fl. .: . .12., t x J ... A 5.62: firhsuxfiréfiv . , V , . . . A V V . z , V . . V . . . . V irratfichhi‘f . :v J... z 5. _ LIBRARY Michigan State University This is to certify that the thesis entitled Homblende Etching as an Indicator of Soil Development and Relative Weathering Among Spodosols presented by Leslie Renee Mikesell has been accepted towards fulfillment of the requirements for the MS. degree in Geological Sciences jeré/aflvZ'A—ZJZ, Zy/é/ Major Professor's Signature /0 £964 {mag/43.4) a; Date MSU is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. I DATE DUE DATE DUE DATE DUE MW 6/01 c:/CIFiC/DateDue.p65-p.15 HORNBLENDE ETCHING AS AN INDICATOR OF SOIL DEVELOPMENT AND RELATIVE WEATHERING AMONG SPODOSOLS By Leslie Renee Mikesell A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Geological Sciences 2002 ABSTRACT HORNBLENDE ETCHING AS AN INDICATOR OF SOIL DEVELOPMENT AND RELATIVE WEATHERING AMONG SPODOSOLS By Leslie Renee Mikesell Homblende etching holds great promise as a relative weathering indicator and as another soil development indicator. This research was designed to test hornblende denticulation amplitude as a weathering indicator, by comparing it to a standard weathering indicator (quartz/ feldspar ratios), and to determine its reliability as an indicator of soil development by comparing it to four soils in various stages of development. The natural weathering of hornblende produces etching, which results in a “denticulated” margin. These features gradually and systematically increase in depth and size. In this study, hornblende denticulation amplitude measurements exhibited a clear trend showing more extensively weathered grains up profile, and therefore should be considered an indicator of weathering. Homblende denticulation amplitude is a record of weathering, which perhaps more accurately reflects weathering intensity rather than duration. The denticulation amplitude measurements also mimicked the soil development order of the four soils showing that it will also serve as a relative weathering indicator among soils of the same age. Due to the fact that it correlated with the rank order of soil development, it can also be used as a soil development indicator. In this study, mineral weathering appears to be related to the degree of soil development in soils representing different degrees of development but of the same age. ACKNOWLEDGMENTS There are many people who have influenced my educational choices, and to those who steered me toward geology, I am eternally grateful. I want to thank my committee, Dr. Michael Velbel, Dr. Randy Schaetzl, and Dr. Gary Weissmann, for guiding me through this process and teaching me more than just geology; I also learned about the educational process and about myself. Thank you for your patience and encouragement, and for sharing your knowledge, expertise, and time with me as I worked toward my goals. A special thanks goes to my advisor, Dr. Velbel, for taking me on as a student after I attended a small, liberal arts college and then took time off for several years before pursuing this degree. I truly appreciate what you did for me. Thank you, Dr. Schaetzl, for your help collecting samples in the field. I know I could not have done that without you. I would like to thank my family for all of their love and support. I am so appreciative of my wonderful parents, Lowell and Betty Moore, who instilled in me a love of education. I am grateful for support from my fabulous sister, Dawn Smithson, who thinks more highly of me than she probably should, and her wonderful husband, Scott, who is a great addition to our family. And last, but certainly not least, I want to thank my awesome husband, Marc, whose unconditional love and unwavering support are what saw me through some rough days during this endeavor. I cannot express in words what you mean to me. I will always be grateful that you are in my life. Special thanks go to many others who also helped me. Many thanks go to Beth Weisenbom, who helped me more than she even knows. You taught me a lot about iii science, dedication, detemiination, and work ethics. I also want to thank Ewa Danielewicz at the Center for Advanced Microscopy who deserves all the credit for the great SEM images in this thesis. Thanks to Bruce Knapp of the USDA for his help in the field finding the sites to sample. I am also so glad to have met Dr. Lina Patino who has been such an encouragement and friend. I want to express my appreciation to the Geology Office and Library staff for the great help they have been. Thank you, all for your support. A special thanks also goes to Dr. Max Reams of Olivet Nazarene University. Dr. Reams is the one who awoke in me the love of geology, nurtured it, and then directed me to Michigan State University to allow it to continue to grow. Thank you. I want to thank the Department of Geological Sciences for financial support through two years of teaching assistantships. Also, this material is based upon work supported by the National Science Foundation under Grant No. 9819148 and Grant No. 6799-00. “Can you fathom the mysteries of God? ...speak to the earth, and it will teach you, ...for does it not know that the hand of the Lord has done this?” Selections from Job 11 and 12. iv TABLE OF CONTENTS LIST OF TABLES ........................................................................................................ viii LIST OF FIGURES ...................................................................................................... ix KEY TO SYMBOLS OR ABBREVIATIONS ............................................................ xvii INTRODUCTION ........................................................................................................ 1 Homblende Etching as a Relative Age Indicator .................................................... 1 Quartz/Feldspar Weathering Ratio ......................................................................... 2 Soil Development .................................................................................................... 3 Research Questions ................................................................................................. 4 Purpose of Study ..................................................................................................... 4 CHAPTER 2 Homblende Weathering ................................................................................................ 6 General Description of Homblende ........................................................................ 6 Homblende Distribution ......................................................................................... 6 Weathering Environment ........................................................................................ 