwi \\\ _ .ooP u .ooF xxx .ooF \\\\ _ooP ',_ ma ‘qzdeq .om \. . .om / /u .om .xv/ +om ,. /. {/11 mD< 5< N3< E._< mzm Sta? .Eamu ho 8:05: m 8 _m>m5 we: Lcmobmn. 81 t 2:9”. .... Sow ...... fl _ ._ o? ._ ‘ ... w 2: .... .__ om Tli.llir. .1 Ho 2: om o «3 00? cm 0 5< Ger 5 .‘MN' _I__#,_ om - com .omw .2: .0m oo— om mum warm? .5qu .6 8:05: a mm. 9.8 $93.20 Ememn. F34. Loom .om we ‘qzdég 82 Apart from the grading in ALfl and ALf2, the sediments show no internal structure. The deposits at Alarka have a matrix supported fabric, in spite of the very high amount of coarse fragments. The coarse fragments are poorly sorted and contain pebbles, cobbles and boulders. The maximum clast size (a-axis) observed at the Alarka locations was more than one meter and the average clast size is about 13 cm (Table 5). The coarse fragments have a variety of shapes but a relatively large percentage (about 50 %) have disk or sub-disk shapes (Table 6). This dominance of flattened fragments probably relates to the metamorphic nature of the bedrock (Table 7). The foliation and fissility planes are loci of weakness which predispose the rocks to break up into diskoid fragments. Most of the coarse fragments (>70 %) are sub-angular or sub-rounded (Figure 18). 5.3.1.3 Hidden Valley site. As compared to the Alarka site, coarse fragments at the Hidden Valley site are less abundant and smaller (Tables 4 and 5). The average size is 6 to 8 cm and the maximum clast size sampled was in the cobble grade. Most clasts are sub-angular or sub—rounded; there are slightly more angular than rounded clasts (Figure 19). All Sphericity classes are represented (Table 6). The sediments of the Hidden Valley site are poorly sorted and matrix supported. The clasts are widely dispersed in the matrix and no grading was discernible. The lowest pit (HLf) is the only one at this site that shows clear stratification (Figure 19). It contains a 25 cm thick, clast supported, diamict. The coarse fragments“ in the diamict consist of about 80 % pebbles and 10 % cobbles (Dcm). The sand fraction of this unit has a grainsize distribution that is clearly different from that of the over and underlying units (Appendix 3). The overlying unit also has a Dmm facies but has a very low concentration of coarse fragments. Sediments at HLa and HUa consist of a material that is similar to the topmost unit of HLf (33% coarse fragments) but has a finer matrix (Dmm). Soil pit HUf has a zone rich in coarse fragments at 72 to 183 cm that coincides with soil horizons ZBtZ and 3Bt3 Table 5: Clast sizes, all sites 83 Soil pit Average a-axis Maximum cm observed, cm 15 105 13 1 15 AUf 14 140 E AUa 11 110 HLf unit 2 6 2] HLa 6 10 HMf 6 16 HUf unit 2 7 17 (upper part) HUf unit 2 6 16 (lower part) HUa 8 21 BLf 10 95 It BLa unit 3 95 BLa unit 1 105 II BUa unit 3 120 ll BUa unit 1 11 125 Table 6: Clast s hericities, all sites 84 * For definition see section 5.2.1 Soil pit Disk* Sub-disk* Spherical* Sub-prism* Prism* % % % % % AU] 29 21 13 13 23 ll ALf2 31 18 16 12 22 AUf 24 38 14 10 14 II AUa 24 29 10 20 16 “ HLf unit 2 14 25 24 17 20 II HMf 14 18 37 25 6 ll HUf unit 2 34 20 23 11 1 1 ll (upper part) HUf unit 2 16 22 18 34 10 fl (lower part) HUa 17 33 31 14 4 BLf 53 22 8 10 BLa unit 3 53 16 6 18 8 BLa unit 1 39 22 18 18 4 BUa unit 3 38 16 24 6 16 Bua unit 1 61 7 13 7 11 85 o o o n o 2 Tm 3V fl :5 «Dm— o w o o w 2 NN ow m :5 «Dm C v N N c N_ NN mm H HE: mam o v o v E 8 0 mo m “E: «am e N~ o NH m N v mo in o oN o o N E 0 mm «DE 9.8m .5305 2 o o o v NM N Nb N :5. “DZ 0.3m comma o o o o o o no N E5 me 3 o 8 E o N o mm 32$ 0 fl 2 ON C ~ ~ 9 am N “E: SE 0 mm o N S c N o «D< 0 it o N E o o N “Due. o N o v o 2 mm mV 812 o o o N E N_ m _ mm 512 “x. s as as s s s 88» as homuncuED 80823802 Hanging 5.8% 50> 320m 85.850 353:0 36:0 HE =om 3% EN 22865: 3.20 K 289 86 3 93mm BESS 36> D nonczoa 83mg 5398 Em; I mD< 5< NS< 5.4 o\o o nxvow. o\o 0v o\o o m o\oow o\ooor mmzw __m 6786 B mmocncsom 87 2 8am 0.. 8N mom | EEO H :5. on UN I o? o :50 N ES «UN EEO | I 8 1 I 02 mwo E | EEO I :m I m as 2 m in M < < momomm 5903 38am newton Eu 0 «Am SI 88 >o=m> $623 3 @093 3032 £85 82835: B 5:38 wofi 0328885 88 (Figure 20). The coarse fragments consist mainly of pebbles and cobbles while relatively few boulders are present. At the bottom of horizon 2Bt2 the coarse fragments are dominated by pebbles. The most appropriate facies code is Dmm. In Hle and HMa there is a gradual transition from highly weathered material at the top of the profile to saprolite, and (locally) hard rock, at the bottom of the pit. The saprolite consists of relatively fresh, brownish yellow gneiss in which the shape of crystals is still recognizable, and reddish, highly weathered, clay rich seams that cross the fresher material. Soil data presented in sections 6.3 and 6.4 confirm that the pedons at these two locations are developed entirely in saprolite. Consequently, no colluvial sediments are present at these locations and they will be excluded from further consideration in the discussion of sediments. 5.3.1.4 Bradley site. Sediments show more evidence for stratification at the Bradley site than at the other two sites. All contacts between the various sedimentary units are planar, horizontal and lack evidence of erosion. As in Hidden Valley, the lowest pit (BLf) contains a diamict layer that is clearly coarser than the overlying units (Figure 21). Unit 1 is matrix supported and poorly sorted (Dmm). The clasts are of pebble and cobble grade and most tend to be disk or sub-disk shaped (Tables 4 and 6). As in the other deposits that were studied, sub-rounded and sub— angular clasts dominate over other roundness classes (Figure 18). By analogy with the Clast poor units in Hidden Valley, unit 2 was also classified as Dmm. It contains only about 2 % coarse fragments. The unit at the bottom of pit BLa is >70 cm thick. It is a clast supported, poorly sorted diamict (Dcm) (Figure 21, Table 4, Appendix 2). The clasts have predominantly a disk and sub-disk shape and are sub rounded or sub angular (Table 6 and Figure 18). The contact with unit 2 is horizontal and appears to be planar. That overlying layer is almost 89 om oswm lllll Illll Illll Ill-I'll I'll] IIIIIIIIUIIIIIII I'll-Illl lIGIIIl-llll lllll III-nil]!!! lllll lllll lllll Ill-I'll III-Ill! Illll IIIIIII. Illlllllll Ill-III] ..IIIIIIIIIIIIIII Illll llllllllllll'll Illllll I'll-Ill momomm «DE “mm vzmm mamN Nkm 15m 71 cm thick matrix supported diamict was observed. Both the matrix and clasts are poorly sorted. Several clasts in this soil pit were in an upright position; imbrication was lacking. Most of the clasts are disk or sub-disk shaped and are sub-rounded to sub-angular (Table 6 and Figure 18). No internal structure was discernible. The contact with the suprajacent loam unit is horizontal and relatively clear. The 27 cm thick loam unit is completely free of coarse fragments. The sand subfractions are dominated by fine and very fine sand. There is no noticeable internal structure or grading and the contact with the overlying diamict was relatively sharp, planar and horizontal. This unit is analogous to unit 2 of BLa, but because no clasts are present it is assigned a Fl facies. The composition of the suprajacent diamict (unit 3) ranged from a relatively high amount (i 25 %) of diskoid pebbles and cobbles near the bottom to a few small, dispersed, more spherical, pebbles near the top (Dmg). The matrix of this unit is sandier and better sorted than the other units in the profile. The shape of the clasts tends to be more spherical and more rounded than in most other units (Table 6 and Figure 18). At one side of the soil pit a small lens of massive, clast supported coarse fragments replaced this unit (Dcm). Pedogenesis and bioturbation have masked the upper and lower contacts of the lens. Unit 4 92 is a matrix supported, reversely graded diamict (Dmi). Near the bottom of the unit the coarse fragments consist of about 10 % pebbles but near the top there are 12 - 20 % pebbles, cobbles and boulders. The unit is poorly sorted and does not possess any internal structure. It is 31 cm thick and extends over the entire length of the pit. Its boundary with the topmost unit is horizontal and appears to be planar. The top 49 cm of the profile consist of loam sediments with 2 - 3 % disk and spherical shaped pebbles. It is poorly sorted, homogeneous and lacks any sedimentary structures (Dmm). 5.3.2. Fabric 5.3.2.1 A-axis Schmidt nets of clast orientations (Appendixes 5 and 6) indicate that scatter exists in the orientation of the clast a-axes. However, the distribution is not uniform (test of uniformity, or = 1 %) and consistent trends exist in the clast orientation. The mean plunge of the a-axis is very gentle (< 23°) at the Alarka and Bradley sites (Table 8). The direction of the azimuth of the a-axis at these sites is approximately opposite to that of the slope (about 180° offset). The deviation from the opposite direction is small (<21°) for all Alarka and Bradley pits except for AUa where it is 45° (Table 8). The deviation is in an easterly direction, except for AUf where it is very small and to the west. This means that, in general, there is a weak, gentle upslope imbrication of the clasts. Only a minority (<23 % in Alarka, 10 % in Bradley) of clast long axes are oriented traversely to the slope of the surface. More detailed analyses of the fabric shape of the distributions (Appendix 5) reveals that the deposits at Alarka and Bradley have two subpopulations. All samples have one subpopulation that is concentrated around the mean a-axis. It is called hereafter "subpopl The second subpopulation ("subpop2") is scattered in AUa, BLa and BUa but appears to fall on a great circle in AU 1, ALf2 and AUf (Appendix 5). The absolute deviation of the 93 N.N.m 5508 wow .mcogcmow com ... movd mend :md o: .o Svmd cwvd S ~N2ON v N o mNN «Dm :od mvmd mmcd 2 _.o cemd mhmd 9 NE: 0 Nm 2 OMN 3m ammd ind m _ md afi .o Nde Svd 5N NfiNv- S» NMN N. o3 VD: 036 N_ #0 fed omNd chd vovd mN 3R2 M2 NE a QR aD< med Need Ommd mm_.o wOmd mmmd mN N\Nw_ N mm m RN .5< moo.“ oovd fad N26 NVNd 000.0 2 mtmf v 3 m NcN NR1? VmNN ommd End 526 mmNd onwd 2 3:2 MN mo w VNN C12 ...—05% omcsa 525$ Oman—m 5355“ Q. coufisoto o 2585 o 85%: “o :oumEoto o *M ...mQNmE *Nm: m5 *8 *Nm 1m 5:5 :ouaSoQ was“ :82 oomtsm mo Ono—m :a =om mama couflzmom :89 .853 620 we soufiaoco “w 033. 94 azimuth of the mean a—axis of subpopl, from the azimuth of the slope, is roughly 180° but it is consistently greater than the deViation of the mean of the whole population (Tables 9 and 10). The best fit great circle of subpop2 is oriented parallel to the slope in AU] but occupies an intermediate orientation, with respect to the slope of the surface, in ALf2 and AUf (Table 9). At the latter two locations the dip of the great circle is downslope. For ALfl and ALf2 it was determined whether clast size could have affected the orientation of the clasts, and thus to which subpopulation a clast belongs. In ALfl the difference in average clast size for the two subpopulations was only about 3 mm (2 %) and not statistically significant (two tailed t-test with on = 10 %). In ALf2 the difference was 1.8 cm, and although this difference was not statistically significant (two tailed t-test with or = 10 %), it does represent a difference of >11 %. Clast orientations at the Hidden Valley site (HUf) are very different from all other distributions. HUf does not have 2 subpopulations, and the a—axes are more concentrated (lower spherical variance) than in any other sample (Appendix 5). The plunge of the mean a-axis is steep (47°) and in the general direction of the slope. The plunge is larger than the gradient of the slope and thus effectively represents a downSIOpe dip. Therefore, there is a weak downslope imbrication at HUf. Clasts dipping upslope are conspicuously absent (Appendix 5). The deviation from the azimuth of the surface slope is relatively large (42°) and to the west. More clasts (27 %) were oriented traversely to the axis of the slope at this location than at any of the other two sites. In a relative sense, the first eigenvalue is small (O.4672) indicating a clustered fabric (Table 8, Figure 22). At the Hidden Valley and Bradley sites K values are small (K<1), which indicates that the fabric is girdle shaped. The extreme weakness of the fabric, however, makes these shapes less meaningful. 95 N.N.m .8508 08 $85.53 8m ... god 82 Bad mm‘ in NM: 1 a E :2 83 Sam memo cm 8 mm a New $12 __ 3.3 ME: 33 w mm «mm m am 5.... __ owcsa 5255“ owcam 5255“ __ 0.358% c o ...M Lakes 3ng 50 89.35 22633 £an 08:33 aoa a mom 55 £38-“ 620 no :osficoto “m a 25d. N.N.m .5508 now £555.35 5m ... QR; m _ co 5mm .5< whwd ommd 832 9 8— a mom “5?. 5.3.0 mom .5 NQwE Om E. w vmm $12 65% omen—m 5:85“ owes—W 5255s 0 £5255N o 0 LS: 95 50 8:225 85:25 8a-“ :82 oats... .6 2.2m i Em *Ammfimvfi fl :oaflawoonsm .moxwé 7.20 96 :ozflcocO ”< a 25¢. Ln(S1/82) 96 Logarithmic ratio plot of eigenvalues, selected sites 3 . 5 ‘ / ID 0'! \ N 1.5 O 0.5 1 1.5 2 2.5 3 3.5 4 Ln(82/83) Figure 22 97 The eigenvector analysis of the two subpopulations of AMI , ALf2 and AUf indicate that their separation on the basis of apparent fabric shape is justifiable. The fabric of the subpopulations is stronger (S 1/S2 or S2/S3 are larger) than the fabric of the total populations (Tables 9 and 10). The K value for subpopl is systematically higher than the K value of the total population, indicating that subpopl is more clustered. The K value for subpop2 is uniformly smaller than that of the whole population and thus points to a better defined girdle for that subpopulation. 5.3.2.2 A-b plane The orientation of the a-b planes of individual clasts is very irregular (Appendix 6). Only in BLa can a clear girdle fabric shape be observed. The orientation of the mean a-b plane, expressed as a best fit great circle, however, is remarkably consistent among the various deposits (Table 10). The plunge is always nearly horizontal, the largest divergence from the horizontal (observed in HUf) being 16°. Emma 5.4.1 General One of the most common features of the sediments is the predominance of diamicts. The shape of the clasts indicates that they were transported by colluvial rather than fluvial processes. The diamicts are interpreted as various types of debris flow deposits. The apparent tabular geometry of the beds of this study conforms to descriptions of many debris flow deposits (Bull, 1972; Nemec and Steel, 1984; Rust and Koster, 1984; Shultz, 1984; Wells, 1984). The weak tendency of the clast long axis to be oriented parallel to the flow and an otherwise disorganized clast fabric are also characteristic for debris flows 98 N.N.m cocoon 08 2:05:53 Sm ... mwo; Gmd m2; VNfio :Nd coed 3 2m a mNN 35 Rod mum: of .5 $06 hNNd Omnd ow mi : OMN 3m 0Nm© a: .o 03.0 N26 NNNd owmd Vb mVN N. o9 5D: cem.m 2 md 36; ofd 2 Nd NNod Vw 5mm a QB nD< Mde 8: Bed 306 C md vwmd ow M: m 0 RN mD< 32m vMNd owmg Em— .o m5 .o Sod cw NM: m NoN N5< 054m NoNd cNo; o3 .c 2N6 N56 mm wv w VNN Ex? owes—m 525.3 omcsa 5252 o .222 n-“ o *M *mmNmE *Nm: m5 *mm *Nm ... _ m %0 36¢ 25 85:5 go 255 E mom 288 2on E “mom 85E 5m use .50 :osficoto ”2 2an 99 (Lawson, 1979;1nnes, 1983; Mills, 1984; Nemec and Steel, 1984; Shultz, 1984; Wells, 1984; Wells and Harvey, 1987; Kochel, 1990). 5.4.2 Alarka site Deposits that have a Dmm lithofacies type (Gms sensu Miall, 1977), such as AUf and AUa, have generally been interpreted as being of debris flow origin (Williams and Guy, 1973; Miall, 1977; 1978; Kochel and Johnson, 1984; Rust and Koster, 1984; Shultz, 1984; Kochel, 1990). Lack of stratification, characteristic of the Alarka site, is often mentioned as a typical feature of debris flow deposits (Nemec and Steel, 1984; Rust and Koster, 1984; Kochel, 1990). It implies that the flows at ALfl and ALf2 arrived in one surge and not as a succession of pulses. Disorganized clast fabric has been noted in highly viscous flows (Nemec and Steel, 1984) and in fluid debris flows (Wells and Harvey, 1987). The lack of many traversely oriented clasts (Nemec and Steel, 1984) and the presence of large matrix supported clasts (Shultz, 1984) supports a plastic, viscous flow origin for the deposits of AUf and AUa. The upslope imbrication observed in a subpopulation of clasts in Alarka also is not uncommon for debris flows (Nemec and Steel, 1984). Normally graded diamicts, such as those of AU] and ALf2, have also been attributed to debris flow activity (Nemec and Steel, 1984; Shultz, 1984; Walton and Palmer, 1988). Shultz (1984) found that beds with a Dmg facies are thinner, finer grained and better organized than Dmm beds. In this study there was no difference in organization or grain size of the Dmg and Dmm facies. Because the bottom of the deposits was not reached, no conclusion can be drawn about a difference in thickness. The deposits are, however, at least twice as thick (>2 m vs. <1 m) as those studied by Shultz (1984). This suggests that the yield strength of the flows was low enough to allow settling of the clasts but it must have been sufficiently high to sustain the deposition of a thick deposit. 100 5.4.3 Hidden Valley site Most sediments at this site have a Dmm facies. As explained above, Dmm facies with disorganized fabric are generally attributed to plastic, viscous debris flows. The low amount and small size of the clasts may relate to the composition of the source material. It has been argued (Velbel, 1987) that the high weathering intensity of the Southern Blue Ridge produces source materials that do not contain sufficient coarse clasts to produce Clast-rich debris flow deposits. The predominance of this finer textured facies at the Hidden Valley site, and not at the other sites, may be due to the presence of less steep, and therefore more stable, slopes in Hidden Valley. Slope failures on stable slopes may be widely separated in time and hence, more time for weathering and breakdown of particles into smaller fragments would be available. This facies is similar to "facies 2" of Walton and Palmer (1988) and to "mud flows" described by Wells (1984). Considering that the grainsize of the transported material is the main difference between mud flows and debris flows, "mud flow" may be the most appropriate term for the Dmm facies at Hidden Valley. Only two beds at Hidden Valley have high amounts of clasts (Figures 19 and 20). One of them, in HLf, has a Dcm facies (Gm, sensu Miall, 1977). Clast supported fabrics are generally associated with fluvial activity or highly mobile, low strength debris flows (see section 5.2). According to Shultz (1984) this facies may comprise the basal sediment or bed load of a mass flow. Walton and Palmer (1988), however, argued that clast supported facies represent low strength flow. Its sharp upper boundary and the textural differences from the suprajacent unit show that this bed is not the basal part of a thicker deposit but a separate bed. Additionally, Nemec and Steel (1984) noted that sharp contacts mark intersurge stream flow or thin fluidal debris flow. This unit, therefore, is likely to be the deposit of a very fluidal debris flow surge or intersurge stream flow. 101 The other coarse fragment rich zone at Hidden Valley, in HUf, has a Dmm facies (Figure 20). The smaller clast size at the bottom of soil horizon 2Bt2 suggests that there could be a sedimentary boundary (lithologic discontinuity) between soil horizon 2Bt2 and 3Bt3. Other characteristics, especially the Sphericity and lithology of the clasts (Tables 7 and 8), seem to support this suggestion but do not allow a final conclusion to be made. If the apparent gradual nature of the change in Clast size is real, the smaller clasts could represent a minor debris flow surge between two larger ones. 5.4.4 Bradley site The relatively clear, non-erosional boundaries such as those observed at the Bradley site have often been attributed to debris flow origins (Gloppen and Steel, 1981; Nemec and Steel, 1984; Wells, 1984; Velbel, 1987; Kochel, 1990). Unit 1 in BLf is similar, though thinner, than the sediments in AUf and AUa (Figure 21). As argued above, the Dmm facies of unit I typically represents plastic, viscous debris flows. Unit 2 contains only about 2 % clasts and can be considered to be a mud flow deposit. The facies of this unit is similar to "facies 1B" of Walton and Palmer (1988). These authors attributed the nongraded nature of the facies to transformation of a turbulent flow into a plug flow and subsequent "freezing" of the material. Schultz (1984) interpreted this facies type as being the product of a plastic, laminar flow. As noted above, Dcm facies such as that of the bottom unit in BLa (Figure 21), have been associated with fluidal and high strength debris flows and streamflood. The poor sorting of the clasts suggests that, in this case, a streamflood origin is not likely. The sharp contact of the next unit, unit 2, with the overlying unit and its red hue (redder than adjacent units) could indicate that it was exposed at the surface for some time. Therefore it could represent fluvial or mud flow activity during the waning stage of a debris flow. Unit 3 has 102 a Dmm facies with few clasts, not unlike "facies 1B" described as lahars by Walton and Palmer (1988) (see above). The bottom unit in BUa has a Dmm facies. The relatively high number of large clasts in a vertical position indicate that this unit was deposited by a high strength flow. The facies of the overlying unit, on the other hand, suggests that it was deposited by a very fluidal mud flow or hyperconcentrated streamflow. Its resemblance to unit 2 in BLa indicates that is probably was deposited during the waning stage of a debris flow. Unit 3 has a Dmg facies. Its well sorted matrix and rounding of the clasts could indicate that some fluvial reworking took place. Fluvial activity is also implied by the presence of a lens shaped body of clast supported coarse fragments. Its lenticular shape and composition indicate that it is a channel fill. Inverse grading, such as in unit 4, has been interpreted as being the result of dispersive pressure (Nemec and Steel, 1984; Walton and Palmer, 1988). As in "facies 1A" of Walton and Palmer (1988), the inverse grading does not occur in the basal zone only but extends over the whole bed. This implies that the flow was sheared over its whole thickness shortly before deposition (Walton and Palmer, 1988) and that it had a relatively high clay content. The topmost unit is similar to the topmost unit in BLf and presents a Dmm facies. A poorly sorted matrix and some dispersed clasts make a debris flow origin, similar to that of unit 2 in BLf, the most plausible mode of formation for this unit. 5.4.5 Fabric The overall isotropic nature of the fabric of the deposits seems to indicate that the flows were highly viscous (Nemec and Steel, 1984) or turbulent (Lindsay, 1968; Owens, 1973). The existence of two possible subpopulations in the Alarka and Bradley sites, however, shows that other interpretations are possible. The two suprpulations could point to the presence of more than one deposit in the zones that were sampled. Although the 103 thicknesses of the described units were not unusual for debris flow deposits and although an effort was made during the fieldwork to identify lithologic discontinuities, the possibility that sedimentary boundaries were either too gradual to detect or masked by pedogenesis cannot be excluded. An alternative, and more likely, explanation is that subpopl (oriented parallel to the flow) represents the clasts that were located closest to the bottom of the flow, and thus closest to the zone of maximum shear, and that clasts of subpop2 were collected elsewhere (higher) in the deposit. This explanation is compatible with the viscous nature of the flows in which the clast orientations originated. Viscous flows are known to possess a basal shear zone that promotes alignment of clasts and an overriding zone of turbulence in which clast alignment is absent (Nemec and Steel, 1984; Shultz, 1984; Watson and Palmer, 1988). The orientation of subpop2 seems to support this alternative explanation. In the pits that are located on the apex of a deposit (AUa, BUa) and at an intermediate slope position (BLa) the a—axes of subpop2 are scattered. This suggests that the clasts were deposited by "freezing" of the flow. In foot positions (ALfl , ALf2, AUf) the clasts of subpop2 lay on a great circle and gently dip downslope. Following the reasoning of Innes (1983) (see section 5.1.6), settling near the foot of a deposit of clasts that initially do not have a preferred orientation, could lead to the positioning of the a-axis of the clasts on a great circle (when projected on a Schmidt net). If the samples were taken on the axis of the deposit the great circles would be perpendicular to the axis. However, the great circles are clearly not perpendicular to the central axis of the deposit (Table 9). The oblique orientation of the great circles with respect to the axis of the deposit can be due to the location of the soil pits on the side, and not the axis, of the deposits. This explanation has to be considered as an hypothesis and cannot be verified until more systematic measurements of debris flow fabric in different topographic positions are available. Because larger clasts tend to settle faster than small clasts, subpop2 should have larger clasts than subpopl if the above 104 hypothesis is correct. The comparison of clast sizes in the two subpopulations of ALfl and ALf2 is, however, inconclusive and does not permit me to reject or accept the hypothesis. The deviation of the mean azimuth of the clasts from the direction of flow is smaller in the debris flow deposits studied in this research than in some lahar facies (Mills,l984) and small glacial debris flows (Lawson, 1979; Dowdeswell and Sharp, 1986; Visser, 1989) (Tables 1 and 9). The deviation is largest in the clast supported facies studied by Mills (1984) but in the clast supported unit of this study the deviation is very small (6°). Mills does not explain his observation but considering that clast alignment is attributed to shear, which may be produced by Clast-Clast interactions (Lindsay, 1968; Nemec and Steel, 1984), it is not surprising that the clast supported unit of this study is closely aligned with the direction of flow. The deviation of the direction of the mean a-axis from the upslope direction is to the east (except in AUf). A similar deviation, consistently to the right of the downslope direction, was also observed by Mills (1983). Mills could not explain this observation but excluded bias in measurement. The reasoning of Innes (1983) could again provide an explanation. If the soil pits in Alarka and Bradley were not located on the longitudinal axis of the deposit but to one side, i.e. the east side, and settling occurs, one would expect to find a systematic deviation to the east. It would be coincidental, however, that all but one of the soil pits in Alarka and Bradley were located to‘the east of the axis of the deposit, especially when location on the central axis was sought during sampling. The dip of the clasts at HUf is downslope and much steeper than at the other locations (Table 8). The dip is more than two times as steep as the dips found in other debris flow deposits (Lawson, 1979; Mills, 1984). The absence of two subpopulations at HUf is compatible with the fluid nature of the flow that presumably deposited the clasts. Fluidal flows usually are turbulent and lack shear zones that produce clast alignment near the bottom (Nemec and Steel, 1984; Shultz, 1984; Watson and Palmer, 1988). Traversely 105 oriented clasts, present in relatively high amounts in this unit, are thought to move by rolling along the bed in fluid flows (Nemec and Steel, 1984; Kohlbeck et al., 1994). The strength of the fabric of the debris flows in this study is comparable to the strength of the fabric of lahars and small glacial debris flows (Lawson, 1979; Mills, 1984). These three types of deposits originate from very similar processes and all have very weak fabrics (Figure 23). Fabric strength, however, is not a foolproof method for the identification of debris-flow-like deposits because other processes, e. g. creep, can produce weak fabrics as well (Figure 23). 5.5 Qonclusions The sedimentary characteristics of the deposits show that they were most probably formed by debris flows and mud-flow-like processes. They consist of massive, unstratified and disorganized diamicts. The debris flow deposits, mainly located at the Alarka and Bradley sites, have high amounts of clasts (generally >50 %) while the mud-flow—like deposits at the Hidden Valley site contain very low amounts of clasts (<3 %). Matrix and clast supported facies were identified, indicating that framework support is not a diagnostic property of the studied deposits. The deposits at Alarka are the most homogeneous and have characteristics of highly viscous to intermediate strength debris flows. A relatively high yield strength for these deposits is consistent with the observation of a lobate topography in the Alarka site. In this respect these deposits resemble debris flow deposits emplaced on arid alluvial fans. At Hidden Valley, mud flow facies predominate but deposits of fluidal debris flows are also present. The paucity of clasts in the mud flows may reflect the composition of the source material and the relative geomorphic stability of the basin. At the Bradley site, deposits are more heterogeneous than elsewhere. Evidence for the stacking of materials from different events is most clearly expressed at this site. Materials deposited by viscous and fluidal 106 mm 2:5 r Bo: mtnmn _m_om_m ficoEmfi cmao 69.25% x595 852 6.5525 8:: Anmtoaazw 653 5:2 ”mango. cmao 53w 25 6633.0. 8:: _ mw\Nm v m.m m m.N N mé — m.o o . _ _ _ .5. 0 exam 2.0-MW oo 00 mEmEco._>cm 652688 $50 92 52m m5 .303 0:2 22:288.. mé m.N mm ZS/LS 390 65:3 :80 356 $5 66:3 3:: m.m m p FQNw 9m m 3 F b .r — P m.o m4 m.N m.m ZS/LS 107 debris flows, and by mud flows, are present. The absence of evidence for surface exposure, except in one unit in BLa, combined with the non-erosional contacts, indicates that the sequences at the Bradley site were deposited in a relatively continuous fashion. Apparently there were no major interruptions in the depositional episodes. The overall fabric of the deposits is isotropic. One subset of the clasts in viscous flows, probably those transported near the bottom of the flow, shows a tendency for a weak (gentle dip) upslope imbrication. The orientation of a second subset of clasts, those transported higher up in the flow, is influenced by their position with respect to the front of the flow. The fabric of only one unit formed by a fluidal debris flow was studied. It is very different from the fabric of the units produced by viscous flows. It is disorganized and has a steep (47°) imbrication downslope. The orientation of the long axis of the clasts is influenced by flow direction and the local slope. The fabric of the deposits, as expressed by the ratio of eigenvalues, is weak. It is comparable in strength to other debris flow deposits described in the literature. However, fabric strength is not a diagnostic property of debris flows because other types of colluvium can possess a weak fabric as well. Settling of clasts and bioturbation may change the original fabric of debris flow deposits. Fabric analyses of debris flow colluvium would, therefore, benefit from more information on the change in fabric over time. 108 6 SOILS 6.1 Literature review 6.1.1 Morphology Uncultivated soils on the Blue Ridge Front of North Carolina often have a complete O—A-E-B-C profile (Graham et al., 1990a). Some of these soils have developed entirely in saprolite or colluvium; most have developed in both materials. Nevertheless, Graham et al. (1990a) did not indicate the existence of lithologic discontinuities in their pedon descriptions. Argillic horizons had formed more quickly in fresh regolith upslope than in highly weathered colluvium on low slope positions (Graham et al., 1990a). Soils in the Joyce Kilmer Memorial Forest in western North Carolina are deep (>90cm) (Daniels et al., 1987a). The A horizons of the soils in this area area thicker and contain more organic matter on north-facing slopes than on south-facing slopes. On north aspects significant amounts of the organic matter are mixed deeply into the subsoil. Weak structural development in the subsoil is attributed by Daniels et al. (1987a) to the absence of wet - dry cycles. Surface horizons however, particularly those on north aspects, have better structure. Sometimes the original rock structure extends well up into the profile, but depth to bedrock is frequently >l.3 m. Relatively few coarse fragments are present in most of these soils, typically <10% by weight. Soils at higher elevations, on the other hand, often contain a dense gravelly subsurface layer (Feldman et al., 1991b). Literature reviewed by Feldman et al. (199 lb) led them to believe that the uniform morphology of high elevation soils of the Southern Blue Ridge has resulted from the dominance of climate in pedogenesis through its effect on slope instability, and that the influence of rock type is secondary. Norfleet and Smith (1989) concur with Feldman et al. (1991b) in that slope and climate are the main factors responsible for differences in soil 109 morphology, rather than parent material, vegetation or age. They reached this conclusion after studying Typic Hapludults on steep to very steep slopes in the Blue Ridge of South Carolina. 6.1.2 Physical and chemical properties. 6.1.2.1 Clay content Soils of the Blue Ridge Front that have at least partially developed in saprolite have the most distinct clay-enriched zones, while entirely colluvial soils have more gradual clay increases (Graham et al., 1990b). Fine/total clay ratios indicate that much of the clay in the Bt horizons of the saprolitic soils is due to illuviation. Clay contents decrease sharply in the saprolite itself (Cr horizon) (Graham et al.,1990b). Enrichment of clay in a B horizon has not taken place, however, in soils under virgin hardwood forest, probably because the climate was too wet (Daniels et al., 1987a). Whole soil clay contents of these soils are higher on north-facing slopes than on south-facing slopes (Daniels et al., 1987a). The accumulation of clay in a thin rind just above the contact of the R horizon was seen by the authors as evidence for illuviation. 6.1.2.2 Bulk density The bulk densities of A horizons of soils in the region are quite low (<1.0 g cm'3) due to high organic matter contents and porous structure (Daniels et al., 1987a; Norfleet and Smith, 1989; Graham et al., 1990b). In the Blue Ridge Front bulk densities of the B horizons are higher than those of underlying saprolite (1.6 g cm'3 as opposed to 1.4 g cm'3) (Graham et al., 1990b), but Daniels et al. (1987a) found that bulk densities generally remained below 1.4 g cm'3 through the porous B horizons, before increasing in the saprolite. 6.1.2.3 Base saturation 110 In the Blue Ridge Front the dominant exchangeable bases of soils are K and Mg, but low levels of Ca are also present (Graham et al., 1990b). Base saturation is low in the acid surface horizons (< 10 %) but increased in the B horizons of the soils (12 - 40 %). On lower slope positions where cations carried downslope by throughflow have accumulated, base saturation is particularly pronounced (i 40 %) (Graham et al., 1990b). In other areas of the Southern Appalachians, exchangeable bases are found to be very low (Norfleet and Smith, 1989; Daniels et al., 1987a). Potassium, directly released from feldspars, is relatively abundant (0.1 - 1 cmolC kg‘l) but Ca and Mg are virtually undetectable in most horizons (Daniels et al., 1987a). 6.1.2.4 Free iron content Parent material has a clear influence on the free (sodium citrate-bicarbonate- dithionite (CBD) extractable) Fe content of soils in the region. In the Blue Ridge Front, soils that are derived from gneiss have less free Fe than those that are derived from schist because the schist has a higher proportion of Fe-bearing minerals (Graham et al., 1990b). High elevation soils in micaceous parent material contain higher levels of CED-extractable Fe than those formed from more siliceous parent material (Feldman et al., 1991b). Soils in colluvium have a more uniform distribution of free Fe due to the colluvial homogenization of preweathered material (Graham et al., 1990b; Daniels et al., 1987b). Free Fe generally reaches a maximum in B horizons but in some pedons there is a decreasing trend with depth (Daniels et al., 1987b; Norfleet and Smith, 1989). Deviations from the expected bulge of free Fe in the B horizons could be attributed to small scale variations in parent material composition and depth of weathering (Norfleet and Smith, 1989). Daniels et al. (1987a) observed that soils on north-facing slopes are consistently higher in extractable Fe than those on opposing south-facing slopes. There also appeared to 111 be a trend of decreasing Fe with increasing elevation. However, the authors did not consider the effect of aspect or slope position on soil chemistry to be strong. 6.1.3 Mineralogy and weathering. 6.1.3.1 Biotite The mineralogy of the sand fraction of some soils in the Southern Blue Ridge of South Carolina is comprised of quartz, feldspar, sericite and muscovite (Norfleet and Smith, 1989). The biotite content increases with depth or there are indications (high amounts of hydroxy-interlayered-vermiculite (HIV)) that it has in the past. In soils of the Blue Ridge Front, biotite has transformed into an interstratified biotite/vermiculite phase (Graham et al., 1989a, b). 6.1.3.2 Chlorite Primary chlorite is present in soils of the Blue Ridge Front (Graham et al., 1989a; b). More chlorite is found in colluvial soils than in saprolitic soils. Comminution during transportation of sand sized primary chlorite may reduce this mineral to the clay fraction, where it persists unaltered (Graham et al., 1989b). The unaltered state can be explained by a lack of intensive weathering. This can account for the difference between colluvial and saprolitic soils but does not explain why there is a relatively large amount of fresh chlorite as compared to other primary mineral contents (Graham et al., 1989b). 6.1.3.3 Vermiculite and HIV Surface horizons of some soils in the region are dominated by vemticulite and hydroxy-interlayered minerals (Daniels et al., 1987b; Norfleet and Smith, 1989). The vermiculite and interlayered minerals decrease with depth, while mica increases (Daniels et al., 1987b). At other sites regularly interstratified mica-vermiculite (RMV) predominates in 112 surface horizons and HIV increases with depth (Feldman et al., 1991b). The authors associated RMV with podzolization processes. The absence of HIV in surface horizons reflects both the mobility of Al-organic complexes out of surface horizons of these mountain soils and the inability of hydroxy-Al interlayers to form in the presence of organic acids (Feldman et al., 1991b). HIV seems considerably less stable at higher elevations than at lower elevations. Colluvial soils usually contain more vermiculite and (HIV) than do saprolitic soils (Daniels et al., 1987b; Graham et al., 1989b). Soils studied by Norfleet and Smith (1989) also had an abundance of vermiculite. The authors did not explicitly state whether or not the parent materials were transported, but the paper leads the reader to believe that they were not. If this is true, their finding is different from that of Daniels et al. (1987b) and Graham et al. ( 1989a, b). The higher precipitation in the study area of Norfleet and Smith (1989) most likely caused a more intense weathering of the micas. 6.1.3.4 Kaolinite Kaolinite in the soils of the region was produced by different processes. In the Blue Ridge front most kaolinite was probably formed as a pseudomorphic alteration product of sand-sized biotite. Comminution later released it into the clay fraction (Graham et al., 1989a). In the Southern Blue Ridge of South Carolina kaolinite contents vary greatly and have no fixed relation with gibbsite contents (Norfleet and Smith, 1989). This could indicate that kaolinite could have been produced by dissolution of primary minerals and subsequent precipitation in a silica-rich soil. In other, more silica-poor, environments kaolinite could form from direct alteration of primary minerals (Norfleet and Smith, 1989). Kaolinitization of sand-sized biotite was considered to be the most likely transformation process by these authors. The kaolinite content in the clay fraction of high elevation soils is nearly constant with depth (Feldman et al., 1991b). Its presence in the non-clay fraction suggests again that it is the result of more than one process, i.e. feldspar alteration in 113 addition to desilication of 2:1 minerals higher in the profile. The low kaolinite content in soils of southwestern North Carolina could be explained by the fact that the soils seldom dry out, thereby limiting kaolinite precipitation (Daniels et al., 1987b). Daniels et al. (1987b) did not find a strong influence of aspect or slope position on soil mineralogy. 6.1.3.5 Gibbsite Gibbsite is most abundant in the clay fraction of C and Cr horizons of saprolitic soils and decreases in abundance toward the surface (Daniels et al., 1987b; Graham et al., 1989a; b; Norfleet and Smith, 1989; Feldman et al., 1991). Occasionally, it reaches its maximum in the Bt horizon and then decreases with depth. Norfleet and Smith (1989) concurred with Graham et al. (1989a, b) in attributing the formation of gibbsite to reprecipitation in lower horizons of Al released by weathering of aluminosilicate minerals. In colluvial soils gibbsite comprises a much lower proportion of the clay fraction and a decreasing trend toward the surface is less evident (Graham et al., 1989a; b). Possible reasons are a homogeneous distribution of previously weathered minerals in the colluvium and a lack of time. The total amount of gibbsite does not change with depth in some soils of the Southern Appalachians, because the clay content decreases with depth (Daniels et al., 1987b; N orfleet and Smith, 1989). This indicates that it is fairly stable in the environment of the South Carolina Blue Ridge (Norfleet and Smith, 1989). The interpretation of Daniels et al. (1987b) was that either gibbsite remains in place following direct conversion from feldspars and is diluted as the soil weathers or that additional gibbsite is forming near the surface. No silt- or sand-sized gibbsite was observed by Daniels et al. (1987b). The decreasing clay contents with depth supports the hypothesis of rapid weathering and desilication from the surface downward. The presence and abundance (approximately 20 % by weight) of gibbsite was noted in weathering profiles of the Coweeta Hydrologic Laboratory area (Campione et al., 1992). 114 According to the authors, the presence of gibbsite in the profiles emphasizes the importance of solution-reprecipitation of Al-rich phases in temperate climates. It suggests that the activity of silica was more critical than the regional climate in controlling regolith formation (Campione et al., 1992). These weathering profiles lack a clear vertical zonation. This may reflect variations in bulk composition of the parent rock and/or variations in the composition of ground water. Bryan (1994) on the other hand, who worked in the same area, found that kaolinite rather than gibbsite was the most abundant secondary clay mineral. Gibbsite was present in the saprolite (C horizons) and declined in abundance upward in the profile. Similar trends are reported by Graham et al. (1989b), Norfleet and Smith (1989) and Daniels et al. (1987b) (see above). Bryan (1994) suggested that the trend was caused by pH and permeability differences between the saprolite and solum. It was not clear why kaolinite would dominate over gibbsite. 6.1.3.6 Iron oxides Hematite and goethite have been observed to coat almandine grains in soils of the Blue Ridge Front (Graham et al., 1989b). Some of the Al released during weathering of the almandine precipitates as gibbsite, a minor component in the grain coatings. Graham et al. (1989b) suggested that the Fe released from etch pits may have locally exceeded the solubility product of ferrihydrite which then precipitates and is altered to hematite. Other parts of the crystal may release Fe more slowly, resulting in lower localized concentrations of Fe, thereby favoring the formation of goethite, which has a lower solubility product than ferrihydrite. Graham et al. (1989b) partly rejected the conclusions of Velbel (1984) that coatings form in saprolite weathering under inorganic circumstances while biochemical processes in the soil produce etch pits and removed earlier formed coatings. The occurrence and abundance of clay-sized hematite in these soils is strongly dependent on the parent rock mineralogy (Graham et al., 1989a, b). Soils derived from 1 15 almandine-bearing parent materials (schists) have higher hematite contents than those developed in material without almandine (gneisses). It has been argued that rock structure, texture and other compositional elements could possibly affect the abundance of hematite in these soils (Bryan, 1994). Colluvial soils have the most evenly distributed hematite, probably as the result of homogenization of preweathered material during colluvial transport (Graham et al., 1989a; b). Although goethite is also a weathering product of almandine, it is found in the clay fraction of all soils, whether or not they are derived from almandine bearing rocks. The additional Fe may have come from the alteration of biotite to kaolinite (Graham et al., 1989a; b). The sites at the Coweeta Hydrologic Laboratory studied by Bryan (1994) differed from those in the Blue Ridge Front (Graham et al., 1990a; b) in that they contained more goethite and less hematite. This can be explained by the smaller crystal size and faster weathering in the Blue Ridge Front but also by climatic and pedologic variations between the two sites (Bryan, 1994). Coweeta is cooler and wetter than the Blue Ridge Front, and thus has conditions that favor the formation of goethite rather than hematite. 6.1.4 Genesis. Graham et al. (1990a) described pedogenesis on the Blue Ridge Front explicitly as a function of regolith type. In the construction of their model of soil formation the authors relied heavily on an earlier model of pedogenesis by Rebertus and Buol (1985). This model can be summarized as follows. In fresh saprolite an initial flush of clays occurs when readily weatherable minerals alter to clays. This clay is illuviated and gradually produces an argillic horizon. The soils then enter a period with little illuviation during which some of the cutans are destroyed by pedoturbation. When biotite becomes extensively kaolinized a new illuviation period starts and the argillic horizon continues to develop. This is followed by another period of minimal illuviation and destruction of cutans. Kaolinitization of muscovite then starts a third period of clay illuviation. Erosion can, of course, interrupt this 116 orderly progression at any time and mix the soil material. At first, clay sized material is produced by physical disruption of weathered minerals during transportation. After deposition of the eroded material, illuviation starts again. Depending on residence time, degree of preweathering and initial clay content, a cambic or argillic horizon forms. During geomorphic remobilization events, clays become depleted and soil development is inhibited. Clay accumulation in the B horizon resumes but it is slower. Finally the material stabilizes. Weatherable minerals are depleted but, given sufficient time, an argillic horizon develops. All pedons in the study area of Graham et al. (1990) in the Blue Ridge Front of North Carolina could be fit into this model. In high—mountain soils of the Southern Blue Ridge, geomorphic processes have resulted in younger, less weathered soil materials overlying older, more chemically weathered soils (Feldman et al., 1991b). Formation of clay minerals from primary aluminosilicates, mixed throughout the soil, was considered to be controlled by biotic and climatic factors by the authors. In most surface horizons the genesis of secondary clays was primarily the result of mica transformation in the following sequence: mica (biotite)--->hydrobiotite--->RMV-—->vermiculite--->smectite. The soils were thought to be undergoing the early stages of podzolization and therefore are very different from adjacent low elevation soils (Feldman et al., 1991b). They are typically associated with refugia of spruce-Fraser fir forests. Elevation also affects the genesis of soils in the South Carolina Blue Ridge through its influence on precipitation (Norfleet and Smith, 1989). The most important factor affecting the removal of silica from these soils is rainfall. High precipitation results in more gibbsite and less kaolinite in the clay fraction (Norfleet and Smith, 1989; Daniels et al., 1987b). 1 17 6.1.5 Classification The deepest soils of the Blue Ridge Front have formed where colluvium has accumulated on the footslopes. These soils are, however, weakly developed and classified as Typic Dystrochrepts or very weakly expressed Hapludults (Graham et al., 1990a). On the slopes and the summits in the Blue Ridge Front the soils are thin and classify as Hapludults. They belong to the Typic group in all landscape positions except on the backSIOpe where they may be Ruptic, Lithic or Entic. Typic Hapludults generally have developed in saprolite parent material (Graham et al., 1990a). The soils with the reddest and most clay—rich Bt horizons are Typic Hapludults on backslopes. At high elevations of the Southern Blue Ridge most soils classify as Typic Haplumbrepts (Feldman et al., 1991b). Some of these soils show the development of a weak, spodic-like horizon. Most soils in the Joyce Kilmer Memorial Forest in western North Carolina contain a cambic horizon and classify as Typic Haplumbrepts on north-facing slopes and as Umbric Dystrochrepts on south-facing slopes (Daniels et al., 1987a). Yellowish-red Typic Hapludults are observed by the authors over Fe-rich phyllites. These Typic Hapludults are a major component of the soils on south-facing slopes below 800 m. Norfleet and Smith (1989) and Daniels et al. (1987b) encountered a soil classification problem in their study areas. Mineralogically, most of the soils studied by these authors could be classified in the oxidic class. This class was intended to group intensively weathered soils together. But, some of the soils also classified as Inceptisols, genetically young soils, or had considerable amounts of easily weatherable minerals. In both cases this apparent contradiction was seen as the result of a flaw in the Soil Taxonomy (Norfleet and Smith 1989, Daniels et al. 1987b). None of the authors checked if there was a correlation between parent material (i. e. saprolite or colluvium) and how well a soil fit into a given mineralogy class. 118 Ms 6.2.1 Data collection As discussed in section 5.2. 1, pits were located on the foot and/or the apex of the deposits. At each location the soil was described and sampled in accordance with Soil Survey Laboratory Staff (1992) and Soil Survey Division Staff (1993) procedures. Morphological properties described for each horizon included horizon boundaries, texture, structure, redoxymorphic features, cutans and coarse fragment content. Estimates of the coarse fragment content were made in volume percent. Duplicate bulk density samples were taken from the horizons with the core method (Soil Survey Laboratory Staff, 1992). Geomorphic features such as slope, aspect, slope geometry and erosion were also recorded. 6.2.2 Data analysis Air dried soil samples were manually crushed with a mortar and pestle. Particle size analysis was performed by dry sieving for the sand and fine gravel (8 - 2 mm) fractions and by the pipette method for the clays. Procedure 3A1 of Soil Survey Laboratory Staff (1992) was followed but without filtering after organic matter removal. Bulk density of the soil samples was obtained by subtracting the weight of the >2 mm particles from the oven dry weight of the sample, and by dividing the difference by the volume of the cores (procedure 4A3 of Soil Survey Laboratory Staff (1992)). Bulk density was also determined for rock fragments collected on the slopes above the deposits. Their volume was determined by immersing them in water in a volumetric flask. Free iron was extracted with the sodium citrate-bicarbonate-dithionite (CBD) method (Mehra and Jackson, 1960) (see section 10.2.2 for details). 119 For clay mineral identification X-ray diffractometry (XRD) was employed. A subsample large enough to yield about 4 g of clay was separated from every horizon-based sample. The appropriate size of the subsample was estimated on the basis of the particle size data. Organic matter was oxidized with H202 at approximately 60° C (Anderson, 1963). The procedure was stopped when frothing disappeared completely. Iron oxides were removed with sodium CBD (Mehra and Jackson, 1960) (see section 10.2.2 for details). The sample was then added to a cylinder, deionized water was added to make the volume 1 L and the mixture was shaken vigorously for one minute. The time required for particles > 2 mt to settle to a certain depth was calculated. After that period of time had elapsed, the clays were siphoned off, the cylinder was filled with water again and the sedimentation and siphoning procedure was repeated. This procedure was repeated until most clays had been removed from the mixture. The clays were then transferred to centrifuge tubes for Mg saturation as described in Whittig et al. (1987). Slides were prepared by adding as much of a concentrated mixture of saturated clays and deionized water to a slide as could be held by surface tension. X-ray analyses were performed on a Philips XRG 3100 scanning diffractometer with Cu Ka radiation. The clay mounts were scanned in 2 sec steps of 005° over a 20 range of 2 to 32°. The relative abundance of the clay minerals was determined with a method based on Rieck et al. (1979) and Griffin and Ingram (1955). First the ratios of the intensity of the 7.15 A, 10 A and 14 A XRD peaks to the intensity of the highest peak were calculated. (For a discussion of the minerals that correspond to these (001) peaks the reader is referred to section 7.2.3). The ratio was then assigned a code of 1 if it was <33 %, 2 if it‘was >33 % and <66 %, and 3 if it was >66 %. This code reflects the relative abundance of the minerals in each sample. Rieck et al. (1979) argued that when ratios of intensities are used, and when the intensities are obtained from the same diffractograms, semi-quantitative comparisons between samples can be made. Therefore, while not a quantitative measure for the mineralogical composition of the samples, the code calculated here permits 120 semi-quantitative comparisons between samples to be made. Markewitch and Pavich (1991) recently employed a similar method, based on peak height ratios, to study relative abundances of clay minerals. 6.3 Results 6.3.1 Soil morphology and classification 6.3.1.1 Alarka site The soils at the Alarka site are very deep. Sola are more than 187 cm thick in three of the four pedons (Table 4). The C horizon was only encountered in AUa, at 161 cm. An A horizon rich in organic matter is present in all pedons, but only on the lowest deposits of the sequence (pedons ALfl and ALf2) is it underlain by an B horizon. (For the nomenclature of pedons the reader is referred to table 2). ALf2 has a 30 cm thick argillic Bt horizon that has a few distinct clay skins. Other B horizons at the Alarka site also have a few distinct clayskins, but they do not classify as argillic horizons because of the low clay enrichment. All B horizons at this site have a hue of 7.5YR (Table 4) and usually have a soil structure with a medium grade. The non-argillic B horizons have an extremely flaggy or gravelly fine sandy loam texture while the argillic Bt horizon has an extremely gravelly sandy clay loam texture. Surficial horizons (A, AE, E, BA) usually have a loam texture. The pedons at ALfl, ALf2 and AUa classify as loamy-skeletal, mixed, mesic Umbric Dystrochrepts. Because of the presence of an argillic horizon in ALf2 the pedon classifies as a loamy-skeletal, mixed, mesic Typic Hapludalf. 121 6.3.1.2 Hidden Valley site The pedons on the lowest deposit at Hidden Valley have relatively thin sola (Table 4). At HLf the C horizon was encountered at 62 cm and at HLa the solum was only 111 cm thick. At the other four pedons the upper limit of the C horizon was encountered at >211 cm. HLf and HLa, and one pedon in Bradley (BLf), are the only soils with thick, gleyed horizons. The gleying indicates that the water table is relatively shallow during a major part of the year and probably impedes a further development of the soils. HMa is the only pedon of the Hidden Valley site in which an B horizon was observed. It also is the only pedon that contains a buried profile and the only one in which a Cr horizon could be observed. In all Hidden Valley pedons except HLf, Bt horizons with cutans and a soil structure with medium grade were observed. Most of these Bt horizons classify as argillic. In the highest pedon, HU a, only the deepest B subhorizon described was argillic. The argillic horizons, except the one at HUa, have a clay texture. The B horizons have hues of 5YR or 2.5YR, redder than the B horizons of the soils at the other two sites. The texture of the soils in Hidden Valley is also distinctly different from that at the other two sites (Table 4). The pedons in Hidden Valley, except for the lowest one (HLf), are dominated by clay loam and clay textures. The lowest pedon has a loamy solum and sandy loam C horizons, the lowest of which is extremely gravelly. This is similar to the textures of horizons at the two other sites which are dominated by loams and fine sandy loams. Taxonomically the pedons on the highest and lowest deposits classify as loamy, mixed, mesic Typic Hapludults. The pedons HMf and HMa are clayey, mixed, mesic Typic Hapludults. 