7 Homblende Weathering Rates ................................................................................ 8 Homblende Weathering Features ........................................................................... 9 Homblende Weathering in Soil .............................................................................. 15 CHAPTER 3 Quartz and Feldspar Weathering .................................................................................. 18 Mineral Persistence ................................................................................................. 1 8 Quartz and Feldspar ................................................................................................ 20 Quartz/Feldspar Ratio ............................................................................................. 22 CHAPTER 4 Spodosols ...................................................................................................................... 24 Podzolization ........................................................................................................... 24 Factors Contributing to Podzolization .................................................................... 25 The Spodosol Profile ............................................................................................... 27 Extractable Iron and Aluminum in Spodosols ........................................................ 28 CHAPTER 5 Study Area .................................................................................................................... 31 Site Location and Description ................................................................................. 31 Site Selection Criteria ............................................................................................. 31 CHAPTER 6 Methods ......................................................................................................................... 37 Sample Preparation ................................................................................................. 37 Homblende Denticulation Amplitude Determination ............................................. 38 Homblende Classification Scheme ................................................................... 38 Unexpected Etching Features ........................................................................... 39 Quartz/Feldspar Ratio Determination ..................................................................... 49 Iron and Aluminum Extractions .............................................................................. 50 Laboratory-Produced Etching Features ........................................................... 50 CHAPTER 7 Results ........................................................................................................................... 69 Soil Classification ................................................................................................... 69 Homblende Denticulation Amplitude ..................................................................... 81 Quartz/Feldspar Ratios ............................................................................................ 85 Extractable Iron and Aluminum .............................................................................. 85 CHAPTER 8 Discussion ..................................................................................................................... 91 Homblende Denticulation Amplitude ..................................................................... 91 Horizon vs. Homblende Denticulation Amplitude ................................................. 92 Influence of pH on Dissolution Rates ..................................................................... 94 Factors that Influence pH ........................................................................................ 95 Kalkaska Heavy Mineral Grains and Clay ....................................................... 96 Quartz/Feldspar Ratios as a Weathering Indicator ................................................. 107 Grain Size and Quartz/Feldspar Ratios ................................................................... 108 Summary ................................................................................................................. 1 10 CHAPTER 9 Conclusions ................................................................................................................... 1 12 Homblende Denticulation Amplitude ..................................................................... 112 Comparison of Homblende Denticulation Amplitude and Soil Development ....... 112 Comparison of Homblende Denticulation Amplitude and Quartz/Feldspar Ratios ...................................................................................... 112 Comparison of Soil Development and Quartz/Feldspar Ratios .............................. 113 Weathering Versus Soil Development .................................................................... 113 Homblende as an Age Indicator ............................................................................. 114 APPENDICIES Appendix A Denticulation Amplitude Measurements Raw Data Table ..................................... 117 Appendix B Quartz and Feldspar Raw Data Table .................................................................... 118 vi Appendix C Extractable Iron and Aluminum ............................................................................. 119 REFERENCES ............................................................................................................. 1 25 vii LIST OF TABLES Table l. ........................................................................................................................... 66 Pedon Data for Laboratory-Produced Alteration Features Table 2. ........................................................................................................................... 70 Physical Properties of Pedons Table 2 (continued) .......................................................................................................... 71 Physical Properties of Pedons Table 3. ........................................................................................................................... 72 Chemical Properties of Pedons Table 4. ........................................................................................................................... 83 Denticulation Amplitude Results Table 5. ............................................................................................................................ 86 Quartz/Feldspar Ratios Table 6. ........................................................................................................................... 89 Fe and Al Extraction Calculations viii LIST OF FIGURES Images in this thesis are presented in color. Figure 1. ........................................................................................................................... 