122 6.3.1.3 Bradley site The lowest pedon in Bradley is similar to its counterpart in Hidden Valley, in that it is shallow (depth C horizon = 59 cm) and has a gleyed horizon, indicating a high water table. The depth to the C horizon in the other two Bradley pedons is >187 cm. An E horizon was only observed at BLa. Bw horizons are described without cutans and with a weakly developed structure. Consequently, no argillic horizons were identified at Bradley. The hue of the B horizons is 7.5YR (Table 4). Superficial horizons generally have a loamy texture while most of the subsurface horizons are coarser textured. The lowest and highest pedons classify as loamy, mixed, mesic Typic Dystrochrepts while HLa is a loamy-skeletal, mixed, mesic Umbric Dystrochrept. 6.3.2 Physical and chemical properties 6.3.2.1 Alarka site Clay enrichment in the B horizons of the soils in Alarka is minimal (Figure 24). Only ALf2 has an argillic horizon. The BC horizons of pedon AUa, which is at the apex of one deposit, contain less clay than the overlying and underlying horizons (Figure 24). The clay content of the BC horizon is comparable to the clay content of the BC horizons of AUf, which is located on the foot of the same deposit, but the content is considerably less than in the horizons of AUf that are at the same depth. The bulk density of AUf is consistently lower than that of AUa (Figure 25). Other properties, such as percentage clayfree silt (Figure 26), do not show any systematic difference between foot and apex of a deposit nor do they show a consistent depth trend. The most striking feature of the soils at the Alarka site is the high amount of coarse fragments (Table 4). Pedons ALfl and ALf2 have horizons with more than 80 % coarse fragments. The highest deposit at Alarka (pedons AUf and AUa) has about 50 % coarse 123 8 9%: 0v.0 m3< .o.. - 5<- - o -- .4 0mm o... .. 00N . .0 Lo .. of 9...... .. 2: ....0 -.-o 0.. - X - f0m. t. 1.0.. .-..-- .. T1 o . O . O 0N0 3.0 00.0 050 0 we ‘urdeg ovd N3< lull Cr_< llDIIl .. 0mm .. 00m -. om? \ .... OOF I -r .. om a a u I.._ L. O omd ON.O o_..o 00.0 m-Eo 0 we ‘urdeg $5 $82 .5qu 6 5082 m mm 39: >90 124 mm 2:9“. mD< IT 5< lull .. 0mm com of J./ . on: H om . . o 2; om; 8; 8.0 8.0 m-Eo 0 mo ‘urdeg Nhr—< ll; Tl Cr_< llllll 0m; 00; 00.0 m-Eo 0 9% 52m? .5098 .6 5:82 m mm 36ch xSm mm 93mm 125 fizzzt--- 3: llloll :1 lol. 00 cm 0m ON 0— o\o .. 0mm .. oom .. om_. LL13 ‘qrdeg .. oow uro ‘qrdeq 5< - <3 2.2+ .-omm .. oom ..omw ..ooP ..0m 0 om ON or o\o 3003 3893 .508 B 5:95.. m mm Em $9390 126 fragments. The trends with depth of the total amount of coarse fragments are almost completely reflected in the fine gravel (8 - 2 mm) fractions (Appendix 3). In ALfl there is a sudden increase in the fine gravel below 50 cm. The depth profiles of the free iron mass in AU] and ALf2 are very similar (Figure 27). The maximum amount of free iron is reached in the Bt horizons at about 80 cm depth. The depth profiles of free iron mass for AUf and AUa are also similar to each other but different from those of ALfl and ALf2. In AUf and AUa there is a steep increase in the amount of free iron just under the surface, with the maximum being reached at less than 60 cm. The steep increase immediately below the surface suggests that erosion may have removed eluvial horizons on the higher deposit at the Alarka site. The results of the particle size fractionation show that there is large variation in the fine gravel and sand fractions between adjacent horizons, especially in ALfl and ALf2 (Appendix 3). 6.3.2.2 Hidden Valley site In Hidden Valley, pedons HMf and HMa have the highest clay enrichment in the Bt horizon (Figure 28). The argillic horizons of these pedons that have developed in saprolite have from 18 % to 25 % more clay than the overlying eluvial horizons. The Cr horizon of HMa has considerably less clay than the overlying horizons. On the lowest and the highest deposits (HLf, HLa and HUf, HUa respectively), clay enrichment in the Bt horizons ranges from 2 % to 16 %. In pedons Hmf and Hua an increase in the clay content can be observed near the bottom of the soil pit. This increase produces a second, deep, maximum of the clay content. No systematic difference in the clay content is present between the pedons located on the foot and the apex of the deposits. On the lowest deposit the pedon on the foot has less clay than the pedon on the apex, while on the highest deposit the pedon on the foot has more clay than the one on the apex. The trend with depth of the clay mass in the two R 9:9... 127 m3< illDIrl 5< [ill 3.? IDIII 2.2 IIIIII 0 0mm 4 0mm 00m .. 00m . of. Am .. on? ..m 3 t 00F w 00— 0m j 0m, . O T a 0 0 No.0 30.0 5.0 000.0 0 30.0 0.0 wood 0 020 0 020 0 uro ‘qrdeq 96 $82 .503 B cozoca m mm mme :0: moi 128 mm 2:9“. SI Ito: 5: rill 0. 0mm ~6/ a CON . fl 0 / A / OMF , a / , o. o 02 .9 \ .9 \ , 0 .0m a . x. d o 00... omd 00.0 m-Eo 0 231171021117: .03 \o \ \ ., Q .03. X 1 58 m x m p «U. Q \ 03 o .. fl w a a ..// /o .om /Q my» . x9 o 3.0 00.0 0-080 BI IrlDll SI ll-III .. 0mm 00m I \ S? m. d .m. \ .. 00* m - .om . 0 00; 0m.0 00.0 m-Eo 0 mzw >m=m> $22: .508 B 5:05: m mm 32: 220 129 pedons on the lowest deposit is very similar and shows a well pronounced clay maximum (Figure 28). This trend is typical of mature soil profiles. The absence of eluvial horizons near the surface may point to their removal by erosion. At HMa deposition occurred and a more complete profile, with an eluvial horizon, became buried (Table 4, Figure 28). Bulk density values show a difference between the foot and apex of deposits but the difference is not consistent (Figure 29). As in Alarka, the pedon on the foot of the lowest deposit (HLf) has a lower bulk density than the pedon on the apex, but the situation is reverse for the highest deposit in Hidden Valley. In pedons HMf and HMa minimum bulk density values are not observed in the A horizon, as was the case in all other pedons studied, but in the AB horizon. The pedons in Hidden Valley have considerably less coarse fragments than those at Alarka and Bradley (Table 4). Only HLf has a horizon (2Cg) with very many coarse fragments (90 %) but the horizon is only 25 cm thick. The free iron mass distribution on the lowest deposit is very irregular (Figure 30). The maximum free iron in the other pedons at Hidden Valley (HMf, HMa, HUf, HUa) is reached in the Bt horizon. The higher in the landscape the pedon is located, the deeper in the soil this maximum is found. As at the Alarka site, many large changes occur with depth in the amount of fine gravel and sand at HLf, HLa, HUf and HUa (Appendix 3). At BUa and Br4 there are only very gradual variations in the texture of adjacent horizons. The only large change in textural composition in these pedons can be observed between the AB and Ab horizons of HMa. 6.3.2.3 Bradley site As evidenced by the absence of argillic horizons, clay translocation is low at this site. The highest pedon, BUa, has two Bw horizons (2Bw2 and 3Bw3) that have less clay than the over and underlying horizons (Figure 31). At the Bradley site clay enrichment 130 mm 9%: 8: tier: 5: ......II .. omm 0 Q, .. 00N , a . mh .omr.m fl 1 3 .u. 7, .. of m 5* \ ..v W/ 0m «6 o oo.m 00; 00.0 m-Eo 0 95 36> .22: 1:01:22: 11.1.. 0 .. 0mm NV 0_ 5 com o a W om_. 4M0 00F &, WW .. om l o oo.N 004 00.0 m-Eo 0 31-101: :Irrsr: W .. 0mm , o , .. 00m _ O _Q _ / .. omw .m-flu A W m. .0 r r. O 00 W .. om o O . o oo.N ooé 00.0 020 0 :82: .508 .6 5:82 m mm. bacon xSm 131 8 $sz mDI liDll SI [Illi— mod we. 0 N00 020 0 .. omm . com of m ..0 .m 0 03 w .. cm 0 o _‘m_>_I |D| :21 Ill \ mod .0. N00 020 0 .% 0mm . CON o 3: |mTl ....I 1.1; m0.0 050 m .omm mzw >m=m> c002: .5005 .6 cozoce m mm mme :0: moi 132 Clay mass as a function of depth, Bradley site Depth, cm 0.00 4o ~~ 60 r— 100 —~ 120 “- 140 —~ 160 “r 180 i + BLf —D_ BLa —’_ BUa Figure 31 133 decreases with increasing elevation of the pedon. For bulk density values, a clear trend with depth could not be identified. The amount of coarse fragments in the soils at Bradley is more variable than at the other two sites (Table 4). The highest amounts (>80 %) are reached in the two lowest pedons. Free iron mass has similar depth profiles in all three pedons (Figure 32). The maximum is reached in the first Bw horizon at about 40 cm depth. The highest pedon has a secondary maximum at the bottom of the described soil profile. At this site, as well as in most pedons of the other two sites (see above), a clear eluvial horizon is absent. This can point to removal of surficial horizons by erosion. Texturally, the pedons in Bradley are more heterogeneous than those in the other two sites (Appendix 3). Large changes in the fine gravel and sand subfractions occur with depth throughout the pedons. This reflects the sedimentological complexity of the parent material. 6.3.3 Clay mineralogy 6.3.3.1 General Kaolinite was recognized by its 7.15 A and 3.57 A XRD peaks and mica was identified by its first order 10 A peak. The 14 A minerals theoretically include chlorite, smectite, vermiculite and hydroxy-interlayered minerals. Identification of each of these species is impossible with the data available but some can be excluded from the list. Chlorite, which is distinguishable from kaolinite by its second order 7.08 A peak and third order 3.54 A peak was not found in the samples. The presence of smectite cannot be excluded on the basis of the data but its occurrence is unlikely in the high intensity weathering environment of the Blue Ridge and in these well drained soils (Borchardt, 134 Free iron mass as a function of depth, Bradley site Depth, cm O 0.005 0.01 0.015 0.02 0.025 4o —4~ 60 i 100 ~— 120 —~ 140 —- 160 —~ 180 .. +BL1‘ +BLa "“""'BUa Figure 32 135 1989). The 14 A minerals present are, therefore, most likely to be vermiculite and hydroxy- interlayered minerals. In general terms, the clay mineralogy varies little within each site but presents clear differences between sites (Table 11). The Alarka site contains more 14A minerals and less kaolinite than the other sites. The Hidden Valley site contains almost no mica, few 14 A minerals but high amounts of kaolinite. The Bradley site has, proportionally, as much mica as the Alarka site, but it has more kaolinite. 6.3.3.2 Alarka site All pedons at the Alarka site contain relatively little mica (Table 11). Kaolinite is present in small to intermediate amounts at AU] and ALf2 and in intermediate to high amounts at AUf and AUa. The relative abundance of 14 A minerals is variable. The BC and C horizons have small to intermediate amounts of 14 A minerals, while higher amounts usually occur in surficial horizons (A, AE, E, BA) (Appendix 4). 6.3.3.3 Hidden Valley site The clay spectrum at the Hidden Valley site is dominated by kaolinite (Table 1 l). Kaolinite is abundant in every horizon of this site. Mica is absent in most horizons. Only pedon HUf has low amounts of mica (Appendix 4). Some 14 A minerals are present in proportionally low amounts in almost all horizons. They are absent from one B horizon in HMa and HUa and have an intermediate abundance in the A horizons of HLf and HUa. 6.3.3.4 Bradley site Kaolinite is also abundant in every horizon of the Bradley site (Table 11). Mica has a low abundance in every horizon (Appendix 4). The relative abundance of the 14 A minerals is higher than at the Hidden Valley site but lower than at the Alarka site. It is intermediate in 136 the A, and E, horizons of BLf and BLa and low in all other horizons of these two pedons. At BUa, 14 A minerals are intermediate in abundance in all horizons. don, all sites*. 14 A minerals Mica Kaolinite 2.0 1.0 1.8 2.3 1.0 2.0 2.0 1.0 2.6 AUa 2.1 1.0 2.5 Mean of site 2.1 1.0 2.2 HLf 1.0 0.3 3.0 Hla 1.0 0.3 2.9 Hle 1.0 0.1 3.0 HMa 1.0 0.3 3.0 HUf 1.0 0.8 2.9 HUa 1.2 0.2 3.0 Mean of site 1.0 0.3 3.0 BLf 1.3 1.0 3.0 BLa 1.3 1.0 3.0 BUa 1.9 1.0 3.0 II Mean of site 1.5 1.0 3.0 * Values in table are pedon averages. Increasing number indicates increasing relative abundance. See section 6.2.2 for definition of codes. 6.4 Discussion 6.4.] Saprolite In HMa a Cr horizon occurred near the bottom of the pit. There were no indications in the particle size distributions for the existence of a lithologic discontinuity higher up in the profile. In HMf a Cr horizon was not identified but the morphology of this pedon is 137 very similar to that of HMa. The B/C horizon of HMf is similar to the C/B horizon of HMa. These horizons consist of highly weathered, reddish, clay-rich seams surrounding angular, brownish yellow fragments of rock that are little weathered. These observations suggest that pedons HMf and HMa have developed entirely in saprolite. Although bulk density does not gradually increase with depth in the saprolite, there is a gradual transition from the bottom of the solum to the bedrock. The bulk density increases from about 1.36 g cm"3 in the C/B and B/C horizons to 1.41 g cm'3 in the BC horizon and to 1.45 g cm'3 in the Cr horizon. Fresh bedrock samples from the Hidden Valley site, which are likely to be similar to the parent material of the saprolite, yielded bulk densities of about 2.65 g cm‘3. The difference in bulk densities between the bedrock and the various transitional horizons represents the loss of mass that accompanies saprolite formation. Assuming isovolumetric weathering the mass losses are 48 % for C/B and B/C horizons, 46 % for BC horizons and 44 % for Cr horizons. This loss in mass is similar to that reported for the Piedmont by Cleaves (1974), for the Blue Ridge Front by Graham et al. (1990a), and for the Blue Ridge of Virginia by Stolt er al. (1992). 6.4.2 Lithologic discontinuities Particle size data for most colluvial soils show that many large variations in the abundance of the sand subfractions exist (Appendix 3). This may be interpreted as evidence for the existence of lithologic discontinuities. At the Alarka site and at [-11.3 and HUa, however, no evidence for discontinuities was observed in the field. There was no evidence, such as buried paleosols, distinct differences in coarse fragment content, or differences in sediment characteristics, for the presence of lithologic boundaries. Clay mineralogy data support the lack of lithologic discontinuities at the Alarka site and at HLa and HUa. The clay mineral spectrum changes with depth (section 6.4.4) but the changes are not closely associated with changes in particle size. The irregular trend with depth of other soil 138 properties, such as clayfree silt content (Figure 26), are not related to clearly distinct variations in particle size distribution either. The apparent discontinuities observed in the particle size distributions at the Alarka site and at HLa and HUa are, therefore, interpreted to be the result of internal variations within the deposit of a single flow, rather than being the result of different deposits stacked on top of each other. Consequently, the numeral prefixes assigned to horizons of the Alarka site and HLa and HU a represent differences in parent material from a pedogenic point of view but they do not necessarily indicate the existence of different depositional units. The internal variations in the deposits can be due to different shear zones in the flows. As shown in section 5.3, the heterogeneity of the parent material at the Bradley site is due to the presence of different sedimentary units. Graham et al. (1990b) and Daniels et al. (1987b) found a more uniform distribution of free iron in colluvium than in saprolite. In the present study, however, no clear difference exists between the trends of free iron in the saprolite and the debris flow deposits. The homogenization of preweathered material during colluvial transport, that Graham et al. (1990b) and Daniels et al. (1987b) use to explain their observations, did not take place in the debris flows. This is confirmed by the particle size data which indicate that, within one flow, vertical differentiation rather than homogenization occurred. 6.4.3 Clay content The soils that have developed in the saprolite have higher clay contents and experienced more clay enrichment than the pedons that have developed in debris flow colluvium (all other pedons studied). In the Cr horizon of HMa, the clay content decreased considerably (from 30 to 1 1 %). Similar observations were made by Graham et al. (1990a) for colluvium of soil creep and slope wash. In the saprolitic soils, changes of free-iron mass with depth closely follow the depth trend of clay mass (correlation coefficient = 0.91). Because free iron tends to move with fine clay in well drained soils, like those 139 studied here (Eswaran and Sys 1970, Soileau and McCracken 1967), and because fine clay is indicative of illuviation, the free iron trend suggests that the clay enrichment in HMf and HMa is primarily due to illuviation. The difference in clay content between the pedons in saprolite and colluvium is related to the type and age of the parent material. As will be shown below (part III), the soils in saprolite appear to be older than those in colluvium, and thus more time has been available for weathering, clay production and clay translocation. Other factors that influence clay formation and translocation, such as climate, vegetation, topography and drainage are similar at the three sites and probably did not play a major role. The soils on the highest deposit in Hidden Valley appear to be similar in age to the saprolitic soils (part III) but have thicker sola. The thick sola reflect clay enrichment over a wide zone and result in a high total clay mass when the amount of clay is integrated for the whole solum (see also figure 35). This suggests that, in saprolitic soils, relatively thin but very clay-rich B horizons often develop, while in colluvial soils thicker, but less clay-rich B horizons form. The highest pedons at the Hidden Valley and Bradley sites (HUa and BUa respectively), and to some extent the lower saprolitic soil (HMf), have a depth trend with two local clay maxima (Figures 28 and 31). One usually expects to find an increase in the clay content in the upper B horizon. In these two pedons, however, the clay content reaches a minimum in the intermediate B horizons and increases both upward and downward from there. At Alarka, C horizons were encountered in the highest soil pit (AUa) and a second clay maximum was not observed. Bisequal soils, with two clay maxima, have been observed elsewhere (Birkeland 1984, Kesel and Spicer 1985, Farming and Farming 1989, Gamble 1991). The two maxima have been linked to different parent materials (Kesel and Spicer 1985, Gamble 1991). As mentioned in section 5.3, lithologic discontinuities are clearly present at BUa. Sedimentary unit 2 (horizon 4Bw4, Figure 21), which is part of the second, deep, clay maximum, contains less coarse fragments, less clayfree very coarse, coarse and medium 140 sand, and more clayfree silt than the overlying unit (Appendix 3). It therefore seems reasonable to assume that textural differences of the parent materials, i.e. more clay in unit 2, determined at least in part the origin of the second, deep, clay maximum at BUa. Consequently, the first, shallow maximum (horizon Bw l) is considered to be most characteristic of pedogenic development at this locality. In Typic Haplorthox that had developed on alluvial fans, Kesel and Spicer (1985) also found profiles with two clay maxima. In these soils, as in the HMa and HUa pedons of the present study, there was no compelling evidence for the presence of lithologic discontinuities. Kesel and Spicer (1985) attributed the relatively low amount of clay between the two maxima to intense and prolonged weathering, of the clays, and subsequent leaching of the weathering products. If the explanation of Kesel and Spicer (1985) applies to HMf and HUa, a clay maximum may have existed in these soils between the two observed bulges in clay content. Therefore, data of the horizon with the highest clay content, whether it belongs to the shallow or deep maximum, will be used hereafter to characterize the most clay-rich horizon of these pedons. More firm conclusions about the origin of the second clay maxima in HMf, HUa and BUa, and the bisequal nature of these soils, could possibly be drawn if textural and fine clay data were available for all horizons, including the C horizons. The most clay-rich horizon (Bwl in BUa, see above) is shallower in soils on the foot of the deposits than in the corresponding soils on the apex of the deposits. In a more general sense it is also true that on these debris flow deposits, and on the saprolite, the clay maxima occur at greater depths in higher landscape positions (Figure 33). For the lowest deposits at the Hidden Valley and Bradley sites this increase in depth with elevation can be explained by a relatively shallow water table at the foot of the deposits. Because the other soils are well drained differences in drainage are not likely to have determined the depth of the clay maxima on the other deposits and the saprolite, although some influence can not be completely excluded. Considering that higher deposits are younger than lower deposits at 141 mm 23mm Em: 2 :m_ Eo: mmmmmLoE co=m>m_m .mzm comm c_5_>> .4 0mm xmam -. com I xmam .. 02. m 62 .m. U- xmam ~08 .. 00_. m xmam I a 000% xm m - 502 008 002 H cm I I xmam 002 a a a . . .1 . + . a h r a o 8 Tu mm mm” H H H mm H W V v V pm». 9 t. e mm W W: e H e m m m mmzm __m .EmEoo >m_m E:E_me 65me 142 the same site (see Part HI), the depth of the clay maximum may be related to age. Age dependence of the depth to clay maximum has been reported from a variety of environments (Levine and Ciolkosz 1983, McFadden and Weldon 1987, Fanning and Farming 1989). The shallow position of the clay maximum in the highest pedons at Alarka (AUa) and Bradley (BUa) (Figure 33) may be due to erosion. It has been argued above (6.3.2.1 and 6.3.2.3), based on the depth profile of free iron, that erosion may have removed surficial horizons at these locations. The shallow position of the clay maximum at AUa and BUa supports this notion. Additionally, pedon BUa is located on the steepest slope at the Bradley site (19 %), which indicates that increased surficial erosion may occur at this location. 6.4.4 Clay mineralogy Although chlorite is present in some of the bedrock at the study sites, it was not found in any of the soil samples. Considering the weatherability of chlorite and the high intensity weathering environment of the area, chlorite has probably been altered to other 14 A minerals. Chlorite was detected in other types of colluvium in the Blue Ridge Front but its presence was hard to explain (Graham et al., 1989b). Slightly higher amounts of 14 A minerals (probably vermiculite and hydroxy-interlayered minerals) are present in surface horizons than in subsurface horizons. This trend is best expressed at Alarka and Bradley. A similar trend was observed by Daniels et al. ( 1987b) and Norfleet and Smith (1989) and was related to an increasing trend with depth of mica. Both trends are explained by these authors by intense alteration of mica to 14 A minerals in surface horizons. In this study, no depth trend for mica was found at Alarka and Bradley and mica was absent at Hidden Valley. The lack of a trend for mica is most likely due to the advanced degree of weathering of the soils. Mica may be present in amounts so small that any variation in its abundance is 143 below the resolution of the quantification method that was used. Colluvial soils have been found to contain more (Daniels et al., 1987b; Graham et al., 1989b) and less (Norfleet and Smith, 1989) vermiculite and HIV than saprolitic soils. In this study, however, there is no difference in the relative amount of vermiculite and hydroxy-interlayered minerals in colluvium and saprolite. Mica has been shown to weather, here and elsewhere, to 14 A minerals such as vermiculite and hydroxy-interlayered minerals (Daniels et al., 1987b; Norfleet and Smith, 1989; Feldman et al., 1991). From common weathering sequences, it is known that upon alteration, 14 A minerals can produce kaolinite. This suggests that the soils at Alarka, which have more mica than the soils at the other two sites, have undergone less weathering than those at the Hidden Valley site, which has proportionally the most kaolinite and the least micas. The degree of weathering is usually related to bedrock composition, time and climate (Ollier, 1984). Differences in bedrock composition are most likely not the main factor that determines the differences in degree of weathering in this case. The bedrock composition at the three sites is similar and variations in the composition within one site are of the same magnitude as the variations between sites (Table 7). If bedrock composition were a chief factor, one would expect clay mineralogy to be variable within each site. That is not the case. Climate is not considered to be an important factor either, because the climate at the three sites is very similar. Detailed climatological observations are not available for the sites but the three sites are located in the same general region, at approximately the same elevation and have comparable aspects and slope. Therefore, the principal factor influencing the degree of weathering, and thus the clay mineral composition, of the studied deposits is time (age of the deposits). This is supported by the results of the relative dating (see part IH). The mineralogical composition of the clays at the Alarka site, and to some extent the Bradley site, includes minerals associated with initial stages of weathering. Their presence suggests that inheritance of clays may have been low in these debris flow deposits, unless 144 both the source material and the described soils are very young. If the source material was very young, and therefore relatively unweathered, it also seems reasonable to assume that inheritance of clays was not high. 6.5 Conclusions Two of the pedons studied have developed in saprolite; these two pedons have higher clay contents and more clay enrichment than do the other colluvial soils. This difference in clay content and clay enrichment is at least in part due to a difference in age (see below). Many lithologic discontinuities exist in the debris flow deposits. The discontinuities at the Alarka site and in two Hidden Valley pedons do not appear to represent sedimentological boundaries, but variations within the deposit of one event. Considering the sedimentological properties of the deposits, the variations can be linked to different shear zones in the flows. At the Bradley site the discontinuities represent lithologic discontinuities in the strict sense and correspond to boundaries between distinct sedimentary units. This variability in the parent material indicates that debris flows render the regolith more heterogeneous rather than homogeneous, as was observed in other colluvium (Graham et al., 1990b). The heterogeneity of the parent material affects the depth trends of most morphological properties of the soils. The only characteristic that shows a systematic trend is the depth to maximum clay and free iron accumulation. The trend is directly related to the topographic position of the pedon, higher pedons having maximum accumulation at greater depths: Depth to maximum clay and free iron accumulations in pedons on the apex of the deposits occur at greater depths than in pedons on the foot of the same deposit. The most likely explanation is an impeded drainage at the lowest topographic positions where vertical water movement in the soils may be restricted, and relative age. The presence of shallow 145 clay maxima in soils on the steepest slopes and the absence of eluvial horizons in some pedons may point to surficial erosion. The clay mineral assemblages show very little variation between the horizons of one pedon, and between pedons at each site. Only 14 A minerals show a weak, decreasing, trend with depth. This trend can not be explained by a more intense alteration of mica to 14 A minerals in surface horizons, as was the case in colluvium studied by Daniels et al. (1978b) and Norfleet and Smith (1989). The relative abundance of 14 A minerals in the soils on the debris flow deposits does not differ from that of nearby saprolitic soils. This finding is different from that of Daniels et al. (1987b) and Graham et al. (1989b) who found that other colluvial soils in the region usually contained more 14A minerals than did saprolitic soils. The main factor influencing the clay mineral assemblages of the soils on the debris flow deposits is time/age. To better understand pedogenesis on debris flow colluvium, more information on the characteristics of its parent material is required. These characteristics, i.e. texture, clay mineralogy and possibly depth profiles of chemical and physical properties, may be inherited from source materials and may be affected by transportation in the flow. It is therefore suggested that pedologic studies comparing recent debris flow deposits and their source material can further the knowledge of pedogenesis on debris flow colluvium. PART III: RELATIVE-AGE DATING 146 147 7 INTRODUCTION Several approaches used to date Quaternary geomorphic surfaces rely upon absolute ages. Often-used methods include thermoluminiscence, fission track, radiocarbon and other radioisotope dating methods (Kochel and Johnson, 1984; Shafer, 1984; 1988; Kochel, 1990). In the absence of suitable material for absolute-age dating, relative-age dating techniques have been used. Relative-age dating techniques have been employed to quantitatively estimate the age of a surface, using knowledge of rates of soil formation (Knuepfer, 1988), and to compare qualitatively the ages of a series of surfaces for which no rate information is available (Whitehouse et al., 1986; Harrison et al., 1990; Phillips, 1990; Birkeland and Berry, 1991; Markewitch and Pavich, 1991; Whittecar and Ryter, 1992; Woodward et al., 1994). Relative-age dating techniques utilizing soils data are based on the principle that the degree of soil development is dependent on five soil-forming factors (Jenny, 1941). Harden (1990) argued that the general magnitude of soil development strongly depends on the time factor, surface stability and possibly parent material. She submitted that the effect of the other factors (vegetation, climate and topography) is less apparent. Even if Harden's (1990) statement underestimates the importance of the other soil forming factors, soil development can be used to compare the relative ages of soils in different locations when those other factors remain more-or-less constant at all sampled locations. In this research, a qualitative (comparative) approach will be used. Clast weathering, color, clay content and iron species will serve as proxies for the degree of soil development and weathering. As argued above (sections 3.2, 3.3.3 and 6.4.4), the climate, vegetation, topography and parent material of the soils at the three sites are similar. Because the sites are located within 13 km of each other, the first three factors can be expected to have been similar in the past as well. 148 Although the mode of origin and the general nature of the parent material is the same, with the exception of the saprolitic soils, variations in the colluvial parent material may be due to compositional differences of the source material of the flows. It cannot be ascertained that the source materials of the different debris flows had the same degree of weathering and soil formation at time zero. Therefore it is conceivable that some of the characteristics of the soils on the debris flow deposits partly reflect the composition of the source material. However, as will be demonstrated in chapter 8, the degree of weathering of the clasts in the deposits is unlikely to have been influenced by preweathering of the source material. The similarity of the results for the other proxies to those of clast weathering suggests that the other proxies, too, may not have been influenced to a large extent by inheritance from the source material. It therefore seems reasonable to assume that time has been the chief factor in the determination of differences in soil development at the three sites, and that relative-age dating techniques can be applied without major obstacles. Strictly speaking, relative-age indicators give an estimate of the relative age of the soil and the present day geomorphic surface on which the soil developed (Phillips, 1990). However, it is assumed here that the mode of origin and the geomorphic evolution of the deposits is similar, and hence, that the relative-age relationships of the soils can be extended to the deposits as well. Therefore, "soil" and "deposit" will be used indiscriminately when references to relative-age are made hereafter. The purpose of the application of relative-age dating techniques is to provide an assessment of the relative ages of the deposits. Although relative ages of other colluvial deposits in the Southern Blue Ridge have been reported (King, 1964; Michalek, 1968; Gryta and Bartholomew, 1977; Mills, 1981), many of them were based on the degree of weathering and/or topographic position of the deposits. One of the very few soil chronosequence studies in the Southern Blue Ridge was carried out by Leigh (1995) on alluvial terraces just south of the present study area. In general, chronosequence studies from warm to hot, humid environments are scarce (Birkeland, 1990). This part of the 149 research may, therefore, contribute to the understanding of pedogenetic pathways in this type of physical environment. 1 50 8 CLAST WEATI-IERING 8.1 Literature review 8.1.1 General Clast weathering has often been used as a means of relative age dating of soils and deposits (Chinn, 1981; Colman and Pierce, 1981; Mills, 1982; Whitehouse et al., 1986; Knuepfer, 1988; Mills, 1988; Whittecar and Ryter, 1992). In many cases granitic boulders have been studied because they decompose relatively rapidly and uniformly in most environments (Birkeland, 1984). Various properties of clasts have been used to represent their degree of weathering. Most frequently, weathering rind thickness has been employed, but percent weathered boulders and depth of near-circular weathering pits have also been used (Birkeland, 1984). 8.1.2 Weathering rind thickness In New Zealand, Chinn ( 1981) showed that the thickness of weathering rinds on cobbles on Holocene geomorphic surfaces increases rapidly with time, described by a power function. Whitehouse er al. (1986) suggested that rind growth is not predicted by a power function for surfaces older than 20 ka because for those surfaces the rate of weathering approaches the rate of inward rind migration. Variations in annual precipitation and minor variations in lithology and mineralogy were found to produce little variation in the rate of weathering rind growth (Colman and Pierce, 1981; Knuepfer, 1988). Knuepfer (1988) observed that even the weathering of chlorite schist occurred at rates comparable to those of other lithologies, i.e. graywacke and high grade metamorphic rocks. The weathering of metamorphic rocks was different from 15 1 that of clasts of other lithologies because metamorphic rocks decomposed principally along foliation planes rather than by inward rind thickening. Nevertheless, weathering rinds on high grade metamorphic rocks have been used with success as a relative age indicator (Mills, 1983). Weathering rind thickness was found to be strongly dependent on depth of rock burial (Birkeland, 1984; Knuepfer, 1988; Mills, 1988). Initial weathering was characterized by rates of subaerial weathering that exceeded those of subsurface weathering; with time, subsurface rates became greater, although more variable than near- surface rates (Birkeland, 1984; Knuepfer, 1988). Measurements of weathering rind thickness have also been used successfully to date mass movements such as rock avalanches in New Zealand (Whitehouse, 1983) and debris flows in the Blue Ridge (Mills, 1983; Whittecar and Ryter, 1992). Mills (1983) identified three topographic levels of debris flow deposits and concluded that the age of the deposits increased from the lowest to the highest level. Weathering rind thickness increased consistently with increasing age. It was, however, not possible to unequivocally assign a deposit to a topographic level on the basis of weathering rind thickness alone because there was some overlap of the rind thickness of adjacent levels (Mills, 1983). Whittecar and Ryter (1992) obtained very similar results. They identified two groups of debris flow deposits at different topographic levels. Greenstone clasts in the soils of the higher level had significantly thicker weathering rinds than the clasts of the lower level, and were interpreted to be older. Other soil properties, such as maximum clay content and maximum free iron content, pointed to a similar age relationship between the two groups of debris flow deposits. 152 8.1.3 Percent weathered clasts The relative amount of weathered clasts has been used most frequently for relative- age dating of moraines (Mahaney, 1978; Burke and Birkeland, 1979; Miller, 1979). Mills (1981; 1982a; b; 1988) employed the method to determine the relative ages of colluvium in the Southern Blue Ridge of North Carolina and the Valley and Ridge of Virginia. Mills (1981; 1982a; 1988) defined a weathered clast as one that could be broken apart by hand. In the Blue Ridge, deposits considered to be young occurred at lower elevations and had considerably fewer weathered clasts than older deposits (Mills, 1981). The percentage weathered clasts correlated well with another relative age indicator, soil color, but did not correlate with the percent clay, which also has been used as a relative age indicator (see chapter 9). This lack of correlation was thought to be due to the breakdown of clasts and the subsequent incorporation of the component grains into the soil matrix (Mills, 1981). In colluvium of the Valley and Ridge the percent weathered clasts was also deemed an accurate measure for discriminating between older and younger deposits (Mills, 1988). In the same environment, however, sandstone weathering rinds proved too variable to use as a weathering index (Mills, 1988). 8.2 Methods As explained in section 5.2.1, 50 clasts were collected at random from 11 pedons. The horizons from which the clasts were collected are listed in Table 3. The clasts were characterized by size, sphericity, roundness, lithology and weathering. Almost all clasts were metamorphic (Table 7) and did not show weathering rinds. Therefore, the percentage weathered clasts was considered as an alternative proxy for relative age. A similar approach was used by Sharp (1969). 153 Determining the percentage weathered clasts has traditionally involved classifying clasts as either weathered or not weathered. In this research project, there was enough variability in the weathering of the clasts to assign them semi-objectively to one of three groups; little, intermediate, and highly weathered. The allocation of a clast to a certain group was based on (i) the physical soundness of the clast as determined by a sharp blow with a rock hammer, (ii) the dominant color of the inside of the Clast and (iii) the alteration of minerals as seen under a 10X hand lens. Clasts in the "little weathered" category typically needed many (> 4) hard blows to be broken, did not show oxidation colors on broken surfaces although staining in cracks was sometimes present, and had no clear alteration of minerals. Clasts in the "intermediately weathered" group did not need many (2- 4) hard blows with a hammer to be broken, had oxidation colors over most of the broken surfaces and had visible alteration of some minerals. The highly weathered clasts were easy to break with S2 gentle strikes with a hammer, were completely brownish or reddish inside and had many crystals that were visibly altered or completely dissolved. For each pedon, the percentage of clasts in each weathering group was determined. The results were then analyzed visually and by cluster analysis of standardized values. Both methods were used in conjunction because it is important to have control over the physical meaning of observations and conclusions. while it is also critical to keep a sense of objectivity. Harbor (1986) also concluded that, for mathematical reasons, it is important to combine a clustering procedure with a subjective analysis of individual data distributions. For the cluster analysis a K-means procedure, with 50 iterations and 4 groups, was used. This procedure produces non-hierarchical clusters by maximizing between-cluster, relative to within-cluster, variation. 154 8.3 Results and discussion The use of clast weathering as a relative—age indicator assumes that all clasts were equally weathered when the deposits were formed and that they have weathered at comparable rates since deposition. In the case of debris flow deposits the first assumption seems warranted. During transportation in debris flows, clasts interact with each other, the matrix, and obstacles on the surface and in the path of the flow. It seems therefore acceptable to assume that weathered clasts are destroyed during transportation and that primarily fresh clasts are incorporated in deposits. Mills (1983) and Graham et al. (1989a, b) also concluded that destruction of preweathered material occurred in debris flows. The rate of weathering (related to the second assumption above), depends chiefly on climate, vegetation, topography and bedrock composition. As argued in the introduction (chapter 7), the climate, vegetation and topography at the three sites are very similar. The lithology of the clasts encountered in the various deposits is also fairly uniform (Table 7). Most sites are dominated by gneisses that usually show little foliation. The main deviation of this mode was observed at AUf and AUa where metasandstone was the dominant lithology. The metasandstone is quartz-rich and thus, more resistant to chemical weathering than most other lithologies that were encountered. However, since most gneisses also contain relatively high amounts of quartz and are better crystallized, the influence of the abundance of metasandstone in the one Alarka deposit may not be large. This is confirmed by the presence of relatively fresh schist at the Alarka site. Schists generally are susceptible to weathering and their presence suggests that the unweathered nature of the metasandstone is not due to its composition alone. In fact, the presence of schist at the Alarka site and its absence in Hidden Valley could be the result, not the cause, of the difference in age indicated in the results of this chapter. For the above reasons, the rate of weathering of the clasts at the three sites will be assumed to be similar. The abundance of the clasts in the three weathering classes is listed in Table 12. 155 There are clear differences in the degree of weathering of the clasts at the three sites. At the Alarka site many clasts are little weathered, intermediate amounts are moderately weathered and few clasts are highly weathered. In Hidden Valley, most clasts belong to the interrnediately or highly weathered classes and very few clasts show little weathering. At the Bradley site clasts are distributed more evenly over the three weathering classes. l:ITable 12: Clast weathering data, all sites. Pedon- Low Medium High Cluster Horizon % % % ii ALfl 50 42 8 1 ALf2 57 37 6 l AUf 20 54 26 2 AUa 18 59 22 2 1' HLf 27 48 25 2 HMf 4 14 82 4 HUf-2Bt2 3 63 34 3 HUf-3Bt3 8 38 54 4 HUa 5 26 69 4 I BLf 14 55 3 l 2 BLa-2Bw2 25 5 l 24 2 BLa-43 W4 5 l 49 0 1 BUa-3Bw3 26 52 22 2 BUa-5Bw5 l 5 70 l 5 2 Within the Alarka and Hidden Valley sites there is a systematic variation in degree of weathering of the clasts between the lowest and highest deposits. Pedons ALfl and ALf2, which are in the lower landscape position at the Alarka site, have more little- weathered and less highly-weathered clasts than AUf and AUa, which are located on the higher deposit. The same trend can be observed at Hidden Valley, where HLf has more 156 little weathered and less highly weathered clasts than the other (higher) pedons. HMf has many more intensely weathered clasts than any of the other locations. Its greater relative age is corroborated by the existence of saprolite at this location (see section 6.4) Three pedons (HUf, BLa and BUa) are represented by samples from two different horizons. Although there is a difference between the two samples from within each pedon, the variation is neither large nor systematic. The only noteworthy difference is the larger amount of little weathered clasts in BLa-4Bw4 (1 1-187+ cm) as compared to BLa-2Bw2 (45-79 cm). This observation can be explained by the generally accepted notion of decreasing weathering intensity with depth. The results of the cluster analyses confirm these visual observations (Table 13). All but one pedon at the Hidden Valley site are concentrated in two clusters. In these two clusters, samples from the other two sites are not represented. These two clusters represent pedons with high amounts of highly weathered clasts. Also all but one pedon of the Bradley site are concentrated in one cluster (cluster #2). In Alarka the two lowest Alarka pedons, AU] and ALf2, do not co-occur with the two higher Alarka pedons. The lowest pedon of Hidden Valley (HLf) was grouped together with the two higher pedons of Alarka and several Bradley pedons. The general concentration of the pedons of one site in one or two groups seems to indicate that the within-site relative-age difference is small. 8.4 nclusions Although the results of this chapter will be combined with those following, they point to the existence of distinct differences in age within and between the three sites. The Alarka site has the least number of weathered clasts and therefore appears to be the youngest of the three. The Hidden Valley site, and especially the saprolite in HMf, has many highly weathered clasts, indicating an older age. All but one of the Hidden Valley pedons were classified in separate clusters, suggesting a similarity in age but also ages that 157 are unlike those of the other two sites. At the Bradley site none of the three weathering classes clearly dominates over the other two, which indicates that the age may be intermediate between the Alarka and Hidden Valley sites. Table 13: Cluster anal ses of clast weatherin ; data. Cluster* Pedons Variable Standardized mean (Clast weatheringgroup) 1 ALfl, ALf2, BLa-4Bw4 Low 1.65 Medium -0.29 High -1.07 2 AUf, AUa, HLf, BLf, Low -O.13 BLa-2Bw2, BUa-3Bw3 Medium 0.58 BUa-5Bw5 High -O.27 3 HUf-2Bt2 Low -1.12 ‘ Medium 1.09 High 0.18 l 4 HMf, HUf-3Bt3, HUa Low -0.97 Medium -1.42 Hih -l.63 * A higher cluster number indicates an older relative age. At the Alarka site the higher deposit appears to be older than the lower deposit. In the cluster analyses the higher deposit was grouped together with the deposits of the Bradley site, suggesting that it may be similar in age to the Bradley deposits. The highest deposit at the Hidden Valley site also appears to be older than the lowest deposit at that site. In summary, the results of this chapter, which have to be considered preliminary, show that there is a difference in age between the three sites, Alarka being the youngest and Hidden Valley being the oldest. At the Alarka and Hidden Valley sites, the lowest deposit is younger than the highest. l 5 8 9 SOIL DEVELOPMENT 2.1 Literature review Soil morphological properties have been used extensively, and successfully, as tools for relative-age determination in geomorphic studies (e.g. Birkeland, 1978; Levine and Ciolkosz, 1983; Harden, 1990; Birkeland et al., 1991; Markewitch and Pavich, 1991; Woodward et al., 1994; Leigh, 1995). Many soil morphological properties have been used to assess the relative age of soils, including percent clay in B horizons (Leigh, 1995), color (Meixner and Singer, 1981; Harden, 1982; Dethier, 1988; Leigh, 1995), decalcification depth (Woodward et al., 1994), clay illuviation (Dethier, 1988; Markewitch and Pavich, 1991; Woodward et al., 1994), solum thickness (Dethier, 1988; Markewitch and Pavich, 1991; Vidic et al., 1991; Leigh, 1995), organic matter in B horizons (Woodward et al., 1994), overall texture (Van Wambeke, 1962; Meixner and Singer, 1981; Harden, 1982; Harrison et al., 1990), consistence (Meixner and Singer, 1981; Harden, 1982) and structure (Harden, 1982; Phillips, 1990). Several problems, however, are associated with the use of soil development as a relative age indicator. Switzer et al. (1988) pointed out that the technique assumes that soil development is continuous but that, on small timescales, soil development may be episodic. Harrison et al. (1990) showed that the degree of variability of soil properties for a given surface and landscape position can be sufficient to raise questions about the usefulness of quantification of soil - time relationships. Phillips (1990) cautioned against the use of soil development as an indicator of the relative age of the parent material of the soil because surficial processes (i. e. erosion and deposition) may have inhibited pedologic development. He argues that soil development should be used to characterize the relative age of contemporary geomorphic features. Harden (1990) came to a similar conclusion and added that an improved comprehension of intrinsic, soil-driven mechanisms is needed to better 159 evaluate soil development. Nevertheless, she concluded that soil morphology has proven to be a convenient tool for assessing degree of development. Most chronosequence and relative-age dating studies of soils have been carried out on moraines or in semi—arid environments. Few studies from humid areas in general, and from the southeastern US in particular, have been performed. Most work in the southeastern US was carried out on the Coastal Plain (Daniels, 1978; Markewitch et al., 1986; 1987; 1988; 1989; Phillips, 1990; Markewich and Pavich, 1991). Markewitch and Pavich (1991) investigated the usefulness of chronosequence studies in a low relief area with a temperate to subtropical, humid climate. They found that in their area soils had developed along two different pathways; one resulting in the formation of Spodosols and one in Ultisols. No clearly distinguishable age trends could be identified in the Spodosols but the morphological properties of Ultisols could be utilized to characterize surface material alteration through time. The best indicators of the degree of soil development among soils of Pleistocene age were solum thickness, thickness of the argillic horizon, color of the B horizon, clay mass in a unit column of the solum and several chemical characteristics. The most useful properties for age comparison of older soils included solum thickness, clay mass and color (Markewitch and Pavich, 1991). These properties increased with increasing time. Properties of Holocene soils were variable and sometimes could be confused with geologic properties of the parent material. This problem was in particular true for horizonation, structure and texture. Clay mass was considered to be the best overall relative age indicator by Markewitch and Pavich (1991). Leigh (1995) studied floodplain and terrace soils about 45 km southwest of the study area of the present research. The floodplain and the lowest terrace were dated by 14C to 1.9-0.6 ka and 13.3 ka respectively. Floodplain soils were Entisols, on the lowest terrace Alfisols had developed and on the higher terraces Ultisols were observed. The B horizons of the higher, and therefore older, terraces contained progressively more clay and were redder than those of lower terraces. 160 221mm 9.2.1 Selection of relative-age indicators The selection of the most appropriate relative age indicators is mainly based on findings by Markewitch and Pavich (1991) and Leigh (1995). Both studies took place in soils and climates that are similar to those of this research. Furthermore, Leigh's (1995) study area is located in the Southern Blue Ridge. Both Markewitch and Pavich (1991) and Leigh (1995) judged soil color and clay content to be good relative-age indicators. To represent color, Markewitch and Pavich (1991) used Munsell hue while Leigh (1995) used the Hurst index. The amount of clay was characterized by Markewitch and Pavich as clay mass in g cm'2, integrated through the thickness of the solum, which is associated with the degree of weathering. Leigh (1995) used the percent clay in the most developed B horizon. Markewitch and Pavich (1991) also ranked thickness of the solum and the argillic horizon as good indicators. All the above observations are not in contradiction with generally accepted principles (Birkeland, 1984), and therefore, the focus of the rest of this chapter will be on color (rubification) indices and measures of clay content as relative-age indicators. Thickness of the solum is not a suitable parameter for this study because the C, and even BC, horizon was not reached in four of the 13 pedons. Moreover, thickness of the solum is sometimes difficult to assess because horizon boundaries may be gradual or diffuse. As an alternative for solum thickness, the depth of maximum clay content will be considered as a thickness criterion. An argillic horizon is only present in six pedons and in the soil pit of two of these, the bottom of the argillic horizon was not encountered. It therefore seems more appropriate to consider presence of an argillic horizon rather than its thickness. 161 9.2.2 Data collection and analyses Field data were collected and basic calculations performed as explained in sections 5.2 and 6.2. Results of these operations are reproduced in Table 4 and Appendices 3 and 4. The most frequently used rubification indices in chronosequence studies are the Buntley-Westin index (Buntley and Westin, 1965), Hurst index (Hurst, 1977) and the redness rating (Torrent et al., 1983). The Buntley-Westin index (BW) quantifies the hue (10R = 7, 2.5YR = 6, 5YR = 5, 7.5 YR = 4), which is then multiplied by chroma. It is closely allied with the intensity of weathering, stage of soil development and drainage (Buntley and Westin, 1965). The Hurst index (HI) quantifies hue (10R = 10, 2.5YR = 12.5, 5YR =15, 7.5YR = 17.5) and multiplies it by the value/chroma fraction. The H1 is a crude proxy for the total amount of iron in a soil (Hurst, 1977). The redness rating (R) is calculated by the formula (10-YR hue) x chroma/value. The R has been linked to hematite content (Torrent et al., 1983; Graham et al,. 1989a). As mentioned in 9.2.1, Munsell YR hue has also been used as a color index (Markewitch and Pavich, 1991), but it is a less refined measure than the above three indices. The calculations were performed on the Munsell color of the reddest horizon of each profile. Because the HI decreases with increasing redness, and presumably increasing age, and the BW and RR increase with increasing redness, the reciprocal of the HI was graphed in Figure 34. To represent the clay content of the soils, three parameters were calculated; percent clay in the most clay-rich horizon (Bwl for BUa, see section 6.4.3), the clay mass of the most clay-rich horizon, expressed in g cm'3 (Bwl for BUa, see section 6.4.3), and the clay mass in g cm‘z, summed over the thickness of the solum. The clay mass was obtained by multiplying the bulk density of the most clay-rich horizon by the percent clay. (An underlying assumption is that all particles in a soil have the same density.) Clay mass is considered to be a better measure for the amount of clay than percent clay because clay 162 mass is an absolute value that permits quantitative comparisons. The clay mass per cm2 of a solum-thick column was derived by multiplying the clay mass of each horizon by its thickness and by integrating the results over the horizons in the solum. A cluster analysis identical to the one explained under 8.2 was performed on the six variables. As mentioned above, the bottom of the solum was not reached in all pedons. Therefore, the clay mass per square centimeter has to be regarded as a minimum value in the case of the two highest pedons at Hidden Valley and Bradley (HUf, HUa, BLa, BUa). 9.3 Results and discussion 9.3.1 Rubification The results of applying the three rubification indices to the various pedons are very similar (Figure 34). The lowest values are present in the Alarka pedons, intermediate values can be observed at the Bradley site and the highest indices are for the pedons at Hidden Valley. The values at the Bradley site are somewhat closer to the values at Alarka than to those at Hidden Valley. The three indices show identical relationships for pedons within the Alarka site; Pedons AU 1, ALf2 and AUa have the same value and AUf, which is on the highest deposit, has a slightly higher value. At Bradley there is no consistent relation between the three pedons. The R increases with increasing elevation but the HI and BW show no trend with elevation. At Hidden Valley the highest value is always reached for HMa, one of the soil pits that was in saprolite. As compared to the other Hidden Valley pedons, HLf and HLa, which are on the lowest deposit, have lower values for the HI and RR and intermediate values for the BW. 163 3 2:9". x85 Lo .8898. 