1 Scanning electron micrograph showing naturally weathered hornblende with etch pits, denticulated margin, and “teeth” parallel to the c-axis. Arrow indicates c-axis direction. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 2. ........................................................................................................................... 11 Scanning electron micrograph of crystallographically controlled etch pits, which form parallel to the c-axis in the direction of cleavage. Arrow indicates c-axis direction. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 3. .......................................................................................................................... 12 Petro graphic micrograph of “sawtooth” termination on a naturally weathered hornblende grain. Image captured with a Pixera digital imaging system. Figure 4. .......................................................................................................................... 12 Scanning electron micrograph of weathered hornblende showing ferruginous microboxwork (B) separated from denticulated hornblende (D) by void space with pendants (P) projecting from the microboxwork into the void space (Figure 6 in Velbel, 1989). Figure 5. .......................................................................................................................... l3 Photomicrograph showing hornblende has been completely removed, leaving only ferruginous boxwork (B) and pendants (P), or “negative pseudomorph” (Figure 9 in Velbel, 1989). Figure 6. .......................................................................................................................... 13 Scanning electron micrograph showing ferruginous boxwork (B) with neoformation pendants (P) (Figure 8 in Velbel, 1989). Figure 7. .......................................................................................................................... 14 Scanning electron micrograph of enlarged lens-shaped etch pits in naturally weathered hornblende. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 8. .......................................................................................................................... l4 Scanning electron micrograph of end-to-end coalescence of etch pits forming grooves or striations in naturally weathered hornblende (Figure 3 in Velbel, 1989). Figure 9. .......................................................................................................................... 19 Mineral-Stability Series in Weathering proposed by Goldich (1938, Table 18). ix Figure 10. ........................................................................................................................ 32 A map of counties in lower Michigan with Kalkaska County highlighted. Figure 11. ........................................................................................................................ 32 A map of Kalkaska County townships with research sites indicated. Sites 1, 2, and 3 are in Blue Lake Township, while Site 4 is in Bear Lake Township. Figure 12. ........................................................................................................................ 33 Map of the morainic system (shown in black) of Michigan with a portion of the Outer Port Huron moraine and the study area for this research in north-central Michigan is indicated in red. (This map is public domain, per a personal communication with Dr. Randy Schaetzl.) Figure 13. ........................................................................................................................ 35 Map showing the ice movement over the Great Lakes region during the Wisconsin glacial stage. The arrows indicate ice flow and the dashed lines indicate ice lobe boundaries. The Outer Port Huron moraine was deposited by the Lake Michigan Lobe, which flowed southward from the Superior east area of the Canadian Shield (modified from Dworkin et al., 1985). Figure 14. ........................................................................................................................ 40 Diagrammatic etching scale showing progress of etching on broken (left) and rounded (right) faces. Depth of etching is shown by indentation in grain margin and is highlighted by shadow (Figure 2 in Locke, 1979). Figure 15. ........................................................................................................................ 41 Class O. Petrographic micrograph of a hornblende grain with no apparent etching. Image captured with a Pixera digital imaging system. Figure 16. ........................................................................................................................ 41 Class 1. Petrographic micrograph of a hornblende grain with etching between 0pm — 4pm. Image captured with a Pixera digital imaging system. Figure 17. ........................................................................................................................ 42 Class 2. Petrographic micrograph of a hornblende grain with etching between 4pm — 8pm. Image captured with a Pixera digital imaging system. Figure 18. ........................................................................................................................ 42 Class 3. Petrographic micrograph of a hornblende grain with etching between 8pm — 12 um. Image captured with a Pixera digital imaging system. Figure 19. ........................................................................................................................ 43 Class 4. Petrographic micrograph of a hornblende grain with etching between 12 um — 16pm. Image captured with a Pixera digital imaging system. Figure 20. ........................................................................................................................ 43 Class 5. Petrographic micrograph of a hornblende grain with etching between 16pm — 20pm. Image captured with a Pixera digital imaging system. Figure 21. ........................................................................................................................ 44 Class 6. Petrographic micrograph of a hornblende grain with etching between 20pm — 24pm. Image captured with a Pixera digital imaging system. Figure 22. ........................................................................................................................ 44 Class 7. Petrographic micrograph of a hornblende grain with etching between 24pm -— 28 um. Image captured with a Pixera digital imaging system. Figure 23. ........................................................................................................................ 