9505 £55 .. 4'18 lnH Will-l "NH 91H NH 911V 10V 311V l-HV (DV'NO ..NF -3 L: -3 L.ow mmm wmwmmm wmmw r o r- or . ow -, om ,. ow .- om -- ow 908 9'18 118 9114 MH aWH "NH 91H l'IH 90V 111V 311V 111V . cod . Nod . #06 . mod .. mod .. OTC -. «v.0 .. .vwd .. ofo .. 220 .. 0N6 maze mmmcumm xmuE c:mm>>->m=::m mmzm =m .mmoEE cosmoEnzm ..xouE .92.. 164 A systematic difference between pedons in toe and apex positions is not apparent (Figure 34). In Alarka, pedon AUf, which is in a toe position, has higher values than the pedon on the apex of the same deposit (AUa). In Hidden Valley there is no difference between the pedons on the toe and apex of the lowest deposit (HLf and HLa) and highest deposit (HUf and HUa), except for the BW of the highest deposit. In the latter case the pedon on the toe has a lower value than the pedon on the apex. As mentioned above, a systematic trend with elevation or geomorphic position is not present in the Bradley site. 9.3.2 Clay content In general, the three parameters representing clay content show the same trends (Figure 35). The lowest values are observed at Alarka, the pedons at Bradley have comparable to slightly higher values, and the highest values are reached at Hidden Valley. The presence of argillic horizons at the Hidden Valley site, and their general absence at the other two sites, substantiates this observation. The highest clay content in a B horizon is at HMf, one of the saprolitic soils, while the highest clay mass in a solum is at HUf. The clay mass in the most clay-rich B horizon and the clay mass in the solum have opposite trends in Alarka but the same trend in Hidden Valley. This indicates that at the Alarka site the B horizons are either clay rich or thick while at the Hidden Valley site they are usually both clay rich and thick. The clay content of the B horizon and clay mass of the solum of pedons on the toe and the apex of a deposit do not show a systematic difference (Figure 35). However, at Alarka the clay content of the B horizon is higher in the apex position (AUa) than in toe position (AUf) while on the saprolite and the highest deposit of Hidden Valley it is lower in the apex position. The low amount of clay in the sola of HLf and BLf is due to a shallow water table which prevents the sola from developing further downward. This means that at mm 2:9". .039 5533305 32.2w 2 can 2m 5m new SI .2 mm:_m> So. 65 202 . 53.0: :2??? .88 B 53:53 ..2 med cozomm mom a H HHH vvvv a H HHH vvww a H HHH vvww mmm wmwmmm mmmm mmn wmwmmn mmau mmn wmwmmn mmau fl o .. fl 0 + fl 9 .om ,fio -o. .. o v . w o . -- ow .. m o z. oo .. v.0 -- on 5 H ow m -90 o4 L. on: -. m6 -. c3 .. no .. om .. 31 L w.o s- cm m-Eo m «-80 a .anw 5 >30 8.2.0: 52188 .86 c_ mums >30 co~toz c2166 60.: :_ >30 28.8 mmzm =m .EoEoo >30 166 these two locations time is not the prime factor that has determined the clay mass in the profile. No inferences can be made for the clay mass in the sola on the highest deposit in Hidden Valley (HUf, HU a) and Bradley (BLa, BUa) because at these locations the lower boundary of the solum was not encountered. This does not, however, appear to influence the conclusions to a large extent. Values for clay mass in the solum of HUf and HUa are among the highest encountered anyway. The results for BLa and BUa indicate that the internal relationship of the site, as indicated by the other two parameters, does not hold for the clay mass in the solum, regardless of the magnitude of the underestimation for BLa and BUa. The results for the two latter pedons, obviously, does affect between-site comparisons. The cluster analysis, which was run on all six soil development indices, grouped the Alarka and Bradley pedons into one cluster. The second cluster contained pedons HLa, HMf, HUf and HUa. Pedons HMa and HLf were each assigned to a one-member cluster. A cluster analyses with three groups classified HLf together with the Alarka and Bradley pedons. This would have left HMa as the pedon that is most different (older) from the other pedons. 2.4 Conclusions The variation in relative age, as indicated by soil development indices, is larger and more consistent between sites than within sites. But, conclusions about the implications of within-site relative-age variations should be evaluated more cautiously than the conclusions about between-site variation (Harrison et al., 1990). Between-site variations consistently point to a relatively young age for the soils at Alarka, an intermediate age for those at Bradley and an older age for the soils at Hidden Valley. Assuming that the rate of increase with time of the color indices and clay content is 167 not logarithmic, the soils at the Hidden Valley site appear to be considerably older than those at the other two sites. Most indices indicate that pedon AUf is slightly older than the other Alarka pedons. This is, however, not confirmed by the clay content in the B horizon. Pedon AUf also was grouped together with the other Alarka pedons in the cluster analyses. At Hidden Valley, there is some evidence to suggest that, at this site as well, the highest deposit is older than the lowest deposit. The saprolite at Hidden Valley, which topographically is located between the two deposits, appears to be older than any of the deposits. The results for the lowest pedon in Hidden Valley, HLf, are typical of relatively young soils. This pedon was grouped singly by the 4-group cluster analysis. Its assignment to the group of the Alarka and Bradley pedons by the three-group cluster analysis suggests that HLf is similar in age to the pedons at these two sites. The lack of systematic variation between pedons on the toe vs. apex of the deposits points to the effectiveness for relative-age dating of the chosen variables. Although there are morphologic differences between pedons on toe and apex positions (see chapter 6), relative-age indicators should show less difference between soils on a single geomorphic surface. Nevertheless, if the relative ages postulated in chapters 8 and 9 (Hidden Valley older than Alarka, higher deposits slightly older than lower deposits) are correct, some of the variations in clay content on a geomorphic surface may be related to age. The available data suggest that B horizons on the apex of a deposit contain more clay than B horizons on the toe of the deposit in young soils but that in old soils they contain less clay on the apex than on the toe. 1 68 10 [RON SPECIES 10.1 Literature review 10.1.1 General Many authors have concluded that certain iron species, their abundance and ratios thereof correlate well with soil age and profile development. The most frequently used iron species include total ferric iron (Fed, extractable with sodium dithionite), ferric iron present in poorly crystalline oxyhydroxide phases and organic complexes (Feo, extractable with acid ammonium oxalate), and total iron (Fet). Regularly employed iron species ratios include Fed/Fet, Fed-Feo/Fet, and Feo/Fed. This literature review summarizes the climatologic, pedologic and geomorphic settings of studies by Alexander (1974), Torrent et al. (1980), Alexander and Holowaychuk (1983a, b), Arduino et al. (1984, 1986), Rebertus and Buol (1985), McFadden and Weldon (1987), Graham et al. (1989a) and Leigh (1995). Alexander (1974) examined a sequence of terraces in a sub-humid climate. The soils on the floodplain were Cumulic Haploxerolls and those on the terraces classified as Typic Argixerolls. In the Typic Argixerolls samples were taken from the argillic horizon. Torrent et al. (1980) studied trends of Fed, Feo, Fed/Fet and Feo/Fed with age in xeric Alfisols on two terrace sequences. The authors sampled the upper part of the Bt horizon because it usually had the most pronounced red color. Alexander and Holowaychuk ( 1983a, b) studied a chronosequence on terraces under a hot, humid climate. The soils on the floodplain were Aquents and Aquepts; on the intermediate terraces they classified as Ustults, and on the highest terrace an Ustox had developed. Arduino et al. (1984) sampled the most weathered horizons of Alfisols, Entisols and Inceptisols on terraces under a Mediterranean climate. The age of the terraces ranged from Middle Pleistocene to 169 Holocene. Rebertus and Buol (1985) studied Hapludults and Dystrochrepts in regolith of the Piedmont and Blue Ridge of North Carolina. Some of the study sites were located 120 km NE of the study area of the present research. Arduino et al. (1986) examined Alfisols, Inceptisols and Entisols on three terrace levels under a Mediterranean climate. The maximum age of the highest terrace was estimated to be 300 ka while the youngest terrace was thought to be at least 10 ka. McFadden and Weldon (1987), working in a Mediterranean climate in SW California, had good absolute age control for 11 terraces ranging in age from 500 ka to 47 a. Characteristics of the area and soils studied by Leigh (1995) were discussed in section 9.1. 10.1.2 Fed and Feo Torrent et al. (1980), who studied two terrace sequences, found an increase with age of Fed, but not of Feo, in soils on one of the sequences. They did not find a trend for these two species in the other sequence. Alexander and Holowaychuk (1983a, b) concluded that the amount of Fed was very similar in the soils they studied, although their data and their discussion indicated that Fed increased as soil age increased. The amount of Feo decreased with terrace age in this study. Arduino et al. (1984) stated that in their study area there was an increasing trend with age of Fed and a decreasing trend with age of Feo. Knuepfer (1988) noted an increase of Feo in young soils (<12-14 ka) but a decrease in older soils, probably due to re-solution of amorphous Fe. Rebertus and Buol (1985) and Leigh (1995) reported that the Fed of B horizons showed an increase with soil development, and thus with soil age. McFadden and Weldon (1987) noted that Fed, Feo and Fet maxima of Holocene soil profiles were reached in the A or Bw horizons and that in Pleistocene soils maximum concentrations of Fed and Feo were reached deeper in the profile, usually in B horizons. From a graph presented by Alexander and Holowaychuk (1983) and from the data of 170 Arduino et al. (1986), similar conclusions about progressively deeper Fe maxima can be drawn. In the study of McFadden and Weldon (1987) Fed content was highly correlated with terrace age (1:094). Fed content often correlates well with soil redness. Graham et al. (1989a) observed a strong correlation of this species with the redness rating and McFadden and Weldon (1987) observed one with the Hurst index (r=0.76 to 0.92 at 0L=2.5% level). 10.1.3 Fed/Fet, (Fed-Feo)/Fet and Feo/Fed In one terrace sequence Torrent et al. (1980) found an increase in Fed/Fet with age but Feo/Fed did not show a trend. In the other sequence there were no apparent age trends of iron minerals but instead a high (i 0.80) Fed/Fet ratio was observed. Torrent et al. (1980) concluded that this implies that the parent materials of the soils must have had similar, and high, Fed/Fet ratios, and thus must have been highly weathered. The sequence that showed trends for Fed/Fet also had increasing amounts of goethite and hematite with increasing terrace age. Feo/Fed is an index of crystallization of secondary forms of iron that in the initial stages of weathering and soil development, when the release of iron from minerals may exceed the rate of crystallization, increases. Thereafter, the ratio decreases with increasing soil age. In the study by Alexander and Holowaychuk (1983a; b) the best age trend was obtained for this index. Feo/Fet decreased as age increased but also decreased substantially from the surface to the subsoil and thence increased slightly below the solum. Arduino et al. (1984) obtained the best trend with age with (Fed-Feo)/Fet. This ratio is a measure of the amount of ferric iron in crystallized phases, as a proportion of the total amount of iron. The largest proportions of Fed-Feo were present in the oldest terraces, indicating that progressively more iron was released from silicates as time advanced 17 1 (Arduino et al., 1984). The ratio was independent of the total amount of iron in the samples and therefore of the general lithology of the basin (Arduino et al., 1984). Rebertus and Buol (1985) concluded that the Fed/Fet ratio was a sensitive indicator of profile development, and therefore, of soil age. The Fed/Fet ratios in the B horizons were much better indicators of relative soil development than those from either C or A horizons. The Fed/Fet ratio also changed nearly systematically with particle size class and solum thickness. As age increased, Fed/Fet and solum thickness increased and overall texture became finer. Arduino et al. (1986) found that both Fed/Fet and Fed-Feo/Fet served to distinguish soils of different age. In this study Feo/Fed was generally found to be only a crude indicator of soil age. McFadden and Weldon (1987) noted that the Fed/Fet ratio of Holocene soils increased slightly with age, while in Pleistocene soils Fed/Fet increased significantly with age. The average Feo/Fed ratio for the solum increased for Holocene and latest Pleistocene soils but decreased for older soils (McFadden and Weldon, 1987). Alexander (1974) observed a similar trend for Feo/Fed. The breakpoint where the trend changed from increasing to decreasing with age occurred after "several thousand years". McFadden and Weldon (1987) concluded that, in their study area, iron oxide content was a potentially excellent indicator of age of Middle Pleistocene or younger deposits. 1 2 8 10.2.1 Selection of relative-age indicators As shown by the literature review, Fe species ratios are generally considered to be better relative age indicators than are individual species data alone. Ratios have the advantage that variations in the parent material are compensated for. If, for instance, a high amount of Fed is the result of a high Fet content in the parent material, and not of the 172 degree of weathering, dividing Fed by Fet will eliminate the influence of the parent material. Therefore, the ratios can be considered "standardized" values. To better compare the effectiveness of the various ratios (Feo/Fed, Fed/Fet and Fed-Feo/Fet) as a relative age indicator, I calculated the ratios that were not determined by the authors. I determined Fed-Feo/Fet for the data of Torrent et al. (1980) and Fed/Fet and Feo/Fed for the data of Arduino et al. (1984). Other authors did not provide raw data that allowed me to determine other ratios than those that they had calculated. Among the authors that discussed Feo/Fed, only Alexander and Holowaychuk (1983), working on soils in a hot humid climate, considered this ratio to be a good relative age indicator. For the data of Arduino et al. (1984) I did not find a good correlation of Feo/Fed with age. Therefore, this ratio will not be used in the present research. Other authors obtained the best relative-age dating results with Fed/Fet and/or Fed- Feo/Fet. From the data of Arduino et al., the ratio calculated by me (Fed/Fet) had a trend that was very similar to the Fed-Feo/Fet trend used by the authors. For the data of Torrent et al., I found a trend for Fed-Feo/Fet that was comparable to the trend of Fed/Fet obtained by the authors. Together with the results from the literature review, this pre-analysis procedure suggests that trends for Fed/Fet and Fed-Feo/Fet are closely related, and that in many cases the ratios lead to similar results. Unfortunately, a quantitative comparison of the effectiveness of the two ratios can not be made because of a lack of quantitative age data. Rebertus and Buol (1985) worked in the Southern Blue Ridge of North Carolina, in a clirnatologic and geologic environment that is very similar to that of the present study. Because they attained better results with Fed/Fet than with Fed-Feo/Fet, the former ratio will be used in the present study. 173 10.2.2 Data collection and analysis Field data and samples were collected as explained in sections 5 .2 and 6.2. Procedures for extracting Fed and Fet were according to Jackson et al. (1986) (procedure 64.3) and Lim and Jackson (1982) (procedure 1-5) respectively. For each horizon a sample of about 10 g was separated from the main sample with sample splitters to obtain a representative sample. The 10 g sample was ground to silt size in an agate mortar. Replicate subsamples of about 4 g were taken for Fed analysis, replicate samples of about 0.2 g were taken for Fet analysis. The subsample for Fed analysis was transferred to a centrifuge tube and a 5:1 mixture of 0.3 M sodium citrate (N a3C6O7.2H20) and 1 M sodium bicarbonate (NaHCO3) was added. Subsequently, the mixture was heated to 75 to 80° C, at which time about 1 g of sodium dithionite (N a2S204) powder was added. After a 15 minute digestion period under repeated stirring, the tube was placed in a centrifuge, spun at 2000 rpm for 6 minutes, and the supernatant decanted. This procedure was repeated to remove any ferric iron that might have been left by the first extraction. The supematants of the two extractions were combined and diluted for measurement on a directly coupled plasma spectrometer. The 0.2 g subsample for Fet analysis was first mixed in a polypropylene bottle with 1 ml of aqua regia (3 parts 12 N HCl and 1 part 15 N HNO3). After about one hour, 10 ml of 48 % hydrofluoric acid (HF) was added and the bottle capped immediately. The mixture was then shaken overnight to dissolve the sample. Afterwards, 100 ml of saturated boric acid was added to decompose any metal fluorides that might have formed and the bottle was shaken again for 6 to 8 hours. Subsequently, the contents of the bottle were diluted to 200 g and analyzed on a directly coupled plasma spectrometer. Results for Fed, Fet and Fed/Fet are provided in Appendix 8. Because there was variation in the Fed/Fet ratio with depth (Figure 36), the profile weighted mean ratio for each pedon was computed. The mean was calculated by the formula 174 (2(ratioi*thicknessi))/solum thickness, where i designates the horizons in the solum. By using the solum, rather than the bottom of the soil pit, C horizons were excluded from the calculations. (C horizons were not encountered in all soil pits). Including them for some pedons only, would have biased the average of those pedons towards a lower value because, generally, C horizons have lower Fed/Fed ratios than do B horizons. A cluster analysis identical to the one explained under 8.2 was performed on the solum means. In this one dimensional case, the routine simply provides a grouping of the pedons along a single axis, providing an objective alternative to the visual assignments of groups on a histogram. 1 . R sults and discussion The general variation of Fed/Fet with depth, demonstrated in Figure 36, can be observed in saprolitic (HMf, HMa) and colluvial (all other pedons) soils alike. This indicates that weathering products do not gradually decrease with depth in these soils. This finding can possibly be explained by variations in the composition of the soil water. As argued above, the effect of compositional changes in the parent material is expected to be small because the ratio has the total amount of iron (Fet) in its denominator. Figure 37 shows that the pedons of the Alarka site have the lowest Fed/Fet ratio and that those at the Hidden Valley site generally have the highest values. The pedons at Bradley have intermediate values. The values within the Alarka site are very similar (25 - 28 %). Also within the Bradley site there is little variation although the ratio slightly decreases as elevation increases. This may be an artifact of the misrepresentation of the ratio at BLa and BUa where the bottom of the solum was not encountered. At the Hidden Valley site the pedons on the lowest deposit have the lowest values (HLf and HLa). The ratio is maximal in one of the saprolitic soils (HMf) and slightly lower in the pedons on the highest deposit (HUf, HUa). mm 2:9". ~31 6:955 :30 m3< ”295w come m2: ”@533 :80 5< 68:3 8:: SI acoesu 8:: :2: ”2mg... 8:: % 175 1 0mm 1 0mm .. CON 1 CON 1. omw G i- om— G a 9 d d m. .m. o 3 r- 00* w 1 Dow w -- om l om L o L 0 ON 0 or o .. mconma 8898 .5qu Lo cozoca m mm 6&8". 176 R 2:9“. m3< 5< N3< 5.2 SI SI MEI :21 SI SI mam 3m Sm iuaOJad 8% __m 8&8; 8236; 2:05 :82 177 The cluster analysis classified all four pedons of Alarka and HLf and BUa, in one group. The three highest pedons in Hidden Valley (HMa, HUf, HU a) were grouped in a second cluster. The third cluster consisted of HLa, BLf and BLa, while pedon HMf was in a cluster alone. When three clusters only were used, HMf was added to the group of HMa, HUf and HUa. These results confirm that there is little variation within the Alarka site and that the pedons on the lowest deposit in Hidden Valley are more similar to pedons in the Alarka and Bradley site than to the other four deposits in Hidden Valley. 10,4 Cooolusions The variation of Fed/Fet with depth, in both saprolitic and colluvial soils, is irregular. However, the results of this chapter corroborate conclusions of earlier chapters. The pedons of the Alarka site have relatively small Fed/Fet ratios, which generally are associated with relatively young soils. The pedons at the Hidden Valley site have Fed/Fet ratios characteristic of older soils. Based on the Fed/Fet ratio, the relative age of the soils at the Bradley site appears to be intermediate to those of the other two sites. The within-site variation at Alarka is smaller than the within-site variation at Hidden Valley and the between-site variation of the three sites. At the Alarka and Bradley sites there is no evidence that the higher deposits are older. In Hidden Valley the lowest deposit has lower Fed/Fet values and therefore appears to be younger. For the Bradley and Hidden Valley sites this conclusion is consistent with the results of chapters 8 and 9. The highest Fed/Fet value was reached in HMf, one of the saprolitic soils. As was the case for the clast weathering and soil development indices, conclusions regarding Fe species ratios have to be considered as partial. They will be combined with other results to come to a firmer end result. 178 11 DISCRIMINANT ANALYSES 11.1 Literature review Qualitative evaluation of relative-age characteristics and data has been the classical approach to the grouping of geomorphic sites (e. g. Sharp, 1969). More recently, various multivariate statistical operations have been used to classify relative-age data into meaningful groups. Miller (1979) used custom-built software to develop an age classification of stations according to six relative-age dating criteria. Mills (1982) applied discriminant analysis to relative-age data from three sites in the Southern Blue Ridge but then evaluated the results subjectively. Switzer et al. (1988) used existing absolute age information and maximum likelihood classification to establish calibration curves and to estimate the age of undated soils. The majority of statistical classifications, however, have been carried out with principal component (PCA), cluster and discriminant analyses. A paper by Dowdeswell and Morris (1983) was one of the first to review these three methods. The authors argued that PCA was an important preliminary step to the grouping of sites through cluster analysis. Advantages were thought to be that PCA routines standardized scores, filters noise and produces orthogonal (independent) new variables. Harbor (1986) agreed with these advantages but pointed out that a major assumption of PCA was that each original variable is normally distributed. In most relative- age studies, however, the ages of deposits are expected to be concentrated in distinct groups that represent different geomorphic surfaces. Therefore, most relative-age data sets will not have a normal distribution. Cluster analysis is usually used to group sites on the basis of PCA scores or standardized values of the relative-age indicators. Most cluster routines combine sites into a hierarchy of groups of decreasing similarity (Dowdeswell and Morris, 1983). A visual inspection of the classification is performed to select the most appropriate level of 179 clustering. If there is no a priori reason to assert that all groups have the same level of internal variance, clusters from different levels may be chosen for further interpretation (Dowdeswell and Morris, 1983). Because cluster algorithms are sensitive to small changes in the data, the technique should be accompanied by a subjective evaluation of the results (Harbor, 1986). The quality of clustering at any level can be tested with discriminant analysis. Discriminant analysis can be considered to be the opposite of clustering because it uses existing groups and constructs functions (new variables) that produce maximum difference between those groups (Davis, 1986). The statistical significance of the separation between the groups produced by the discriminant functions, and thus the quality of the initial grouping, can be assessed by Wilks's lambda (Dowdeswell and Morris, 1983) and Hotelling's T2 test (Davis, 1986). Harbor (1986) submitted, however, that there is a circularity in the testing of the output of cluster analysis by discriminant analysis. He contended that, because of purely mathematical reasons, discriminant analysis always gives a high significance between groups derived from cluster analysis when the data set conforms to all assumptions of the methods. Harbor (1986), therefore, stated that the only valid means of using discriminant analysis with groups of relative-age data is to use one set of data to generate the clusters and another to test them, or to use groups based on personal judgment. In other words, discriminant analysis can be used as a quality check for a non statistical grouping of data into discrete classes. 11 Met ds 11.2.1 Construction of relative-age groups Not all relative-age indicators used in this study are normally distributed (Lilliefors test, or = 10 %). Therefore, it seemed best not to use PCA, and subsequently cluster 1 80 analysis, to group the pedons. It seemed better to use subjective but physically meaningful data, rather than objective but statistically inappropriate data, for further grouping and processing. Furthermore, considering the caution by Harbor (1986) about the appropriate use of dicriminant analyses, I chose to group the pedons based on the findings of the previous three chapters. All relative-age proxies indicated that, in general, the youngest deposits were present in Alarka and the oldest in Hidden Valley. Within the Alarka site clast weathering data clearly pointed to a greater age for the highest deposit (AUf, AUa) than for the lowest. The soil development indices seemed to confirm this for AUf but not for AUa. The iron species ratio showed very similar values for all Alarka pedons. These observations are inconclusive but because of the strong evidence from clast weathering, the pedons of the Alarka site were separated in two groups. Pedons ALfl and ALf2 were classified in the youngest group (G1) and AUf and AUa were the first members of a slightly older group (G2). Within the Hidden Valley site all relative-age indicators (clast weathering, soil development, iron species ratio) pointed to a younger age for the pedons on the lowest deposit (HLf, HLa), as compared to the higher pedons. The cluster analyses reported in the previous three chapters systematically grouped HLf with the pedons of the Alarka site. Pedon HLf was therefore classified in G2. Because of the similarity of HLa to HLf, it too was classified in G2. The other four Hidden Valley pedons (HMf, HMa, HUf, HUa) were assigned to a separate group (G3) that may contain the oldest soils. At the Bradley site there is no apparent age difference between deposits at different topographic levels. The clast weathering and soil development data indicate that the ages of the three Bradley deposits may be similar. It was therefore decided to classify the three Bradley pedons together. The results from chapters 8 and 10 show that they are more closely associated with G2 than with G] or G3. It would be possible to assign a new group to these soils. However, it is not clear whether this new group would be slightly younger 181 or older than G2. Because the resolution of the employed methods probably does not permit to make this distinction, the soils of the Bradley site were classified in group G2. 11.2.2 Data analysis Once the three groups above were delineated, three variables were excluded from the discriminant analysis to avoid redundancy. In chapter 8 the number of clasts in three weathering classes was expressed as a percentage of the total number of clasts. This implies that the value of the third class is completely determined by the outcome for the two other classes. One of the classes (medium weathered) was therefore eliminated from the data set. The three rubification indices (chapter 9) had very similar trends. To avoid an undue influence of the color as a relative-age indicator, one of the indices was removed from the data. I chose to remove Buntley-Westin index because it was developed for grassland soils, which have a morphology that is quite different from the soils studied here. However, the similarity of the three color indices (Figure 34) indicate that removing any one of them would have been acceptable. The percent clay in the most developed B horizon and the clay mass in the most developed B horizon both provided very similar trends as well (chapter 9). One of these two variables, percent clay, was therefore excluded from the data set. The seven variables remaining (two weathering classes, two color indices, two parameters for clay content and Fed/Fet) and a dummy variable for the group assignment were then submitted to the discriminant analysis routine of SYSTAT. Mathematically, the computation of the discriminant functions is similar to that of principal components. The discriminant functions can be considered as new variables which maximize between group variance and minimize within group variance. To visually represent the groupings as described by the new variables, the scores for the discriminant functions were plotted on a graph (Figure 38). The scores are the new values of the pedons for the new variables. 182 11.3 Rosults and oiscossion The grouping explained above (G1, G2, G3) was described by the discriminant functions in a statistically significant manner (W ilks's lambda, T2, or = 1 %), indicating that there is not only a subjective, physical base for the assignment of the pedons to their respective groups, but that there also is an objective, statistical justification to do so as well. Figure 38 illustrates the quality of the classification. Discriminant function 1 by itself could provide a good means of separating the three groups. The scores of G3 on function 1 are clearly different from the scores of the other groups. Although the scores of G1 and G2 on function 1 are distinct, these two groups can be better separated when the discriminating power of function 2 is added. The graph also shows that although there is considerable variation in group G2, splitting it in two or more distinct groups would be difficult. This confirms the subjective notion, mentioned in section 11.2.1, that subdividing G2 would not be appropriate. This does not imply that the ages of the pedons in G2 are equal; they obviously are not. It implies that the ages are sufficiently similar to be unseparable with the methods used. As a test, three other grouping schemes were subjected to discriminant analysis. In the first, the saprolitic soils were assigned to a separate group. In the second, the Bradley site was classified separately, and in the third test HLf was added to G1. Separation of the new groups in the three tests was not achieved by the discriminant analysis with the statistical confidence observed in the initial run. Also the plots of the scores showed that the discriminating power of the functions was less than in the original run. This indicates that the statistically most appropriate grouping was chosen in the initial classification. 183 mm 9:9“. .>_co 8:06 Lo 5:85:53 82:35 .2 9m mmma___m ”902 F 8:05.: EmEEtomE m; or m o m- o T m T T “ i\\l/ .1 n n m- m3< \ :1 9m. .. m- :m m 3:. \3/ .. 7 x L . \ U. _ 54.. L. N0 \\ / w. L.. .\ *3 II /.. .. o w mam- . Sruzrn J. m / \x , .5 .. F m. /../ \ .l‘. w --l / LEII \. .. N W. \ 5.27. m0 / .. m a _, $2- 8 L .\ . v mmzw __m 6908 EmEEcomE. 184 11. Esti ati n f solut a es Studies in the Southern Blue Ridge and in the Blue Ridge of Virginia have tried to relate relative-age data to absolute ages (Mills, 1983; Whittecar and Ryter, 1992; Leigh, 1995). This was accomplished by comparing soil morphological properties and relative ages to characteristics of dated soils. This kind of comparison is tenuous and probably can only provide order of magnitude estimates of the ages. Nevertheless, even very approximate numerical ages could help evaluate the results of the cosmogenic radiosiotope dating discussed in the part IV. Soils with hues redder than 2.5YR were inferred to be older than 100 ka by Whittecar and Ryter (1992) and Leigh (1995), and about 375 ka by Mills (1983). In the Valley and Ridge of Virginia, Mills (1988) estimated soils with similar colors to be >600 ka old. These soils had well developed argillic horizons (Whittecar and Ryter, 1992), with >45 % clay (Mills, 1983). Since pedons HMf, HMa, HUf and HUa have hues of 2.5YR and 10R, well developed argillic horizons and >40 % clay in at least one B subhorizon, the age of the saprolitic soils and of the highest deposit at the Hidden Valley site is estimated at >100 ka, and even could be >600 ka. Soils in the Southern Blue Ridge that were dated by 14C to 13 ka were Alfisol-like, and had about 20 % clay and a 10YR color in the B horizon (Leigh, 1995). Soils in the Blue Ridge of Virginia that were estimated to be about 10 ka by Whittecar and Ryter (1992) had weak argillic horizons, many unweathered clasts, and about 25 % clay and a 5YR hue in the B horizon. Soils judged to be of the same age had about 20 % clay and 7.5 to 10 YR colors in the B horizon (Mills, 1983). These characteristics are comparable to those of the pedons at the Alarka and Bradley sites, which therefore could be of the order of 10 ka old. It would be inappropriate to further distinguish between these soils with this method. 185 11.5 Conclusions Evaluation of clast weathering, soil development and iron species ratios allowed for a classification of the debris flow deposits in three age-related groups. Discriminant functions were effective in distinguishing between the three groups of deposits. Testing of the discriminant analysis indicated that the grouping was statistically significant at a high level of confidence. Other tested groupings could not be accepted with the same level of confidence. This indicates that the proposed, physically based, qualitative grouping of the deposits is statistically the most appropriate. Based on comparison with soils of known age in the region, the absolute age of the youngest group (ALfl , ALf2) can be expected to be of the order of 10 ka. The oldest soils (HMf, HMa, HUf, HMa) may be >100 ka, and even >600 ka old. 1 86 12 CONCLUSIONS The results of the various relative-age proxies used in the previous chapters are in good agreement. They classified the deposits into uniform groups that were physically meaningful and statistically significant. The proxies for relative age indicate that the deposits of the three sites were emplaced at different periods in time. Generally, the debris flows at the Alarka site occurred more recently than those at the Hidden Valley site. The lowest deposits at these two sites appeared to be younger than the highest deposits. Other investigators working in the Blue Ridge have found a similar association between relative age of debris flow deposits and topographic position (Mills, 1981; 1983; Whittecar and Ryter, 1992). A distinction in age between the oldest (higher) deposit at Alarka and the youngest (lowest) deposit of Hidden Valley could not be made. The discriminant analysis showed that assigning a similar age to these deposits is statistically meaningful. The deposits at the Bradley site appear to be comparable in age to the higher deposit at Alarka and the lowest at Hidden Valley. The soil and weathering characteristics of the saprolite at the Hidden Valley site point to a greater age for the exposure of this saprolite than for any of the colluvial deposits. Most relative-age proxies show larger within-site variation at the Hidden Valley site than at the other two sites. If this inference is correct, it indicates that either there is a greater difference in age of the deposits at the Hidden Valley site or that as soils become older more variation is introduced in their development. Acceptance of the latter option would contradict the principle of converging pedogenetic pathways. Mills (1982) mentioned, however, that perhaps small differences in parent material and microenvironment of deposits may become more important with greater age and hence may introduce more variation. Absolute dating of the deposits could possibly resolve this contention. 187 Two soil properties discussed in chapter 6 conform to these age relationships. Clay mineralogy (Table 1 1) showed that more kaolinite was present at the Hidden Valley site than at the Alarka site and that the higher deposit at the Alarka site contained more kaolinite than the lower deposit. Because a high kaolinite content is usually associated with an advanced degree of weathering, and thus an older age, these observations strengthen the age relationships postulated above. The clasts at the Hidden Valley site are smaller and less abundant than at the Alarka site (Table 5, Figure 17). Although this may reflect the size and abundance of the clasts in the source material, it also could be the result of longer weathering of the Hidden Valley deposits. If, as postulated above, the paucity of clasts at the Hidden Valley site is the result of longer weathering, and was not inherited from the source material, the term "mud flow" assigned to these deposits in section 5.4.3 may not be appropriate. If these deposits originally contained large amounts of coarse fragments, "debris flow" may be the best designation after all. The soil properties used in this research as proxies for relative age may have been inherited, in part, from the source material of the debris flows. However, the close agreement of the results of the various, independent, relative-age indicators, suggests that inheritance did not affect these soil properties considerably. A study comparing pedogenic characteristics of source areas and deposits of very recent debris flows could establish the importance of inheritance for pedogenesis on debris flow colluvium more firmly. PART IV: CONCLUSIONS 188 189 Debris flow deposits are a common geomorphic phenomenon in the landscape of the Southern Blue Ridge of North Carolina. In this study, sedimentary, pedological and age characteristics of these debris flow deposits were examined. In spite of the widespread occurrence of the deposits, detailed investigations of these properties have not often been carried out in the region, nor have they been carried out frequently in other temperate to warm humid environments. Part of the relevance of this study, therefore, lies in the generation of characterization data of previously poorly understood landforms. In addition, this study was designed to help understand pedogenesis on debris flow deposits and to identify periods of increased landscape stability/instability in the region through an analysis of relative ages of the deposits. A reconnaissance survey showed that colluvial deposits occur commonly in many landscape positions throughout the Little Tennessee River basin. Debris flow deposits are most prevalent in relatively small, low order stream valleys with steeply-sloping backslopes. Typically, these deposits have a lobate topography, and large boulders, many more than a meter in diameter, can often be observed at their surface. Three sites with debris flow deposits were selected for further study based, among other criteria, upon the presence of a sequence of deposits at the site, degree of surface erosion, differences in degree of soil development in and between sites and the uniformity of the local bedrock. The sedimentary characteristics of the deposits show that most were formed by debris flow activity. They typically consist of massive, matrix supported diamicts, a facies often observed in debris flows by other researchers. In some soil pits of the Hidden Valley site very few clasts were encountered. These deposits are interpreted to be mud flow deposits. However, this site also had the oldest relative age and the paucity of coarse fragments may be due to prolonged weathering. Facies characteristics indicate that the viscosities of the debris flows at the three sites were different. The Alarka site appears to have experienced the most viscous flows, as evidenced by the fabric, lobate topography and thickness of the deposits. At the Hidden 190 Valley and Bradley sites more fluidal debris flows occurred. In the case of Hidden Valley this may be due to a finer texture of the source material. The finer texture may relate to more intense weathering of the source material, and/or could point to a greater stability for this basin. The Bradley site has the most heterogeneous sediments. Various sedimentary units are present in the three pits at this site. However, there is no evidence for prolonged surface exposure. This suggests that there were no major interruptions in the most recent depositional episode at Bradley. Particle size fractionation data show that the parent materials of the soils are very heterogeneous. This heterogeneity was reflected in many pedons by the presence of lithologic discontinuities. These lithologic discontinuities were in most cases interpreted as the result of internal variations within one flow, rather than the result of the stacking of consecutive flows. It is postulated that the internal variations represent different shear zones in the flows. The presence of shear zones is consistent with the results of the sedimentary fabric analysis. The observation of heterogeneous parent materials in debris flow deposits contrasts with those of Daniels et al. (1987) and Graham et al. (1990b) who observed homogenization in colluvial soils. Although these authors did not completely describe the origin of the colluvium they studied, this contrast indicates that pedogenesis is not uniform in all colluvial deposits of the region. The only parameter that shows a systematic spatial trend is the depth to maximum clay accumulation. Depth to maximum clay accumulation increases as elevation increases. This pattern holds for situations when two pedons are located on one deposit and also for two deposits within one site. The trend most likely reflects impeded vertical movement of vadose water in lowest topographic positions and age differences between the deposits. Shallow clay maxima on steep slopes and apparently incomplete soil profiles point to removal of surficial horizons by erosion at some locations. Two of the pedons that were studied in Hidden Valley have developed in saprolite. These pedons showed no evidence for lithologic discontinuities and contain more clay and 191 have more pronounced clay maxima than do colluvial soils. The difference in clay content and accumulation may in part be due to a greater age of the saprolitic soils. The clay mineralogy within every pedon varies little. Of the principal clay mineral groups, only 14 A minerals show a decreasing trend with depth. No consistent trends with depth were apparent for the other clay minerals. Within the Alarka site there was a small difference in the clay mineral assemblages between the lower and higher deposit. Larger differences exist between sites. These within and between site differences most probably relate to differences in age of the deposits. The qualitative and statistical analysis of the various relative-age proxies produced very consistent results. This consistency and the accordance of the results with sediment and soil characteristics, such as content and size of coarse fragments, and clay mineralogy, indicate that the results are dependable. It also suggests that the chosen variables are appropriate relative-age indicators for debris flow deposits in warm to temperate humid environments. This research, as well as studies by Mills ("1981; 1983) and Whittecar and Ryter (1992), shows that higher deposits are slightly older than lower deposits in the same area. In the Little Tennessee River basin the within-site variance in age is smaller than the between-site difference in age. In spite of the generally large between-site differences, one deposit at each of the three sites appeared to have a relative age that was very similar to that of one deposit at the other sites. This similarity in age may be due to an inadequate resolution of the applied relative-age dating methods. However, if real, the similarity would suggest that an external factor, operative across all the three sites, triggered the debris flows at approximately the same time. In the tectonically stable Southern Blue Ridge this factor is most likely to be climate (Mills, 1981; 1982b; Kochel and Johnson, 1984; Velbel, 1987). The two most recent deposits are present at the Alarka site. Deposits of comparable age were not observed at the other two sites. It is not likely that deposits of that age have been present at those other sites, they cannot be covered and would have been exposed to 192 less erosion than the older deposits. Their absence at the Hidden Valley and Bradley sites therefore suggests that climate alone cannot explain the occurrence of debris flows in the region. Local, intrinsic factors must affect the timing of debris flow activity. Further study of relative ages of debris flow deposits in the region could help to identify the relative importance of intrinsic and external factors that produce debris flows. Large numbers of deposits of the same relative age would point to a dominance of external over internal factors. Absolute ages would allow for temporal correlations between sites on a larger spatial scale and could clarify during which periods in geologic history the climate of the Southern Blue Ridge was most favorable for the formation of debris flows. Considering that materials datable with traditional methods, such as 14C, are absent from the studied deposits, it is suggested that methods based on the decay of cosmogenic radioisotopes may be the only viable dating approach in the near future. PART V: APPENDIXES APPENDIX 1 Generalized description of selected pedons of reconnaissance survey 193 194 Table 14: Field ' of selected of reconnaissance orrzon orst cm Munsell consistence fragments color none none CW’ e“! e“! conunon none none none none none none none none e“! Table 14: Field orrzon 'on of selected cm 195 Olst Munsell color multicolored multicolored of reconnaissance cont consistence fragments none ew none many none common common ew 196 Table 15: Particle size of selected of reconnaissance Pedon Horizon exture Sand Silt Clay VCS CS MS FS VFS of fine % % % % % % % % earth1 197 Table 15: Particle size anal of selected of reconnaissance cont' Pedon Horizon Texture Sand Silt Clay VCS CS MS FS VFS of fine % % % % % % % % earthl Abbreviations for texture class: 1 = loam, c = clay, cl = clay loam, 31 = sandy loam, scl = sandy clay loam. APPENDIX 2 Longitudinal profiles, Alarka and Bradley sites 198 199 mm 939“. D Em 0 Co 8282 L2 m 659; mom a E .8565 o 09. cor o: ONF oow ow om ov cm 0 . . . . r r . r r . . . t t . . . or- In, m- /ll/ m o // E ///| m: mzm 9.82 $65 353358.. 0 N w ‘uogreAala 200 ow 9:9“. E .8865 0mm oom omw cow 0 new 0 Co c0580. L8 9 939E mow ON- I I I T or- III! ITII or VIII om I I I I om. 1' U l I ov mzw >285 .265 3:635:04 . UJ ‘uogreAaB APPENDIX 3 Particle size fractionation data 201 202 0.: .98 «.3 3 «.m m.» «.m« «.8 o.« a; «8« Q: F.o« 0.3 as 4.8 m; _.m« «.8 «.m 5 6m« of m.o« m.«_ E ..m I: v.«« 9% «.« 3 .-«am «.3 :« of 2 me 3: «.8 4.3 «.« o; «-«;m I: «.t «.2 «.« «a «.3 tom «.5 m.« «4 Em 3 o..«« «.2 8.8 5. «.2 «.5 9mm «.4 m; m 3 mi 3: «.m me 3: «.8 Qt“ m.« E < 5< 93 o.o« 3: mi 3 «.2 «.8 «.«m m m «.« «0% «.«F :« v.2 «.m g «.8 95 Q; «.« «4 6m« «.m 2: t: «a 9m «.m« WK «.3 we 3. sm« «.2 «.3 4.: E 3 «.2 «.8 95 3. m « m as «.8 Q: E 3. «.2 «.8 98 _.« 3 «< 3 0.2 «.3 E g «.t «.8 ..mm « m «.« 2 «:< «.o 9: «.5 «.3 m.«_ v.2 «.8 «.3 «.«_ s m «0mm «.w «.5 «.t «.3 o.«F sm m._« 28 m S o a 5mm «.8 Z: 3: «.m «m m? «.m« 93 o S I: E« «.« «.2 me: «8 1m 3: 98 98 o; «.o m 3 0.: «.2 we «.« fit «.8 «.3 We 90 m< t: «.3 m.«_ «.0 gm of 0.5 o.«m «.« v « < 2.2 Er «y 5:8 8: so as. 2 ES .99 so .x. mus _ mm m2 _ mo _ wo> H .86 _ Em _ 2mm 5.: v-« 7:5 m4 .8on 88m. 8% ..m .98 36 225a um: 29...» 203 «.« 3.3 «.3 «.« 3..« «.3« «.«« «.«« «.3 3 « «o« «.« 3.3« «.3 «.« «.3. «.«« «.«3 «.«« «.3 «.« o3. «.« «.«3 «.« «.3. 3.3 3.3« «.«« «.3.« « « «.« «o« «.« «.3 «.3 «.33 «.« «.«« «.«3 «.3« «.« «.« o« «.« 3.3 «.33 3.« 3.« «.«« «.«« «.3. 3..« «.« o 3.3. «.« «.« «.« «.3 «.«« 3.3.« «.«« « « «.« om «.« «.« 3.« «.« «.« «.«3. «.«« «.«« 3 « «.« «3m «.« «.« «.« «.3. «.3. «.33. «.«« «.«« «.« 3.« 33m «.« «.« «.« «.« 3.« «.3.« «.«« «.«« «.« «.« Sm «.« «.«3 3.33 «.« «.3. 3..«« «.«« «3.3. 3 « 3.« < S: «.33 «.«« «.3 «.« «.« «.« «.«« «.«« «.« « « o« «.« «.«« «.3.« «.«3 « 3 « « «.33 «.«« «.« «.3. «o« «.3 «.«« «.«3 «..« «.« «.3 3.3 «.«« 3 « «.« «o «.« 3.3 «.«3 «.« «.« «.«« «.«« «.«3. 3 3 3.« 8 «.« «.3 «.«3 «.« «.« «.«« «.«« «.«3. 3 « «.« 33m «.« «.3 «.«3 «.« «.v «.3 «.«« «.«« « « «.« < :1 «.«3 «.3 «.«3 «.« « 3. «.«3 3.«« «.«« 3 3 «.« o« «.33 «.3 «.«3 «.« « « «.« «.«« 3.«« 3 « «.« o« «.«3 «.«« «.«3 «.« 3 « «.« «.«« «.«« «.« «.3 «om« «.« «.«« «.«3 «.« 3 « «.« «.«« «.«« «.3. 3. « 6m« «.3 «.«« «.33 «.« « « «.3 «.3.« «.«« 3.« «.3 3m «.« «.«3 3.33 3.« 3 « «.«« «.«« «.«« « « «.« 3.35 3..« «.«3 «.3 «.« « « «.«« «.«« «.«3. «.« « 3 «ism «.« «.«3 3.33 «.« « 3. «.3 «.«« «.«« «.« 3 3 < «3. es «vv £8 «cc 30 .x. as 30 o\.. «”5 _ mm «s. _ «o _ wo> _ >90 _ 33w _ ««.«« EE 3.-« E: «-3. .855: Sump 33353 8% =3. «3% «Na «.«EE ”«3 «33 204 «.« «.3« «.«3 «.«3 «.« «.33 «.«3 «.«« «.« «.« eo «.« 3.3 «.«3 «.33 3.3. «.«« «.«3 «.«« 3.« «.« 3-««5 «.« «.«« «.«3 «.« «.3 «3.« «.«3 «.«« «.« «.« «-««B 3..« «.3 «.«3 «.« «.3 «.«« 3.«3 «.«« 3.« «.« 3-«3m «.« «.3 3.3 «.« «.« 3..«« «.3« «.«3. 3.« 3.« «-«« «.« «.«3 3.33 . «.« «.« «.«« «.«« «.«« «.3 «.« ««m «.« «.«3 «.33 «.« «.« «.«« «.«« 3.«3. «.« «.3. «3m «.« «.«3 «.3 «.« «.« «.«« «.«« «.«3. «.« «.« «< «.« «.3 «.«3 «.« «.« «.«« «.«« «.«« «.3 «.« «< «.« «.3 «.33 «.« «.« «.«« «.«« «.«3. «.3 «.« < «.2... «.« «.« «.« «.« «.« «.«3. «.«« «.«« 3.« «.« 8 3.« «.«3 «.« «.« «.3. «.33. «.«3 «.«3. «.« 3.« 3.03« «.« «.«3 3.« «.« «.« «.«3. «.«3 «.33. «.« «.« «-o3m «.« «.« «.« «.« «.« «.«3. «.«« «.«« «.3 «.« «3m «.« «.« «.« «.« «.« «.3« «.«3 «.«« «.« «.« 33m «.« 3.3 «.«3 3.« «.« «.«« «.«« «.«3. 3.« 3.« «< «.« «.«3 «.«3 «.« «.« «.«3 «.«« 3.«.. «.3 «.3 < 32: «ES va ztmm 0:: 30 ..\o 29338.92 30 .x. «”5 _ «u. _ «s. _ «o _ «o> — «go _ 3__« _ ««.«« EE «-« _es «-3. .8551 «8mm 6.383 «3.« ...m «.«« «am 22:33 ”«3 «.««3 205 «.« «.«« « 3 « « «.« «.«3 «.«« «.3« « «3 « « ¥««o« «.«3 «.«« 3 «3 3 « «.« «.3 «.3« «.«« «.3 «.« 38 «.3 3.3 3.« «.« «.3. «.«« «.«« 3.3. «.3 «.« am «.« «.«3 « « «.« «.« «.«« «.«« «.«3. I. «.« < :m «.« «.«3 3.3 «.« «.3. «.«« v.3« 3.3. 3.3 «.« 3m« «.3. «.«3 3.33 «.« «.« «.«« «.«« «3.3. «.« 3.« 3.2m« «.« 3.3 «.3 «.33 «.3. «.«3 «.«« 3.3« «.« «.« «.sm« «.« «.«3 «.«3 «.3 3.« «.3 3.«« «.«« «.« «.« «am «.« «.«3 3.3 «.«3 «.3. «.«3 «.«« «.«« «.3 «.« 3am «.« 3..«3 «.«3 «.« «.« 3.«« «.«« 3.3« «.« «.« _ ««.o L 3__« F ««.«« EE 3.-« _EE «-3. .828: :88 6.383 8% =« «3.«« «Na «.«Eaa ”«3 «33 co.«_._o: «88 «5 c_ Egon E9336 3.« 3.9.32 «waEmm 0233 2 .29 N. 3.« 3- . v6 adv mdF FA: adv NSF 3mm wém N4": N4: TmBmm 206 Ndw wum— o.w m.\. 06 N.mF . m.Nm adv o.m N.m N-m>>mm ma; «41 m6 m.v N6 NSF Ndm wtmv m.N o; Esmv m6 m.ow 0.: wfi 9w N.m_. «.5 5mm m.w 06 mgmm ma: m.m_. Nd o.m 0.0 Tm: Adm Ndv v6 N.N NBmN «INF #4; OK 9v m4. mtmw No.3. ®.mv m.N TN :sm N.o_. Ndr o.w m4. mé VKF m.mm 56v NS 0.0 Q< mDm o.w adv 3: m.o_. m6 v.33: 0.0m m.om N4; N4; vgmw «.«3 «.«3 « « « « «.« «.«« «.«« «.«« 3.3 «.« «;m« 3.« 3.3 «.33 «.«3 «.«3 «.«3 «.«« «.«« «.« «.« «;m« 3.« «.«3 «.« «.« «.« «.«3 «.«« «.3« «.« «.« 3;« 3.« «.3 «.« «.« «.« «.«3 «.«« «.«3. 3.3. 3.« m «.« «.«3 «.« «.« «.3. «.«3 «.«« «.«3. «.« «.« «< Sm «E3: «vq «:8 «3.« 30 ..«o «.32 30 «x. «Us _ mm _ «s. _ «o _ «o> _ «so _ Em _ «3.«« EE 3.-« EE «-3. .3555: «88 «383 «««w =« ««.«« «Na «38 ”«3 «BS Table 17: Clayfree article size data, all sites 207 Pedon Horizon sand I silt I VCS] (B I NB 1 FS 1 VFS * % offfine earth 1<2 mm) (Clal ree) ALf1 A 62.6 37.4 3.6 7.5 15.1 23.1 13.3 AE 55.7 44.3 3.8 7.8 13.6 20.5 10.0 2E 62.2 37.8 6.6 7.7 14.1 22.3 11.4 ZBt 68.2 31.8 6.6 10.7 16.6 23.0 11.4 3801 76.3 23.7 13.2 15.4 18.9 19.6 9.1 3802 88.1 11.9 14.8 21.1 24.6 20.3 7.2 ALf2 A1 64.1 35.9 4.4 8.6 18.0 23.0 10.2 A2 62.3 37.7 5.5 8.7 14.7 22.5 10.9 E 62.8 37.2 6.8 8.7 14.0 20.6 12.7 ZBt 66.7 33.3 7.5 9.2 14.9 24.2 11.0 2801 74.2 25.8 7.0 10.7 16.1 25.7 14.8 3802 78.1 21.9 10.2 13.8 18.7 23.8 11.5 AUf A 58.7 41.3 5.6 8.4 13.3 21.4 9.9 E 62.9 37.1 5.6 7.9 13.9 27.3 8.2 8w1 63.2 36.8 6.5 8.8 13.9 21.7 12.4 Bw2-2 65.1 34.9 5.1 8.7 14.1 25.2 12.0 Bw2-1 72.3 27.7 6.3 8.8 15.8 25.4 16.1 2801 71.4 28.6 10.3 10.9 15.3 22.0 12.9 2802 69.5 30.5 6.3 9.6 16.1 25.5 11.9 AUa A 62.1 37.9 5.2 8.2 13.9 23.7 11.1 Bw-2 62.1 37.9 7.8 9.6 14.1 20.8 9.8 Bw-1 64.1 35.9 6.5 7.9 14.3 22.9 12.4 Bt 70.0 30.0 8.1 9.4 14.7 25.1 12.7 2801 59.0 41.0 6.7 8.7 13.2 22.0 8.3 2802 68.9 31.1 6.7 10.2 15.3 22.5 14.2 20 69.6 30.4 10.6 10.8 14.6 20.8 12.8 30 65.7 34.3 5.4 9.5 14.4 21.9 14.5 HLf A 63.2 36.8 5.0 11.5 16.5 20.6 9.7 Bw 60.3 39.7 4.5 12.0 15.8 18.7 9.3 80 63.4 36.6 4.4 10.8 17.3 20.7 10.2 _01 84.0 16.0 4.4 10.4 22.1 34.4 12.7 2293 87.6 12.4 11.0 20.5 26.4 22.2 7.5 30 63.2 36.8 3.9 10.8 15.2 21.5 12.0 Table 17: Clayfree 208 particle size data, all sites (cont'd) Pedon Horizon sand I silt I VCS I (B I NB I PS I VFS 3 °/o of fine earth (<2 mm) (clar ree) HLa A 57.4 42.6 5.9 11.6 14.5 17.7 7.7 B/A 44.0 56.0 3.2 5.6 11.8 14.7 8.7 811 50.7 49.3 7.3 7.3 11.4 16.2 8.6 Bt2 45.0 55.0 1.3 5.3 11.1 16.8 10.5 H: 46.5 53.5 2.3 8.8 13.7 15.2 6.4 C 63.6 36.4 3.0 11.7 16.8 20.7 11.4 2C 79.4 20.6 6.4 15.3 21.7 25.2 10.8 301 49.9 50.1 1.5 6.8 10.0 17.7 13.8 4C 75.4 24.6 5.7 11.7 19.0 27.7 11.3 509 73.6 26.4 6.9 11.1 18.3 24.7 12.6 HMf A 60.0 40.0 6.4 9.6 15.8 19.8 8.4 AB 59.2 40.8 5.0 9.7 16.8 19.2 8.4 311 58.9 41.1 6.2 10.8 14.5 19.4 7.9 812 61.5 38.5 6.6 12.6 14.2 16.9 11.3 B/C-2 69.8 30.2 9.4 14.5 13.7 21.4 10.8 B/C-1 68.4 31.6 6.8 1517 16.4 18.6 11.0 H) 61.4 38.6 8.4 13.2 13.0 18.7 8.2 HMa A 62.7 37.3 4.2 10.1 18.4 21.8 8.3 AB 60.9 39.1 4.3 8.8 16.2 22.8 8.8 Ab 64.8 35.2 5.4 10.4 17.7 21.4 9.9 51b 57.2 42.8 4.8 7.5 14.5 19.8 10.5 E2b 60.0 40.0 4.0 7.4 15.9 23.1 9.7 Btb-2 66.0 34.0 3.7 9.1 16.5 24.0 12.8 Btb-1 71.3 28.7 2.7 9.8 19.2 26.3 13.3 C/Bb-2 79.9 20.1 2.6 11.6 20.9 31.5 13.3 C/Bb-1 82.6 17.4 6.2 16.1 21.7 25.8 12.6 Clb 85.8 14.2 10.2 19.5 22.1 24.4 9.6 HUf A 55.7 44.3 4.3 10.8 15.4 17.5 7.7 BA 57.9 42.1 4.0 10.8 15.3 18.9 8.9 BW1 61.8 38.2 4.8 13.6 16.3 19.2 7.9 28w2-2 65.3 34.7 5.3 14.6 17.4 19.5 8.4 2Bw2-1 63.7 36.3 4.2 13.9 16.5 18.9 10.1 3Bt3-2 69.9 30.1 9.5 13.3 16.0 21.6 9.5 3Bt3-1 67.9 32.1 11.6 11.7 15.1 21.5 8.0 4Bt4 65.8 34.2 7.9 10.7 16.0 19.6 11.6 209 Table 17: Clayfree particle size data, all sites (cont'd) Pedon Horizon sandI silt | vcs | <8 I MS | F8 | vr=s % of fine earth j<2 mm) (clar ree) 3 HUa A 63.7 36.3 5.0 10.1 18.2 21.1 9.3 BA 65.7 34.3 5.0 10.6 19.6 22.4 8.0 Bw1 67.9 32.1 5.7 12.9 19.8 20.1 9.4 8w2 68.2 31.8 7.1 15.9 19.2 18.0 7.9 28w3 60.8 39.2 4.8 13.3 17.0 16.8 8.9 38w4 60.2 39.8 7.2 10.6 15.5 21.4 5.4 SBt 65.2 34.8 6.5 11.3 16.9 21.0 9.4 BLf A 54.9 45.1 6.7 6.5 10.0 19.6 12.2 8w 56.4 43.6 6.1 7.4 11.0 17.6 14.4 Q‘L 63.3 36.7 4.5 7.1 11.8 27.4 12.3 2gL2 70.1 29.9 7.6 10.8 15.6 27.0 9.1 BLa Ap 56.5 43.5 5.5 5.8 9.6 23.2 12.4 E 51.7 48.3 5.8 5.5 8.3 17.5 14.7 8w1 55.5 44.5 6.9 5.8 9.5 17.8 15.5 28w2 64.2 35.8 7.6 9.0 12.7 23.4 11.3 38w3 52.4 47.6 5.1 5.5 8.3 17.2 16.3 48w4 59.8 40.2 12.0 8.9 9.8 16.6 12.4 BUa Ap 51.8 48.2 5.5 6.7 10.3 18.2 11.1 8w1 55.9 44.1 10.8 8.8 10.1 15.9 10.3 28w2 63.9 36.1 12.1 11.6 11.6 18.6 10.1 38w3 67.4 32.6 12.2 12.8 13.8 17.8 10.8 4Bw4 48.9 51.1 2.8 4.6 8.2 19.9 13.2 58w5-2 65.0 35.0 11.5 12.1 13.1 18.4 9.9 58w5-1 72.1 27.9 17.0 14.8 14.8 16.9 8.6 APPENDIX 4 Sorting and bulk density data 210 211 Table 18: SortinLand bulk density data, all sites Pedon Horizon 01 I 03 sortin; bulk density estimated? mm 9 cm-3 ALf1 A 0.017 0.230 3.68 0.97 n AE 0.017 0.220 3.60 1.18 n E 0.017 0.230 3.68 1.38 y 231 0.018 0.290 4.01 1.38 y 3301 0.045 0.550 3.50 1.38 y 3BC2 0.070 0.660 3.07 1.38 y ALf2 A1 0.017 0.260 3.91 0.98 n A2 0.017 0.240 3.76 1.01 n E 0.015 0.240 4.00 1.05 n 281 0.002 0.240 10.95 1.20 y 2BC1 0.025 0.300 3.46 1.20 y 3BC2 0.038 0.430 3.36 1.20 y AUf A 0.016 0.230 3.79 0.80 n E 0.020 0.240 3.46 1.25 n BW1 0.017 0.240 3.76 1.45 y BW2-2 0.022 0.240 3.30 1.45 y BW2-1 0.017 0.240 3.76 1.45 y 2BC1 0.040 0.490 3.50 1.45 y 28C2 0.040 0.330 2.87 1.45 y AUa A 0.013 0.230 4.21 1.10 n BW-2 0.002 0.240 10.95 1.18 n BW-1 0.009 0.230 5.06 1.38 y 31 0.018 0.270 3.87 1.38 y 2BC1 0.027 0.260 3.10 1.38 y 2BC2 0.036 0.320 2.98 1.38 y 2C 0.036 0.380 3.25 1.38 y 3C 0.027 0.260 3.10 1.38 y HLf A 0.012 0.270 4.74 1.38 n BW 0.005 0.260 7.21 1.52 n H: 0.005 0.260 7.21 1.53 n 99 0.040 0.330 2.87 1.57 n 2C9 0.100 0.610 2.47 1.65 y 3C 0.033 0.280 2.91 1.65 y 212 : Sorting and bulk density data,all sites (cont'd) Table 18 Pedon Horizon Q1 I Q3 sorting bulk density estimated? mm 9 cm-3 HLa A 0.005 0.240 6.93 1.45 n B/A 0.004 0.090 4.74 1.48 n 811 0.002 0.125 7.54 1.44 n Bt2 0.002 0.060 5.35 1.40 n H) 0.003 0.125 7.07 1.31 n C 0.003 0.220 8.86 1.36 n 20 0.006 0.370 7.85 1.48 n 3C}; 0.004 0.120 5.48 1.35 n 4C 0.008 0.300 6.12 1.44 n 5C9 0.012 0.300 5.00 1.45 y HMf A 0.015 0.260 4.16 1.29 n AB 0.003 0.240 8.94 1.28 n Bt1 0.002 0.120 7.75 1.49 n Bt2 0.002 0.150 8.66 1.54 n BIC-2 0.002 0.160 8.94 1.37 n BIC-1 0.002 0.230 10.72 1.36 n a) 0.002 0.230 10.72 1.41 n HMa A 0.003 0.210 9.17 1.31 n AB 0.003 0.200 8.94 1.29 n Ab 0.003 0.250 9.13 1.43 n E1b 0.006 0.210 5.92 1.45 n E2b 0.003 0.210 8.37 1.58 n Btb-2 0.002 0.170 9.22 1.41 n Bib-1 0.002 0.170 9.22 1.41 n C/Bb-2 0.003 0.240 9.80 1.35 n C/Bb-1 0.003 0.330 11.49 1.35 n Crb 0.066 0.520 2.81 1.45 n HUf A 0.003 0.220 8.56 1.33 n BA 0.003 0.180 8.49 1.45 n Bw1 0.002 0.190 9.75 1.38 n 2Bw2-2 0.003 0.190 8.72 1.44 n 28w2-1 0.003 0.210 9.17 1.44 n 3Bt3-2 0.002 0.220 10.49 1.47 n 3813-1 0.002 0.220 10.49 1.47 n 4Bt4 0.002 0.200 10.00 1.54 n 213 Table 18: SortinLand bulk densiy data,all sites (cont'd) Pedon Horizon Q1 I 03 sorting bulk density estimated? mm gem-3 l-Ma A 0.010 0.290 5.39 1.16 n BA 0.008 0.280 5.92 1.47 11 BM 0.016 0.340 4.61 1.50 n 8w2 0.026 0.420 4.02 1.55 n 28w3 0.010 0.330 5.74 1.62 n 38w4 0.003 0.210 9.17 1.66 n 381 0.003 0.240 8.94 1.56 BLf A 0.011 0.190 4.16 1.39 n Bw 0.005 0.190 6.16 1.58 n 091 0.022 0.220 3.16 1.46 n 2093 0.031 0.330 3.26 1.50 y BLa Ap 0.015 0.190 3.56 1.27 n E 0.015 0.240 4.00 1.52 n Bw1 0.015 0.330 4.69 1.58 n 2Bw2 0.027 0.410 3.90 1.37 n 38w3 0.010 0.150 3.87 1.59 n 4Bw4 0.027 0.400 3.85 1.57 n BUa Ap 0.016 0.200 3.54 1.15 n Bw1 0.016 0.180 3.35 1.38 n 28w2 0.022 0.200 3.02 1.47 n 38w3 0.026 0.260 3.16 1.46 n 4Bw4 0.014 0.170 3.48 1.45 n SBw5-2 0.022 0.450 4.52 1.48 n 58w5-1 0.016 0.250 3.95 1.48 n APPENDIX 5 Stereograms for Clast a-axis 214 215 Stereograms of long axis of 018318, lower deposits at Alarka site ALfl Pro ectian . . . . . . . . . Schmidt Num6er oF Sample Paints 30 151 Eigenvalue . 0 580 2nd Eigenvalue 0 251 3rd Ei envalue . . . . . . 0.169 LN l E / E2 1 . 0 839 LN 1 E2 / E3 ) . 0 396 (LNlEl/EZl] / 1LNlE2/E3ll . 2.116 S herical variance . 0 4456 Rgar . . 0 5544 o Clast a-ax1s orientation ALf2 Pro ectian . . . . , Schmidt Num6er aF Sample Points 30 let Eigenvalue . 0 624 2nd Eigenvalue . 0 250 3rd Ei envalue . . . . . . 0.125 LN l E / £2 1 . 0 915 LN 1 E2 / E3 1 . . . . . 0.690 (LNlE1/E21] / [LNlEZ/EEll . 1 325 S herical variance . 0 6618 R or . . . . . . . . . . . . 0.3382 . Clast a-axi: orientation . Figure 41 216 Stereograms of long axis of clasts, higher deposit at Alarka site N AUf Pro ectron Schmidt Num6er of Sample Paints . . 30 151 Eigenvalue . . . . . . . 0.550 2nd Eigenvalue . 0.313 3rd E1 envolue . . . . . . . 0.137 LN I E / E2 1 . 0.564 LN ( E2 / E3 1 . 0.827 lLNiEl/EZl] / lLNiEZ/E3ll . 0.681 S herical variance . 0.5727 R or . . 0.4273 + Clast a-axrs orientation AUa Pro ection . . . . . . . . . Schmidt Num er oF Sample Paint: 30 lst Eigenvalue ’ . 0.505 2nd Eigenvalue . 0.276 3rd E1 envalue . . . . . . . 0.219 LN ( E / E2 1 . 0 606 LN 1 E2 / E3 1 . 0.228 lLNlEl/EZl] / lLNlE2/E3ll . 2.657 S herical variance . 0.5330 850r . . 0.4670 0 Clast a-axis orientation Figure 42 217 Stereogram of long axis of clasts, Hidden Valley site HUf Pro ectian Schmidt Num6er oF Sample Points . . 29 lst Eigenvalue . . . . . , . 0.469 2nd Eigenvalue . 0 352 3rd Ei envalue . 0 179 LN ( E / E2 1 . . . . . 0.287 LN 1 E2 / £3 1 . 0 678 (LNlEl/EZl] / lLNlEZ/E3ll . 0.424 S herical variance 0 4080 REar 0 5920 o Clast a-aXis orientation Figure 43 218 Stereograms of long axis of clasts, Bradley site BLa Pro ection ..... . . . . Schmidt Num er oF Sample Points 1 lst Eigenvalue ....... 0.592 2nd Eigenvalue . 0 315 3rd Ei envalue . 0 093 LN 1 E / E2 1 . . . . . . . 0.630 LN 1 E2 / E3 1 . 1 221 1LN1El/E211 / 1LN1E2/E3ll . 0.516 S herical variance 0 6406 850r . 0 3594 4 Clast a-oxis orientation BUa Pro ection . . . . . . . . . Schmidt Num er of Sample Pr 113 31 lst Eigenvalue . 0 500 2nd Eigenvalue . 0 354 3rd Ei envalue . . . , . . . 0.146 LN 1 E / E2 1 . 0 347 LN 1 E2 / E3 1 . 0 884 1LN1E1/E21] / 1LN1E2/E3ll 0.392 S hericol variance 0 6157 Rgar . 0 3843 a Clast o-axis orientation Figure 44 22159 Stereograms of long axis of clasts, subpopl 1LN1E1/EZ)] / iLNiEZ/Eaii‘ Spherical variance 8 or [\Iafl Pro ectian . . . . . . . . Schmidt Num6er of Sample Paints . 16 Mean Lineatian Azimuth 76.2 Mean Lineatian Plunge . . . 30.4 lst Eigenvalue . 0 748 2nd Eigenvalue 0 152 3rd Ei envalue . . 0.100 LN 1 E / E2 1 1 592 LN 1 E2 / E3 1 . 9.421 0 0 o Clast o-axis orientation lLNlEl/EZl] / (LNiEZ/EBJI S erical variance . . R or IXIJEZ Pro ection . . . . . . . . . Schmidt Num er 0? Sample Paints 13 ”can Lineatian Azimuth 103 1 Neon Lineatian Plunge . . . 18.8 let Eigenvalue . 0 885 2nd Eigenvalue . . . 0 081 3rd Ei envalue . . . 0.034 LN 1 E / E2 l . 2 390 LN 1 E2 / E3 1 . . 0 878 (LNlEl/EZl] / 1LN1E2/E3ll 2 720 S herical variance 0 0631 R or . . . . 0.9369 9 Clast a-axis orientation Atllf Pro ection Schmidt Num er oF Sample Points 13 neon Lineatian fizimuth . . . 60.2 Neon Lineatian Plunge 15 4 lst Eigenvalue . 0.896 2nd Eigenvalue 0.091 3rd Ei envalue 0 013 LN 1 E / E2 1 . . 2.292 LN 1 E2 / E3 1 {.914 . 0. 0. 2 Clast a-OXis orientation Figure 45 220 Stereograms of long axis of clasts, subpop 2 ALfl Pro ectian . . . . Schmidt Num er aF Somple Paints . 13 Great Circle Azimuth . . . . 232.0 Great Circle Plunge . 32 3 lst Eigenvalue . 0 532 2nd Eigenvalue . . . . . . 0.406 3rd Ei envalue . 0 062 LN 1 E / E2 l . 0 270 LN 1 E2 / E3). 1.878 1LN1E1/Ele / 1LN1E2/E3ll 0 144 Spherical. variance . . . . 0 4223 . . . . . . . . . 0 S777 . Clast a-axis orientation ALf2 Pra ectian . Schmidt Num er 0F Sample Paints . . 15 Great Circle Azimuth . . . . 231.7 Great Circle Plunge 25 6 lst Eigenvalue . . . . . . 0.598 2nd Eigenvalue . 0 363 3rd Ei envalue . 0 039 LN 1 E /E2 1 . . . . . . 0.499 LN 1 E2 / E3 l . 2 241 1LN1E1/E2l] / 1LN1E2/E3ll 0 223 S herical variance . . . 0 3462 WE . 0 6538 . Clast O‘OXIS orientation AUf Pro ectian . . . . Schmidt Num er aF Sample Points 16 Great Circle Azimuth . . . 181.7 Great Circle Plunge 33 8 lst Eigenvalue 0 529 2nd Eigenvalue . . . . . . . 0.426 3rd Ei envalue 0 044 LN 1 E /E2 l 0 216 LN 1 E2 / E3 1 . 2 262 1LN1E1/E2l] / 1LN1E2/E3ll 0 096 S herical variance . . . . 0.3453 WE . . . . . 0 6547 o Clast a-axis orientation Figure 46 APPENDIX 6 Stereograms for Clast a-b plane 221 222 Stereograms of poles of a-b plane, lower deposits Alarka site ALfl Projection Schmidt Number of Sample Paints . . 30 lst Eigenvalue . . . . . . 0.612 2nd Eigenvalue . 0 219 3rd Ei envalue . . . . . . 0.169 (E / E2 l . 1 026 LN 1 E2 / E3 1 0 262 1LN1E1/E2l] / 1LN1E2/E3ll 3 910 S herical variance 0 2718 R or . 0 7282 o Orientation of a-b plane ALf2 Pro ectian Schmidt Num er of Sample Points 0 lst Eigenvalue . 0 691 2nd Eigenvalue . . . . 0.173 3rd Ei envalue . 0 137 LN 1 E / E2 l . . . . . . . 1.386 LN 1 E2/ E3 l . 0.234 1LN1E1/E21] / 1LN1E2/E3ll 5 919 S herical variance . . . . 0.1989 WE . . . . 0.8011 0 Orientation aF a-b plane Figure 47 223 Stereograms of poles of a-b plane, upper deposit at Alarka site AUf Pra ectian . . . . . . . Schmidt Num6er of Sample Paints 30 lst Eigenvalue . . . . . . . 0 584 2nd Eigenvalue . 0 317 3rd Ei envalue . 0 099 LN 1 E / E2 1 . . . . . . . 0 610 LN 1 E2 / E3 1 . 1 166 1LN1E1/E2ll / 1LN1E2/E3ll 0 523 Spherical variance . . . . . 0.3072 Rbor . . 0 6928 a Orientation of o-b plane AUa Pro ectian Schmidt Num er 0F Sample Points 28 lst Eigenvalue . 0 622 2nd Eigenvalue . 0 219 3rd Ei envalue . 0 160 LN 1 E / E2 1 . . . . . . . 1 044 LN 1 E2 / E3 1 0 316 1LN1E1/E2l] / 1LN1E2/E3ll 3 300 S herical variance . . . . 0.2440 R or . 0 7560 a Orientation of a-b plane Figure 48 224 Stereograms of poles of a—b plane, Bradley site BLa Pro ectian Schmidt Num6er of Sample Paints . . 29 let Eigenvalue . 0 660 2nd Eigenvalue . . . . 0.217 3rd Ei envalue . . . . . . 0.124 LN 1 E / E2 l . 1 113 LN 1 E2 / E3 l . . . . . . 0.561 1LN1E1/E2)] / 1LN1E2/E3ll 1 983 S herical variance 0 2202 R or . . . . . . . . . 0.7798 0 Orientation 0F a-b plane AUa Pro ectian . . . . . Schmidt Num6er aF Sample Paints 29 lst Eigenvalue . 0 730 2nd Eigenvalue . . . 0.227 3rd Ei envalue 0 043 LN 1 E / E2 1 1 166 LN 1 E2 / E3 l . . . . . . . 1.673 1LN1E1/E2l] / 1LN1E2/E3ll 0 697 S herical variance 0 1776 Rgar . . . . . . . .. . . . 0.8224 0 Orientation 0F a-b plane Figure 49 APPENDIX 7 Clay mineralogy data 225 226 Table 19: data Pedon Horizon 14 minerals mica kaolinite relative abundance* ddNOOOJN NN-‘NNN 101003me NNNNNN 2 2 3 3 2 2 1 2 2 3 2 3 2 3 NM-‘NNQNW Add—L—L—L—Ld NNWNQQWN A—L—L—L-LN AOOOO-‘ 0303030003“) 227 Table 19: data cont' Pedon Horizon 14 min mica kaolinite relative abundance* d—t—L—L-‘d—L—t—t—L C-‘doo wwwwwwwwmw HOOD OCOOOO-e 0003003000303 1 1 1 0 1 O 1 1 1 0 1 0 1 O 1 0 1 1 ODODODCDCDODODCDCO d-l-J-‘d—A—L—b 4044404.; wwwwwwmw 228 Table 19: Cla data cont' Pedon 14ofizon 14 nfin nflca kaofinfie relative abundance* Bw1 2Bw2 38w3 4Bw4 58w5-2 58w54 * See section 6.2.2 for definition APPENDIX 8 Iron species data 229 230 Table 20: Iron species data. all sites Pedon Horizon % Fed‘ % Fet" 100'Fed/Fet ALf1 A 0.78 2.41 32.4 AE 0.66 2.53 26.1 E 0.73 2.71 26.9 2Bt 1.00 3.86 25.9 3BC1 0.90 3.75 24.2 3BC2 0.84 3.58 23.4 ALf2 A1 0.89 2.84 31.4 A2 0.77 2.72 28.4 E 0.80 2.94 27.0 2Bt 1.24 4.07 30.4 2BC1 1.02 4.02 25.3 3BC2 0.80 3.68 21.7 AUf A 0.74 3.50 21.1 E 1.03 4.34 23.8 Bw1 1.35 4.51 29.8 8w2-2 1.09 4.33 25.2 Bw2-1 1.14 3.56 32.2 2BC1 1.18 3.57 33.2 28C2 0.99 3.57 27.6 AUa A 0.96 2.91 33.0 Bw-2 1.20 4.31 27.7 Bw-1 1.12 4.08 27.4 Bt 0.98 5.06 19.3 2BC1 1.07 4.12 25.9 2BC2 0.92 3.44 26.7 2C 0.76 3.31 22.9 3C 0.85 3.85 22.2 HLf A 1.45 3.84 37.7 Bw 1.86 7.77 23.9 m 1.27 4.81 26.5 Cg 0.83 3.76 21.9 2Cg 0.79 3.30 23.9 3C 4.83 7.82 61.8 231 Table 20: iron species data, all sites (cont'd) Pedon Horizon % Fed‘ % Fet' 100‘Fed/Fet HLa A 1.95 5.66 34.4 B/A 2.35 6.50 36.2 Bt1 1.80 8.81 20.4 Bt2 4.06 9.17 44.3 E) 2.35 8.38 28.0 C 1.97 7.28 27.1 2C 2.26 6.06 37.3 309 1.48 ~ 4.79 30.9 40 2.56 6.25 41.0 5Cg 0.85 4.54 18.8 HMf A 1.72 4.19 41.1 AB 1.86 4.54 40.9 Bt1 3.45 6.85 50.3 Bt2 3.71 7.43 49.9 BIC-2 3.37 6.79 49.7 BIC-1 3.03 6.58 46.0 K) 3.70 7.83 47.2 HMa A 2.67 5.41 49.4 AB 2.12 4.86 43.6 Ab 2.35 6.28 37.3 E1b 1.79 ' 4.06 44.0 E2b 1.98 4.80 41.2 Btb-2 2.70 6.10 44.3 Btb-1 3.13 5.17 60.6 C/Bb-2 1.94 5.70 34.0 C/Bb-1 2.42 6.54 36.9 Crb 1.55 4.26 36.5 HUf A 1.95 4.84 40.4 BA 2.29 5.80 39.5 Bw1 1.80 6.31 28.5 2Bw2-2 2.19 6.35 34.5 2Bw2-1 2.80 6.55 42.8 3B13-2 3.82 8.16 46.9 3Bt3-1 3.59 . 6.83 52.5 4Bt4 2.91 7.31 39.8 232 Table 20: Iron species data, all sites (cont'd) Pedon Horizon % Fed" °/o Fet* 100*Fed/Fet HUa A 2.01 5.72 35.1 BA 2.20 6.03 36.5 Bw1 2.29 5.48 41.8 Bw2 1.98 4.68 42.3 28w3 1.99 6.73 29.6 3Bw4 2.45 4.99 49.1 3Bt 2.69 6.64 40.5 BLf A 1.26 3.84 32.8 Bw 1.46 4.43 33.1 091 1.20 4.34 27.6 2092 1.04 4.27 24.4 BLa Ap 1.11 3.74 29.7 E 1.35 4.19 32.2 Bw1 1.68 4.80 34.9 2Bw2 1.52 4.54 33.5 38w3 1.49 _ 4.47 33.4 4Bw4 1.14 4.02 28.2 BUa Ap 1.19 3.72 31.9 Bw1 1.26 4.05 31.2 2Bw2 1.10 3.77 29.2 38w3 1.04 3.43 30.3 4Bw4 1.07 3.77 28.4 5Bw5-2 1.29 4.17 30.9 5Bw5-1 1.41 5.36 26.3 * Weight °/o of fine fraction PART VI: LIST OF REFERENCES 233 LIST OF REFERENCES Acker, L. L., and R. D. Hatcher. 1970. Relationships between structure and topography in northwest South Carolina. South Car. Div. Geol. Geol. Notes 14:35-48. Alexander, B. B. 1974. Extractable iron in relation to soil age on terraces along the Truckee River, Nevada. Soil Sci. Soc. Am. J. 38:121-124. Alexander, B. B., and N. Holowaychuk. 1983. Soils on terraces along the Cauca river, Colombia: 1. Chronosequence characteristics. Soil Sci. Soc. Am. J. 47:715-721. Alexander, E. B., and N. Holowaychuk. 1983. Soils on terraces along the Cauca river, Colombia: H. The sand and clay fractions. Soil Sci. Soc. Am. J. 47:721-727. Anderson, J. U. 1963. An improved pretreatment for mineralogical analysis of samples containing organic matter. Clays Clay Min. 10:380-388. Arduino, B., E. Barberis, F. Ajmone Marsan, E. Zanini, and M. Franchini. 1986. Iron oxides and clay minerals within profiles as indicators of soil age in northern Italy. Geoderrna 37:45-55. Arduino, E., E. Barberis, F. Carraro, and M. G. Fomo. 1984. Estimating relative ages from iron-oxide/total-iron ratios of soils in the western Po valley, Italy. Geoderrna 33:39- 52. Battiau-Queney, Y. 1989. Constraints from deep crustal structure on long-term landforrn development of the British Isles and Eastern United States. Geomorphology 2:53-70. Birkeland, P. W. 1978. Soil development as an indication of relative age of Quaternary deposits, Baffin Island, N.W.T., Canada. Arct. Alp. Res. 10:733-747. Birkeland, P. W. 1984. Soils and Geomorphology. Oxford University Press, New York. 372 pp. Birkeland, P. W. 1990. Soil-geomorphic research - a selective overview. In P. L. K. Knuepfer, and L. D. McFadden (eds) Soils and Landscape Evolution. Geomorphology 3:207-224. Birkeland, P. W., M. E. Berry, and D. K. Swanson. 1991. Use of soil catena field data for estimating relative ages. Geology 19:281-283. Bockheim, J. G. 1980. Solution and use of chronofunctions in studying soil development. Geoderrna 24:71-85. Bogucki, D. J. 1976. Debris slides in the Mt. Le Conte area, Great Smoky Mountains National Park, USA. Geogr. Ann. 58A: 179-191. Borchardt, G. 1989. Smectites. pp. 657-728 In J. B. Dixon, and S. B. Weed (eds.) Minerals in Soil Environments. Soil Sci. Soc. Am., Madison, WI. 234 Bouchard, M., and M. J. Pavich. 1989. Characteristics and significance of pre- Wisconsinan saprolites in the northern Appalachians. Zeitschrift Geomorph. Suppl. bnd. 72:125-137. Brabb, E. E. 1984. Minimum landslide damage in the United States, 1973-1983. USGS. Open file report. 84-486. 4 pp. Brabb, E. E. 1989. Landslides: Extent and economic significance in the United States. pp. 25-50. In E. E. Brabb, and B. L. Harrod (eds.) Landslides: Extent and Economic Significance. Balkema, Rotterdam. Braun, E. L. 1972. Deciduous Forests of Eastern North America. Hafner, New York. 427 pp. Brimhall, G. H., O. A. Chadwick, C. J. Lewis, W. Compston, I. S. Williams, K. J. Danti, W. E. Dietrich, M. E. Power, D. Hendricks, and J. Bratt. 1991. Deformational mass transport and invasive processes in soil evolution. Science 255:695-702. Brimhall, G. H., and W. E. Dietrich. 1987. Constitutive mass balance relations between chemical composition, volume, density, porosity and strain in metasomatic hydrochemical systems: Results on weathering and pedogenesis. Geochim. Cosmochim. Acta 51:567- 587. Bryan, D. S. 1994. Factors controlling the occurrence and distribution of hematite and goethite in soils and saprolites derived from schists and gneisses in western North Carolina. M.S. thesis. East Lansing, Michigan State U. 125 pp. Bull, W. B. 1962. Relations of alluvial-fan size and slope to drainage-basin size and lithology in western Fresno County, California. USGS Prof. Paper 450—B. 53 pp. Bull, W. B. 1963. Aluvial fan deposits in Western Fresno County, California. J. Geol. 71:243-251. Bull, W. B. 1972. Recognition of alluvial-fan deposits in the stratigraphic record. pp. 63-83. In J. K. Rigby, and W. K. Hamblin (eds) Recognition of Ancient Sedimentary Environments. Soc. Econ. Paleont. Mineral. Spec. Pub]. 16 Bull, W. B. 1990. Stream-terrace genesis: implications for soil development. pp. 351 - 367. In P. L. K. Knuepfer, and L. D. McFadden (eds.) Soils and Landscape Evolution. Geomorphology 3. Buntley, G. J ., and F. C. Westin. 1965. A comparative study of developmental color in a Chestnut-Chernezem-Brunizem soil climosequence. Soil Sci. Soc. Am. Proc. 29:579- 582. Buol, S. W., F. D. Hole, and R. J. McCracken. 1980. Soil Genesis and Classification. Iowa State U. Press, Ames. 406 pp. Burke, R. M., and P. W. Birkeland. 1979. Reevaluation of multiparameter relative age dating techniques and their applications to the glacial sequence along the eastern escarpment of the Sierra Nevada, California. Quat. Res. 11:21-51. Caine, N. 1980. The rainfall intensity-duration control of shallow landslides and debris flows. Geogr. Ann. 62Az23-27. 235 Campbell, R. H. 1975. Soil slips, debris flows and rainstorms in the Santa Monica Mountains and vicinity, southern California. USGS Prof. Pap. 851. 51 pp. Campione, M. A., D. E. McVey, T. L. Gerke, W. D. Briggs, Y. Zhang, J. B. Maynard, and W. D. Huff. 1992. Non-systematic weathering profile in the Blue Ridge Mountains, N.C.: Role of geochemical variables. Geol. Sochm. Abstr. Progs. 24:214. Cannon,S. H., and S. Ellen. 1985. Rainfall conditions for abundant debris avalanches in the San Francisco Bay region, California. Calif. Geol. 38:267-272. Chadwick, O. A., G. H. Brimhall, and D. M. Hendricks. 1990. From a black box to a grey box -- a mass balance interpretation of pedogenesis. In P. L. K. Knuepfer, and L. D. McFadden (eds) Soils and Landscapo Evolution. Geomorphology 3:369-390. Chassie, R. G., and R. D. Goughnour. 1976. National highway landslide experience. Highway Focus 8:1-9. Chinn, T. J. H. 1981. Use of rock weathering rind thickness for Holocene absolute age dating in New Zealand. Arct. Alp. Res. 13:33-45. Ciolkosz, E. J ., B. J. Carter, M. T. Hoover, R. C. Cronce, W. J. Waltman, and R. R. Dobos. 1990. Genesis of soils and landscapes in the Ridge and Valley province of central Pennsylvania. In P. L. K. Knuepfer, andL. D. McFadden (eds) Soils and Landscape Evolution. Geomorphology 3: 245- 261. Clark, G. M. 1979. Problems in correlation of the Appalachian Quaternary record: Examples and future research possibilities south of the glacial border. p. 5. In L. A. Raymond (ed.) Proceedings of the Second Annual Conference on the Quaternary Histog of the Southeastern United Stem. Appalachian State University, Dept. Geology. Clark, G. M. 1987. Debris slide and debris flow historical events in the Appalachians south of the glacial border. pp. 125-138. In J. E. Costa, and G. F. Wieczorek (eds) Debris Flows/avalanches: Process, Recognition, and Mitigation. Reviews in Engineering Geology VII. Geol. Soc. Am., Boulder. Clark, G. M. 1989. Central and southern Appalachian water and wind gap origins: Review and new data. Geomorphology 2:209-232. CleavesE. T. 1989. Appalachian Piedmont landscapes from the Permian to the Holocene. pp. 159-179. In T. W. Gardner, and W. D. Sevon (eds) Appalachian Geomogphology. Elsevier, Amsterdam. Cleaves, E. T., A. E. Godfrey, and O. P. Bricker. 1970. Geochemical balance of a small watershed and its geomorphic implications. Geol. Soc. Am. Bull. 81:3015-3032. Colman, S. M., and K. L. Pierce. 1981. Weathering rinds on andesitic and basaltic stones as a Quaternary age indicator, western United States. USGS Prof. Paper 1210. 56 pp. Costa, J. E. 1984. Physical geomorphology of debris flows pp. 268-317. In J. E. Costa, and P. J. Fleisher (eds) Developments and Applioations of Geomogphology. Springer-Verlag, Berlin. 236 Costa, J. E., and E. T. Cleaves. 1984. The Piedmont landscape of Maryland: A new look at an old problem. Earth Surf. Proc. 9:59-74. Cruden, D. M., S. Thomson, B. D. Bomhold, J. Y. Chagnon, J. Locat, S. G. Evans, J. A. Heginbottom, K. Moran, D. J. W. Piper, R. Powell, D. Prior, and R. M. Quigley. 1989. Landslides: Extent and economic significance in Canada. pp. 1-23. In E. E. Brabb, and B. L. Harrod (eds) Landslides: Extent and Economic Significance. Balkema, Roterdam. Culley, J. L. B. 1993. Density and compressibility. pp. 529 - 539. In M. R. Carter (ed.) Soil Sampling and Methods of Analysis. Lewis Publ., Boca Raton, Fl. Daniels, R. B. 1978. Age of soil landscapes in the Coastal Plain of North Carolina. Soil Sci. Soc. Am. J. 41298-105. Daniels, W. L. 1985. Virgin hardwood forest soils of western North Carolina. Ph.D. thesis. Blacksburg, Virginia Polytechnical Inst. and State U. 179 pp. Daniels, W. L, L. W. Zelany, andC. J. Everett. 1987. Virgin hardwood forest SOllS of the Southern Appalachian Mountains 1. Soil morphology and geomorphology. Soil Sci. Soc. Am. J. 51: 722- 729. Davis, J. C. 1986. Statistics and Data Analysis in Geology. John Wiley and Sons, New York. 646 pp. Day, F. P., D. L. Phillips, and C. D. Monk. 1988. Forest communities and patterns. pp. 141-149. In W. T. Swank, and D. A. Crossley (eds) Forest H r 10 and Ecolo Coweeta. Ecological Studies: Analysis and Synthesis Springer-Verlag, New York. Day, P. R. 1965. Particle fractionation and particle-size analysis. pp. 545-567. In C. A. Black (ed.) Methods of Soil Analysis. Am. Soc. Agron. Monogr. 9, Madison, WI. DeGraff, J. V. 1994. The geomorphology of some debris flows in the Southern Sierra Nevada, California. Geomorphology 10:231-252. Delcourt, P. A. 1980. Quaternary alluvial terraces of the Little Tennessee River valley, east Tennessee. pp. 110- 121, 175- 212. In J. Chapman (ed) The 1272 Archaoological and eolo ical Investi ations in th Tellica Res rv ir. Report of Investigations 29. Dept. of Anthropology, University of Tennessee. Delcourt, P. A., and H. R. Delcourt. 1984. Late Quaternary paleoclimates and biotic responses in eastern North America and the western North Atlantic Ocean. Paleogeo. Paleoclim. Paleoeco. 48:263-284. Dethier, D. P. 1988. The soil chronosequence along the Cowlitz River, Washington. USGS Bull. 1590-F. 47 pp. Dietrich, W. B., and R. Dorn. 1984. Significance of thick deposits of colluvium on hillslopes; A case study involving the use of pollen analysis in the coastal mountains of northern California. J. Geol. 92: 147-158. Douglas, J. E., and W. T. Swank. 1975. Effects of management practices on water quality and quantity, Coweeta Hydrologic Laboratory, North Carolina. USDA Forest Service. Central Technical Report. N-13. 174 pp. 237 Dowdeswell, J. A. 1983. Multivariate statistical approaches to the analysis of late Quaternary relative age data. Progr. Phys. Geo. 7:157-176. Dowdeswell, J. A. 1984. Late Quaternary chronology for the Watts Bay area, Frobisher Bay, southern Baffin Island, N. W. T., Canada. Arct. Alp. Res. 16:311-320. Dowdeswell, J. A., and M. J. Sharp. 1986. Characterization of pebble fabrics in modern terrestrial glaciogenic sediments. Sedimentology 33:699-710. Ellen, S. D., and R. W. Fleming. 1987. Mobilization of debris flows from soil slips, San Francisco Bay region, California. pp. 31-40. In J. E. Costa, and G. F. Wieczoreck (eds) Debris Flows/avalanches: Process, Mitigation, Recognition, and Mitigation. Reviews in engineering geology VII. Geol. Soc. Am., Boulder. Eschner, A. R., and J. H. Patric. 1982. Debris avalanches in eastern upland forests. J. Forestry 345-347. Eswaran, H., and C. Sys. 1970. An evaluation of free iron in tropical basaltic soils Pedologie 20:62-85. Evans, L. J ., and W. A. Adams. 1975. Quantitative pedological studies on soils derived from Silurian mudstones. IV. Uniformity of the parent material and evaluation of internal standards. J. Soil Sci. 26:319-335. Evans, L. J., and W. A. Adams. 1975. Quantitative pedological studies on soils derived from Silurian mudstones. V. Redistribution and loss of mobilized constituents. J. Soil Sci. 26:327-335. Evans, S. G. 1982. Landslides and surficial deposits in urban areas of British Columbia: A review. Can. Geotech. J. 19:269-28. Eyles, N., C. H. Eyles, and A. D. Miall. 1983. Lithofacies types and vertical profile models; an alternative approach to the description and environmental interpretation of glacial diamict and diamictite sequences. Sedimentology 30:393-410. Fanning, D. S., and M. C. B. Farming. 1989. Soil Mosphology, Genesis and Classification. John Wiley and Sons, New York. 395 pp. Feldman, S. B., L. W. Zelazny, and J. C. Baker. 1991a. High-elevation forest soils of the Southem Appalachians: 1. Distribution of parent materials and soil-landscape relationships. Soil Sci. Soc. Am. J. 55:1629-1637. Feldman, S. B., L. W. Zelazny, and J. C. Baker. 1991b. High-elevation forest soils of the Southern Appalachians: H. Geomorphology, pedogenesis and clay mineralogy. Soil Sci. Soc. Am. J. 55:1782-1791. Gamble, E. E. 1991 Geomo hic tu in he rGasc nade River Basin Laclede and Texas Counties, Missouri. Natl. Soil Survey Center, Soil Survey Invest. Rep. No 43. 120 pp. Gerrard, A. J. 1990. Soil variations on hillslopes in humid temperate climates. In P.L.K. Knuepfer, and L. D. McFadden (eds) Soils and Landscape Evolution. Geomorphology 3:225 - 244. 238 Gloppen, T. G., and R. J. Steel. 1981. The deposits, internal structure, and geometry in six alluvial fan-fan delta bodies (Devonian-Norway). pp. 49-69. In F. G. Ethridge, and R. M. Flores (eds) Recent and Anoient Nonmarine Depositionol Environments. Soc. Econ. Paleont. Miner. Spec. Publ. 31. Graham, R. C., and S. W. Buol. 1990. Soil-geomorphic relations on the Blue Ridge Front: II. Soil characteristics and pedogenesis. Soil Sci. Soc. Am. J. 54:1367-1377. Graham, R. C., R. B. Daniels, and S. W. Buol. 1990. Soil-geomorphic relations on the Blue Ridge Front: 1. Regolith types and slope processes. Soil Sci. Soc. Am. J. 54: 1362- 1367. Graham, R. C., S. B. Weed, L. H. Bowen, D. D. Amarasiriwardena, and S. W. Buol. 1989a. Weathering of iron-bearing minerals in soils and saprolite on the North Carolina Blue Ridge Front: H. Clay mineralogy. Clays Clay Min. 37:29-40. Graham, R. C., S. B. Weed, L. H. Bowen, and S. W. Buol. 1989b. Weathering of iron- bearing minerals in soils and saprolite of the North Carolina Blue Ridge front: I. Sand-size primary minerals. Clays Clay Min. 37: 19-28. Grant, W. H. 1987. Debris avalanches and the origin of first-order streams. pp. 103 - 110. In W. T. Swank, and D. A. Crossley (eds) Forest Hydrology and Ecology at Coweeta. Ecological Studies 66. Springer-Verlag, New York. Grantham, J. H., and M. A. Velbel. 1988. The influence of climate and topography on rock-fragment abundance in modern fluvial sands of the southern Blue Ridge Mountains, North Carolina. J. Sed. Petr. 58:219-227. Gray, R. E., H. E. Ferguson, and J. V. Hamel. 1979. Slope stability in the Appalachian Plateaus of Pennsylvania and West Virginia: Rock slides and avalanches. pp. 447-471. In B. Voight (ed.) Developments in Geotechnical Engineering. 143. Elsevier, New York. Greswell, S., D. Heller, and D. N. Swanston. 1979. Mass movement response to forest management in the Central Oregon Coast range. USDA For. Serv. Gen. Tech. Rep. PNW-84. 35 pp. Griffin, G. M., and R. L. Ingram. 1955. Clay minerals of the Neuse river estuary. J. Sed. Petr. 25: 194-200. Gryta, J. J ., and M. J. Bartholomew. 1977. Evidence for late Cenozoic debris-avalanche type deposits in Watauga County, North carolina. Geol. Soc. Am. Abstr. Progs. 9: 142- 143. Guccione, M. J. W. 1982. Stratigraphy, soil development and mineral weathering of Quaternary deposits, midcontinent, USA. PhD. thesis. Boulder, U.Colorado. 302 pp. Hack, J. T. 1982. Physiographic divisions and differential uplift in the Piedmont and Blue Ridge. USGS Prof. Paper 1265. 49 pp. Hack, J. T., and J. C. Goodlett. 1960. Geomorphology and forest ecology of a mountain region in the central Appalachians. USGS Prof. Pap. 347. 83 pp. Hadley, J. B., and R. Goldsmith. 1963. Geology of the eastern Great Smoky Mountains, North Carolina and Tennessee. USGS Prof. Paper 349-B. 118 pp. 239 Hadley, J. B., and A. E. Nelson. 1971. Geologic map of the Knoxville quadrangle, North Carolina, Tennessee, and South Carolina. USGS Misc. Geol. Invest., Map I-654. Hallet, B., and J. Putkonen. 1994. Surface dating of dynamic landforms: Young boulders on aging moraines. Science 265:937-940. Hamilton, W. 1961. Geology of the Richardson Cove and Jones Cove quadrangles, Tennessee. USGS Prof. Paper 349-A. 55 pp. Harbor, J. M. 1986. A comment on certain multivariate techniques used in the analysis of late Quaternary relative age data. Phys. Geog. 72215-225. Harden, J. W. 1982. A quantitative index of soil development from field descriptions: Examples from a chronosequence in central California. Geoderrna 28: 1-28. Harden, J. W. 1990. Soil development on stable landforms and implications for landscape studies. Geomorphology 3:391-398. Harrison, J. B. J., L. D. McFadden, and R. J. Weldon III. 1990. Spatial soil variability in the Cajon Pass czhronosequence implications for the use of soils as a geochronological tool. Geomorphology 3: 399-416. Harvey, A. M. 1990. Factors influencing Quaternary alluvial fan development in southeast Spain. pp. 247-270. In A. H. Rachocki, and M. Church (ed.) Alluvial Fans: A Field Approach. John Wiley and Sons, Chichester. Hatcher, R. D. 1979. The Coweeta Group and Coweeta syncline. Southeastern Geol. 21: 17-29. Hatcher, R. D. 1980. Geologic map and mineral resources summary of the Prentiss quadrangle, North Carolina. North Carolina Dept. Nat. Res. Comm. Development. Hatcher, R. D. 1988. Bedrock geology and regional geologic setting of Coweeta Hydrologic Laboratory in the eastern Blue Ridge. pp. 83-92. In W. T. Swank, and D. A. Crossley (eds) Forest Hydrology and Ecology at Coweeta. Ecological Studies: Analysis and Synthesis 66. Springer-Verlag, New York. Hooke, R. L. 1987. Mass movements in semi-arid environments and the morphology of alluvial fans pp. 505 -529. In M. G. Anderson,-and K. S. Richards (ed.) Slope Stability. John Wiley and Sons, Chichester. Hungr, 0., G. C. Morgan, D. F. VanDine, and D. R. Lister. 1987. Debris flow defences in British Columbia. pp. 201-.222 In J. E. Costa, andG. F. Wieczorek (eds) Debris Flow/avalanches: Process, Recognition, and Mitigation. Reviews in Engineering Geology VII. Geol. Soc. Am, Boulder. Hursh C. R. 1941. The geomorphic aspects of mudflows as a type of accelerated erosion in the southern Appalachians. Trans. Am. Geophys. Union Part 2:253-254. Hurst, V. J. 1977. Visual estimation of iron in saprolite. Geol. Soc. Am. Bull. 88:174- 176. Innes, J. L. 1983. Debris flows Progr. Phys. Geo. 7:469-501. 240 Jackson, M. L., C. H. Lim, and L. W. Zelazny. 1986. Oxides, hydroxides and aluminosilicates. pp. 101-150. In A. Klute (ed.) Methods of Soil Analysis, Part 1: Physical and Mineralogical Methods. Am. Soc. Agr., Madison, WI. Jacobson, R. B., A. J. Miller, and J. A. Smith. 1989. The role of catastrophic events in central Appalachian landscape evolution. Geomorphology 2:257-284. Jahns,R.H. 1978. Landslides. pp. 58-65. In Geophysics Study Committee Geophysical Predictions. Natl. Acad. Sci., Washington, DC. Jenny, H. 1941. Factors of Soil Formation. Mc Graw-Hill, New York. 186 pp. Jian, L., and Y. Jianmo. 1983. The main features of the mudflow in Jiang-Jia ravine. Zeitschrift Geomorph. 27:325-341. Jibson, R. W., and D. K. Keefer. 1993. Analysis of the seismic origin of landslides: Examples from the New Madrid seismic zones. Geol. Soc. Am. Bull. 105:521-536. Johns, W. D., R. E. Grim, and W. F. Bradley. 1954. Quantitative estimates of clay minerals by diffraction methods. J. Sed. Petr. 24:242-251. Kadomura, H., H. Yamamoto, T. Imagawa, H. Takahashi, and R. Suzuki. 1980. Erosion and mudflows at Usu volcano after the 1977 - 78 eruption. Environm. Sci. 3: 155-184. Keefer, D. K., and A. M. Johnson. 1983. Earth flows: Morphology, mobilization and movement. USGS Prof. Paper 1264. 56 pp. Kellerhals, R., and M. Church. 1990. Hazard management on fans, with examples from British Columbia. pp. 335-354. In A. H. Rachocki, and M. Church (eds) Alluvial Fans: A Field Approach. John Wiley and Sons, Chichester. Kesel, R. H., and B. E. Spicer. 1985. Geomorphologic relationships and ages of soils on alluvial fans in the Rio General Valley, Costa Rica. Catena 12: 149 -166. King, P. B. 1964. Geology of the central Great Smoky Mountains, Tennessee. USGS Prof. Paper 349-C. 149 pp. Knuepfer, P. L. K. 1988. Estimating ages of late Quaternary stream terraces from analysis of weathering rinds and soils. Geol. Soc. Am. Bull. 100: 1224-1236. Koch, C. A. 1974. Debris slides and related flood effects in the 4-5 August, 1938, Webb Mountain cloudburst: Some past and present environmental geomorphic implications. M.S. thesis. Knoxville, U. Tenessee. 112 pp. Kochel, R. C. 1987. Holocene debris flows in central Virginia. pp. 139-156. In J. E. Costa, and G. F. Wieczorek (eds) Debris Flows/Avalanches: Process, Recognition, and Mitigation. Reviews in Engineering Geology VII. Geol. Soc. Am., Boulder. Kochel, R. C. 1990. Humid fans of the Appalachian mountains. pp. 109-129. In A. H. Rachocki, and M. Church (eds) Alluvial Fans: A Field Approach. John Wiley and Sons, Chichester. ' 241 Kochel, R. C., and R. A. Johnson. 1984. Geomorphology and sedimentology of humid- temperate alluvial fans, central Virginia. pp. 109-122. In E. H. Koster, and R. J. Steel (eds) Sedimentology of Gravels and Conglomerates Memoir 10 Can. Soc. Petr. Geol., Calgery. Kohlbeck, F., J. Mojica, and A. E. Scheidegger. 1994. Clast orientation of the 1985 lahars of the Nevada del Ruiz, Colombia and implications for depositional processes. Sed. Geol. 88: 175-183. Krohn, J. P., and J. E. Slosson. 1976. Landslide potential in the United States. Calif. Geol. 29:224-231. Krumbein, W. C. 1941. Measurement and geological significance of shape and roundness of sedimentary particles. J. Sed. Petr. 11:64-72. Kunze, G. W., and J. B. Dixon. 1986. Pretreatment for mineralogical analysis. pp. 92 -95. In A. K1ute(ed.) Meth s f oil Anal isP I:Ph ic an Mi eral ' Methods. Am. Soc. Agron., Madison, WI. La Roca Cervigon, N., and V. Calvo-Cases. 1988. Slope evolution by mass movements and surface wash (Valls D'Alcoi, Alicante, Spain). Catena Supplement 12295-102. Lawson, D. E. 1979. A comparison of the pebble orientations in ice and deposits of the Matanuska glacier, Alaska. J. Geol. 87:629-645. Lecce, S. A. 1990. The alluvial fan problem. pp. 3-24. In A. H. Rachocki, and M. Church (eds) Alluvial fans: A Field Approach. John Wiley and Sons, Chichester. Leigh, D.S. 1995. Stratigraphy and soil chronosequence of the Brasstown sites: A model for age assessment of alluvium in the Southern Blue Ridge, USA. Catena in press Levine, E. L., and E. J. Ciolkosz. 1983. Soil development in till of various ages in northeastern Pennsylvania. Quat. Res. 19:85-99. Lim, C. H., and M. L. Jackson. 1982. Dissolution for total elemental analysis. pp. 1-12. In A. L. Page (ed.) Methods of Soil Analysis. Part 2: Chemical and Microbial Properties. Am. Soc. Agr. and Soil Sci. Soc. Am., Madison, WI. Lindsay, J. F. 1968. The development of clast fabric in mudflows. J. Sed. Petr. 38: 1242-1253. Little, L. P., and W. T. Ward. 1981. Chemical and mineralogical trends in a chronosequence developed on alluvium in eastern Victoria, Australia. Geoderrna 25:173- 188. Losche, C. K., R. J. McCracken, and C. B. Davey. 1970. Soils of steeply sloping landscapes in the southern Appalachian Mountains Soil Sci. Soc. Am. Proc. 34:473-478. Mahaney, W. C. 1978. Late-Quatemary stratigraphy and soils in the Wind River Mountains, western Wyoming. pp. 223-264. In W. C. Mahaney (ed.) Quatema_ry Soils. Geo Abstracts, Norwich. Mark, D. M. 1973. Analysis of axial orientation data, including till fabrics. Geol. Soc. Am. Bull. 84:1369-1374. 242 Markewich, H. W., and M. J. Pavich. 1991. Soil chronosequence studies in temperate to subtropical, low-latitude, low-relief terrain with data from the eastern United States. Geoderrna 51 :213-239. Markewitch, H. W., W. C. Lynn, M. J. Pavich, R. G. Johnson, and M. J.C. 1988. Analyses of four Inceptisols of Holocene age, east-central Alabama. USGS Bull. 1589- C. 29 pp. Markewitch, H. W., M. J. Pavich, M. J. Mausbach, R. L. Hall, R. G. Johnson, and P. P. Hearn. 1987. Age relations between soils and geology in the Coastal Plain of Maryland and Virginia. USGS Bull. 1589-A. 34 pp. Markewitch, H. W., M. J. Pavich, M. J. Mausbach, and R. G. Johnson. 1989. A guide for using soil and weathering profile data in chronosequence studies of the Coastal Plain of the eastern United States. USGS Bull. 1589-D. 39 pp. Markewitch, H. W., M. J. Pavich, M. J. Mausbach, B. N. Stuckey, R. G. Johnson, and V. M. Gonzalez. 1986. Soil development and its relation to the ages of morphostratigraphic units in Horry County, South Carolina. USGS Bull. 1589-B. 61 PP- May, R. W., A. Dreimanis, and W. Stankowski. 1980. Quantitative evolution of clast fabric within the Catfish Creek Till, Bradtville, Ontario. Can. J. Earth Sci. 17:1064-1074. Maynard, J. B. 1992. Chemistry of modern soils as a guide to interpreting Precambrian paleosols. J. Geol. 100:279-289. McCahon, T. J ., and L. C. Munn. 1991. Soils developed in Late Pleistocene tills, Medicine Bow Mountains, Wyoming. Soil Sci. 152:377-388. McFadden, L. D., and R. J. Weldon. 1987. Rates and processes of soil development on Quaternary terraces in Cajon Pass, California. Geol. Soc. Am. Bull. 98:280-293. Mehra, O. P., and M. L. Jackson. 1960. Iron oxide removal from soils and clays by a dithionite-citrate system buffered with sodium bicarbonate. Clays Clay Min. 7:317-327. Meierding, T. C. 1984. Correlation of Rocky Mountain Pleistocene deposits by relative dating methods: A perspective. pp. 146-159. In W. C. Mahaney (ed.) Cogelation of Quaternary Deposits. Geo Books, Norwich, England. Meixner, R. E., and M. J. Singer. 1981. Use of a field morphology rating system to evaluate soil formation and discontinuities. Soil Sci. 131:114-123. Menges, C. M. 1990. Soils and geomorphic evolution of bedrock facets on a tectonically active mountain front, western Sangre de Christa Mountains, New Mexico. In P. L. K. Knuepfer, and L. D. McFadden (eds) Soils and Landscapo Evolution. Geomorphology. 3:301 - 332. Miall, A. D. 1977. A review of the braided river depositional environment. Earth Sci. Rev. 13:1-62. Miall, A. D. 1978. Lithofacies types and vertical profile models in braided river deposits: asummary. pp. 597-604. In A. D. Mia11(ed.) Fluvial Sedimentology. Can. Soc. Petr. Geol. Memoir 5. 243 Miall, A. D. 1985. Architectural-element analysis: a new method of facies analysis applied to fluvial deposits. Earth Sci. Rev. 22:261-308. Michalek, D. D. 1968. Fanlike features and related periglacial phenomena of the southern Blue Ridge. PhD. thesis. Charlotte, U. North Carolina. 198 pp. Miller, C. D. 1979. A statistical method for relative-age dating of moraines in the Sawatch Range, Colorado. Geol. Soc. Am. Bull. 90:1153-1164. Miller, C. D. 1979. A statistical method for relative-age dating of moraines in the Sawatch range, Colorado. Geol. Soc. Am. Bull. 90:1153-1164. Mills, H. H. 1981. Some observations on slope deposits in the vicinity of Grandfather Mountain, North Carolina, USA. Southeastern Geol. 22:209-222. Mills, H. H. 1982. Long-term episodic deposition on mountain foot slopes in the Blue Ridge province of North Carolina: Evidence from relative-age dating. Southeastern Geol. 23: 123-128. Mills, H. H. 1982. Piedmont-cove deposits of the Dellwood Quadrangle, Great Smoky Mountains, USA. Zeitschrift Geomorph. 26:163-178. Mills, H. H. 1983. Clast-fabric strength in hillslope colluvium as a function of slope angle. Geogr. Ann. 65Az255-262. Mills, H. H. 1983. Pediment evolution at Roan Mountain, North Carolina, USA. Geogr. Ann. 65A:111-126. Mills, H. H. 1984. Clast orientation in Mount St. Helens debris-flow deposits, North Fork Toutle river, Washington. J. Sed. Petr. 54:626-634. Mills, H. H. 1988. Surficial geology and geomorphology of the Mountain Lake area, Giles county, Virginia, including sedimentological studies of colluvium and boulder streams. USGS Prof. Paper 1469. 57 pp. Mills, H. H., G. R. Brakenridge, R. B. Jacobson, W. L. Newell, M. J. Pavich, and J. S. Pomeroy. 1987. Appalachian mountains and plateaus. pp. 5-50. In W. L. Graf (ed.) Geomogphic Systems of North America. Centennial Spec. Vol. 2. Geol. Soc. Am., Boulder, Colorado. Mokma, D. L., M. L. Jackson, J. K. Syers, and P. R. Stevens. 1973. Mineralogy of a chronosequence of soils from graywacke and mica-schist alluvium, Westland, New Zealand. New Zealand J. Sci. 16:769-797. Morton, D. M., and R. M. Campbell. 1974. Spring mudflows at Wrightwood, southern California. Quat. J. Engin. Geol. 7:377-384. National Research Council. 1985. Reducing Losses from Landsliding in the Qnited States Natl. Acad. Press, Washington, DC. 41 pp. Neary, D. G., and L. W. Swift. 1987. Rainfall thresholds for triggering a debris avalanche event in the southern Appalachian Mountains. pp. 81-92. In J. E. Costa, and G. F. Wieczorek (eds) ebris Flows/Av lanches: Pr cess Rec ition an Miti ation. Reviews in engineering geology VII. Geol. Soc. Am., Boulder. 244 Neary, D. G., L. W. Swift, D. M. Manning, and R. G. Burns 1986. Debris avalanching in the southern Appalachians: An influence on soil formation. Soil Sci. Soc. Am. J. 50:4565-471. Nelson, F. E. 1985. A preliminary investigation of solifluction lobes. Catena 12:23-33. Nemec, W., and R. J. Steel. 1984. Alluvial and coastal conglomerates: Their significant features and some comments on gravelly mass-flow deposits. pp. 1-31. In E. M. Koster, and R. J. Steel (eds) edimentolo of ravels an on 1 crates Memoir 10 Can. Soc. Petr. Geol., Calgary. - Norfleet, M. L., and B. R. Smith. 1989. Weathering and mineralogical classification of selected soils in the Blue Ridge Mountains of South Carolina. Soil Sci. Soc. Am. J. 53:1771-1778. O'Loughlin, C. L. 1974. The effect of timber removal on the stability of forest soils. J. Hydro]. 13:121-134. Ollier, C. 1984. Weathering. Longman, London. 270 pp. Ono, Y. 1990. Alluvial fans in Japan and South Korea. pp. 91-108. In A. H. Rachocki, and M. Church (eds) Alluvial Fans: A Field Approach. John Wiley and Sons, Chichester. Owens, 1. 1973. Alpine mudflows in the Nigel Pass area, Canadian Rocky Mountains. PhD. thesis. Toronto, U. Toronto. 167 pp. Owens, 1. F. 1974. Interrelationships between climatic and material characteristics in the occurrence of alpine mudflows in the Nigel Pass area, Canadian Rocky Mountains. Akad. Wiss. in Gottingen, II. Math-Phys. 29:313-318. OwensJ. P., M. M. Hess, C. S. Denny, and E. J. Dwomik. 1983. Postdepositional alterartion of surface and near-surface minerals in selected Coastal Plain formations of the Middle Atlantic States. USGS Prof. Paper 1067-F. 26 pp. Pal, D. K., S. B. Deshpande, K. R. Venugopal, and A. R. Kalbande. 1989. Formation of di- and tn'octahedral smectite as evidence for paleoclimatic changes in southern and central peninsular India. Geoderrna 45: 175-184. Pavich, M. J. 1986. Processes and rates of saprolite production and erosion on a foliated granitic rock of the Virginia Piedmont. pp. 552-590. In S. M. Coleman, and D. P. Dethier (eds) Rates of Chemical Weathering of Rocks and Minerals. Academic Press, New York. Pavich, M. J. 1989. Regolith residence time and the concept of surface age of the Piedmont "Peneplain". Geomorphology 2: 181-196. Peatross, J. A. 1986. Morphometric analysis of slope stability controls during the Hurricane Camille storm in Virginia. M.S. thesis. Charlottesville, U. Virginia. 185 pp. Philbrick, S. S. 1962. Old landslides in the upper Ohio Valley. Geol. Soc. Am. Spec. Pap. 68:245-246. 245 Phillips, J. D. 1990. Relative ages of wetland and upland surfaces as indicated by pedogenic development. Phys. Geog. 11:363-378. Pierson, T. C. 1981. Dominant particle support mechanisms in debris flows at Mt. Thomas, New Zealand, and implications for flow mobility. Sedimentology 28:49-60. Pierson, T. C. 1984. A rheologic classification of subaerial sediment-water flows. Geol. Soc. Am. Abstr. Progs. 16:623. Pierson, T. C., and J. E. Costa. 1987. A rheologic classification of subaerial sediment- water flows pp. 1-12. In J. E. Costa, and G. F. Wieczorek (eds) Dobris Flows/Avalanches: Process, Recognition, and Mitigatioa. Reviews in Engineering Geology VH. Geol. Soc. Am., Boulder. Pomeroy, J. S. 1980. Storm-induced debris avalanching and related phenomena in the Johnstown area, Pennsylvania, with references to other studies in the Appalachians. USGS Prof. Paper 1191. 24 pp. Pomeroy, J. S. 1982. Landslides in the Greater Pittsburgh region, Pennsylvania. USGS Prof. Paper 1229. 48 pp. Pomeroy, J. S. 1982. Mass movement in two selected areas of western Washington County, Pennsylvania. USGS Prof. Paper 1170-B. 17 pp. Pomeroy, J. S. 1984. Storm-induced slope movements at East Brady, northwestern Pennsylvania. USGS Bull. 1618. 16 pp. Pomeroy, J. S. 1986. Slope movements in the Warren-Allegheny Reservoir area, northwestern Pennsylvania. USGS Bull. 1650. 15 pp. Pomeroy, J. S., and R. E. Thomas. 1985. Geologic relationships of slope movements in northern Alabama. USGS Bull. 1649. 13 pp. Powers, M. C. 1982. Comparison chart for estimating roundness and sphericity. Data sheet 18. Am. Geol. Inst. Rachocki, A. H. 1981. Alluvial Fans: An Attempt at an Empirical Approach. John Wiley and Sons, Chichester. 161 pp. Radbruch-Hall, H., R. B. Colton, W. E. Davies, .1. Lucchitta, B. A. Skipp, and D. J. Varnes 1982. Landslide overview map of the conterrninous United States. USGS Prof. Paper 1183. 24 pp. Rapp, A. 1963. The debris slides at Ulvédal, western Norway. An example of catastrophic slope processes in Scandinavia. Akad. Wiss. in Gdttingen, II. Math-Phys. 1 3: 1 95-210. Rapp, A., and L. Stromquist. 1976. Slope erosion due to extreme rainfall in the Scandinavian mountains. Geogr. Ann. 58A:]93-200. Ratte, C. A., and D. D. Rhodes. 1977. Hurricane induced landslides on Dorset Mountain, Vermont. Geol. Soc. Am. Abstr. Progs. 9:311. 246 Raymond, L. A. 1976. Glacial, periglacial and pseudo-glacial features in the Grandfather Mountain area, North Carolina. Southeastern Geol. 18:213-229. Rebertus, R. A., and S. W. Buol. 1985. Iron distribution in a developmental sequence of soils from mica gneiss and schist. Soil Sci. Soc. Am. J. 49:713-720. Reed, J. C. 1964. Geology of the Linville Falls quadrangle, North Carolina. U S Geol. Survey. Bull. 1161-B. 53 pp. Reineck, H. B., and I. B. Singh. 1986. Depoitional Sedimentag Environments Springer-Verlag, Berlin. 551 pp. Reneau, S. L., W. E. Dietrich, C. J. Wilson, and J. D. Rogers. 1984. Colluvial deposits and associated landslides in the northern San Francisco Bay area, California, USA. pp. 425-430. In Proceedings, IV International Symposium on Landslides. Can. Geotech. Soc., Toronto. Rieck, R. L., H. A. Winters, D. L. Mokma, and M. M. Moertland. 1979. Differentiation of surficial glacial drift in southeastern Michigan from 7-A/10-A X-ray diffraction ratios of clays. Geol. Soc. Am. Bull. 90:216-220. Rogers, N. W., and M. J. Selby. 1980. Mechanisms of shallow translational landsliding during summer rainstorms: North Island, New Zealand. Geogr. Ann. 62A: 1 1-21. Ruhe, R. V. 1956. Geomorphic surfaces and the nature of soils. Soil Sci. 82:441-455. Rust, B. R., and E. H. Koster. 1984. Coarse alluvial deposits. pp. 53-70. In R. G. Walker (ed.) Facies Models. Geoscience Canada, Reprint Series 1 Geol. Soc. Can., Toronto. Schneider, R. H. 1973. Debris slides and related flood damage resulting from Hurricane Camille, 19-20 August and subsequent storm, 5-6 September, 1969, in the Spring Creek drainage basin, Greenbrier County, West Virginia. Ph.D. thesis. Knoxville, U Tennessee. 131 pp. Schuster, R. L. 1978. Introduction. pp. 1-11. In R. L. Schuster, and R. J. Krizek (eds) Landslides. Analysis and Cootrol. Transp. Res. Board, Special report 176 Natl. Acad. Sci., Washington, DC. Schuster, R. L., and R. W. Fleming. 1986. Economic losses and fatalities due to landslides. Assoc. Eng. Geol. Bull. 23:11-28. Scott, R. C. 1972. The geomorphic significance of debris avalanching in the Appalachian Blue Ridge Mountains. Ph.D. thesis. Athens, U. Georgia. 120 pp. Selby, M. J. 1976. Slope erosion due to extreme rainfall: A case study from New Zealand. Geogr. Ann. 58A:]31-138. Selby, M. J. 1982. Hillslope Materials and Pgocesses Oxford U. Press, Oxford. 264 PP- Shafer, D. S. 1984. Late-Quatemary paleoecolog'ic, geomorphic, and paleoclimatic history of Flat Laurel Gap, Blue Ridge Mountains, North Carolina. M.S. thesis. Knoxville, U. Tennessee. 148 pp. 247 Shafer, D. S. 1988. Late Quaternary landscape evolution at Flat Laurel gap, Blue Ridge Mountains, North Carolina. Quat. Res. 3027-11. Sharp, R. P. 1969. Semiquantitative differentiation of glacial moraines near Convict Lake, Sierra Nevada, California. J. Geol. 77:69-91. Sharpe, C. F. S. 1938. Landslides and Rolated Phenomena. Pageant, New Jersey. 238 PP- Shultz, A. W. 1984. Subaerial debris-flow deposition in the Upper Paleozoic Cutler Formation, Western Colorado. J. Sed. Petr. 54:759-772. Sidle, R. C., and D. N. Swanston. 1982. Analysis of a small debris slide in coastal Alaska. Can. Geotech. J. 19:167-174. Smith, G. A. 1986. Coarse-grained nonmarine volcaniclastic sediment: Terminology and depositional processes. Geol. Soc. Am. Bull. 97: 1-10. Soil Survey Division Staff. 1993. Soil Survey Manual. Handbook No. 18. US. Dept. of Agric., Washington, DC. 437 pp. Soil Survey Laboratory Staff. 1992. Soil Survey Laboratory Methods Manual. Soil Survey Investigations Report No. 42. US. Dept. Agric., Washington, DC. 400 pp. Soileau, J. M., and R. J. McCracken. 1967. Free iron and coloration in certain well- drained Coastal Plain soils in relation to their other properties and classification. Soil Sci. Soc. Am. Proc. 31:248—255. Soller, D., and J. P. Owens. 1991. The use of mineralogical techniques as relative age indicators for weathering profiles on the Atlantic Coastal Plain, USA. Geoderma 51:111- 13 1 . Stolt, M. H., J. C. Baker, and T. W. Simpson. 1991. Micromorphology of the soil- saprolite transition zone in Hapludults of Virginia. Soil Sci. Soc. Am. J. 55: 1067-1075. Stolt, M. H., J. C. Baker, and T. W. Simpson. 1992. Characterization and genesis of saprolite derived from gneissic rocks of Virginia. Soil Sci. Soc. Am. J. 56:531-539. Stolt, M. H., J. C. Baker, and T. W. Simpson. 1993. Soil-landscape relationships in Virginia: 1. Soil variability and parent material Uniformity. Soil Sci. Soc. Am. J. 57:414- 421. Swanson, F. J ., M. M. Swanson, and C. Wods. 1981. Analysis of debris avalanche erosion in steep forest lands. pp. 67-94. In T. R. H. Davies, and A. J. Pearce (eds) Erosion and Sediment Transport in Pacific Rim Steeplgds I.A.H.S. 132, Christchurch, New Zealand. Swanston, D. N. 1979. Effect of geology on soil mass movement activity in the Pacific Northwest. pp. 89-116. In C. T. Youngberg (ed.) Eorest Soils and Landuse. U. Colorado, Fort Collins. Swift, L. W., G. B. Cuningham, and J. E. Douglas. 1988. Climatology and hydrology. pp. 35-55. In W. T. Swank, and D. A. Crossley (eds) Forest Hydrology and Ecology at Coweeta. Ecological Studies: Analysis and Synthesis 66. Springer-Verlag, New York. 248 Switzer, P., J. W. Harden, and R. K. Mark. 1988. A statistical method for estimating rates of soil development and ages of geologic deposits: A design for soil-chronosequence studies Math. Geol. 20: 49- 62. Takahashi, T. 1981. Debris flows. Annu. Rev. Fluid Mech. 13:57-77. Talvenheimo, G., and J. L. White. 1952. Quantitative analysis of clay minerals with the X-ray spectrometer. Anal. Chem. 24: 1784-1789. Tanaka M. 1976. Rate of erosion in the Tanzawa mountains, central Japan. Ann. Geol. 58A:155-163. Tanaka, S., K. Sakamoto, J. Takagi, and M. Tsuchimoto. 1968. Search for Aluminium 26 induced by cosmic-ray muons in terrestrial rock. J. Geophys. Res. 73:3303-3309. Thomas, D. J. (in press). Soil Survey of Macon County, North Carolina. USDA-SCS, US Govt. Print. Off., Washington, DC. 289 pp. Torrent, J ., U. Schwertmann, H. Fechter, and F. Alferez. 1983. Quantitative relationships between soil color and hematite content. Soil Sci. 136:354-358. Torrent, J ., U. Schwertmann, and D. G. Schulzefi 1980. Iron oxide mineralogy of some soils of two terrace sequences in Spain. Geoderrna 23:191-208. Ui T., S. Kawachi, and V. E. Neall. 1986. Fragmentation of debris avalanche material during flowage - Evidence from the Pungarehu Formation, Mount Egmont, New Zealand. J. Volc. Geothermal Res. 27:255-264. Van Wambeke, A. R. 1962. Criteria for classifying tropical soils by age. J. Soil Sci. 13:124-132. VanDine, D. F. 1985. Debris flows and debris torrents in the Southern Canadian Cordillera. Can. Geotech. J. 22:44-68. Vames, D. J. 1958. Landslide types and processes. Highway Research Board. Special Report. 29. 20—47 pp. Vames, D. J. 1975. Slope movements in the Western United States. pp. 1-17. In Guelph symposium on Geomorphlogy. Mass Wasting. Geo Abstracts, Norwich. Vames, D. J. 1978. Slope movement types and processes. pp. 12-33. In R. L. Schuster, and R. J. Krizek (eds) L n slid .Anal is an ontr 1. Special Report 176 Natl. Acad. Sci., Washington, DC. Velbel, M. A. 1984. Natural weathering mechanisms of almandine garnet. Geology 12:631-634. Velbel, M. A. 1985. Geochemical mass balances and weathering rates in forested watersheds of the southern Blue Ridge. Am. J. Sci. 285:904-930. Velbel, M. A. 1987. Alluvial fan origin for terrace deposits of the southeast Prentiss quadrangle, near Otto, North Carolina. Southeastern Geol. 87-103. 249 Vidic, N., M. Pavich, and F. Lobnik. 1991. Statistical analyses of soil properties on a Quaternary terrace sequence in the upper Sava River Valley, Slovenia, Yugoslavia. Geoderrna 51:189-212. Visser, J. N. J. 1989. Stone orientation in basal glaciogenic diamictite: Four examples from the Permo-Carboniferous Dwyka Formation, South Africa. J. Sed. Petr. 59:935- 943. Walton, A. W., and B. A. Palmer. 1988. Lahar facies of the Mount Dutton Formation (Oligocene-Miocene) in the Marysvale volcanic field, southwestern Utah. Geol. Soc. Am. Bull. 100:1078-1091. Wang, C., and R. W. Arnold. 1973. Quantifying pedogenesis for soils with discontinuities. Soil Sci. Soc. Am. Proc. 37:271-278. Weaver, W. B., and S. A. Schumm. 1974 Fan-head trenching: An example of a geomorphic threshold. Geol. Soc. Am. Abstr. Pros. 6:481. Wells, N. A. 1984. Sheet debris flow and sheetflood conglomerates in Cretaceous cool maritime alluvial fans, South Orkney Islands, Antarctica. pp. 133-146. In E. H. Koster, and R. J. Steel (eds) Sedimentology of Gravels and Conglomerates. Memoir 10 Can. Soc. Petr. Geol., Calgary. Wells, S. G., and A. M. Harvey. 1987. Sedimentologic and geomorphic variations in storm-generated alluvial fans, Howgill Fells, northwest England. Geol. Soc. Am. Bull. 98: 182-198. Whipple, K. X., and T. Dunne. 1992. The influence of debris-flow rheology on fan morphology, Owens Valley, California. Geol. Soc. Am. Bull. 104:887-900. White, W. A. 1950. Blue Ridge Front - a fault scarp. Geol. Soc. Am. Bull. 61: 1309- 1346. Whitehouse, I. E. 1983. Distribution of large rock avalanche deposits in the central Southern Alps, New Zealand. New Zealand J. Geol. Geophys. 26:271-279. Whitehouse, I. E., M. J. McSaveney, P. L. K. Knuepfer, and T. J. Chinn. 1986. Growth of weathering rinds on Torlesse sandstone, Southern Alps, New Zealand. pp. 419-435. In S. M. Colman, and D. P. Dethier (eds) Rates of Chemical Weathering of ngs and Minerals. Academic Press, Orlando. Whittecar, G. R. 1985. Stratigraphy and soil development in upland alluvium and colluvium, north-central Virginia Piedmont. Southeastern Geol. 26:117-129. Whittecar, R. G., and D. W. Ryter. 1992. Boulder streams, debris fans and Pleistocene climate change in the Blue Ridge Mountains of Central Virginia. J. Geol. 100:487-494. Whittig, L. D., and W. R. Allardice. 1986. X-ray diffraction techniques. pp. 339-342. In A. K1ute(ed.) e h ofs il a] sis. Part 1: Physical and Mineralogical Methods. Am. Soc. Agr. and Soil Sci. Soc. Am., Madison, WI. 250 Wieczorek, G. F. 1987. Effect of rainfall intensity and duration on debris flows in central Santa cruz Mountains, California. pp. 93-104. In J. E. Costa, and G. F. Wieczorek (eds) Debris Flows/Avalanches: Process, Recognition, and Mitigation. Reviews in Engineering Geology VII. Geol. Soc. Am., Boulder. Williams, G. P., and H. P. Guy. 1971. Debris avalanches - A geomorphic hazard. pp. 25-46. In D. R. Coates (ed.) Environmental Geomorphology. State Univ. New York, Binghamton, NY. Williams, G. P., and H. P. Guy. 1973. Erosional and depositional aspects of hurricane Camille in Virginia, 1969. USGS Prof. Paper 804. 80 pp. Wolman, M. G., and R. Gerson. 1978. Relative scales of time and effectiveness of climate in watershed geomorphology. Earth Surf. Proc. 3: 189-208. Woodcock, N. H. 1977. Specification of fabric shapes using an eigenvalue method. Geol. Soc. Am. Bull. 88:1231-1236. Woodward, J. C., M. G. Macklin, and J. Lewin. 1994. Pedogenic weathering and relative-age dating of quaternary alluvial sediments in the Pindus Mountains of Northwest Greece. pp. 259-301. In D. A. Robinson, and R. B. G. Williams (eds) Rock Weathering and Landform Evolution. John Wiley and Sons, Chichester. Yaalon, D. H. 1971. Soil forming processes in time and space. pp. 253-257. In D. H. Yaalon (ed.) Paleopedology. Israel U. Press, Jerusalem. Zeiruda, Q., and V. Mencl. 1982. Landslides and their Control. Developments in Geotechnical Engineering 31. Elsevier, Amsterdam. 324 pp. nrcuran sran UNIV. LIBRARIES 1111111111111111111111111111111111111111111111111 31293014216513