45 Class 8. Petrographic micrograph of a hornblende grain with etching between 28pm — 32 um. Image captured with a Pixera digital imaging system. Figure 24. ........................................................................................................................ 46 Petrographic micrographs of weathered hornblende grains with holes. Arrows point to holes. Images captured with a Pixera digital imaging system. Figure 25. ........................................................................................................................ 47 Petrographic micrographs of weathered hornblende grains with fractures. Images captured with a Pixera digital imaging system. Figure 26. ........................................................................................................................ 48 Petrographic micrographs of weathered hornblende grains with both holes and fractures. Images captured with a Pixera digital imaging system. Figure 27. ........................................................................................................................ 53 Scanning electron micrograph of an untreated hornblende grain. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 28. ........................................................................................................................ 53 Scanning electron micrograph of the indicated area on Figure 27 showing smooth surfaces with no etching features. Figure 29. ........................................................................................................................ 54 Scanning electron micrograph of a hornblende grain treated with sodium hypochlorite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 30. ........................................................................................................................ 54 Scanning electron micrograph of the indicated area on Figure 29 showing smooth surfaces with no etching features. xi Figure 31. ........................................................................................................................ 55 Scanning electron micrograph of a hornblende grain treated with sodium citrate- dithionite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 32. ........................................................................................................................ 55 Scanning electron micrograph of the indicated ‘Area 1’ on Figure 31 showing laboratory-produced alteration features. Figure 33. ........................................................................................................................ 56 Scanning electron micrograph of the indicated ‘Area 2’ on Figure 31 showing large laboratory-produced alteration feature. Figure 34. ........................................................................................................................ 57 Scanning electron micrograph of a hornblende grain treated with sodium citrate- dithionite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 35. ........................................................................................................................ 57 Scanning electron micrograph of the indicated area on Figure 34 showing laboratory-produced alteration at the edges. Figure 36. ........................................................................................................................ 58 Scanning electron micrograph of a hornblende grain treated with sodium citrate- dithionite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 37. ........................................................................................................................ 58 Scanning electron micrograph of the indicated area on Figure 36 showing laboratory-produced alteration on surfaces. Figure 38. ........................................................................................................................ 61 Stereo pair 1. Polaroid pictures of scanning electron images were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 39. ........................................................................................................................ 62 Stereo pair 2. Polaroid pictures of scanning electron images were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. xii Figure 40. ........................................................................................................................ 63 Stereo pair 3. Polaroid pictures of scanning electron micrographs were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 41. ........................................................................................................................ 64 Stereo pair 4. Polaroid pictures of scanning electron micrographs were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 42. ........................................................................................................................ 73 The Kalkaska pedon is the most strongly expressed Spodosol in the development sequence of this study. Photo taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. Figure 43. ........................................................................................................................ 74 Selected information from the Soil Profile Description Sheet of the Kalkaska pedon. Figure 44. ........................................................................................................................ 75 The strong Rubicon pedon is the second best expressed Spodosol in the development sequence of this study. Picture taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. Figure 45. ........................................................................................................................ 76 Selected information from the Soil Profile Description Sheet of the strong Rubicon pedon. Figure 46. ........................................................................................................................ 77 The weak Rubicon pedon is the third best expressed Spodosol in the development sequence of this study. Photo taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. Figure 47. ........................................................................................................................ 78 Selected information from the Soil Profile Description Sheet of the weak Rubicon pedon. Figure 48. ........................................................................................................................ 79 The Grayling pedon is the least strongly expressed Spodosol in the development sequence of this study. Photo taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. xiii Figure 49. ........................................................................................................................ 80 Selected information from the Soil Profile Description Sheet of the Grayling pedon. Figure 50. ........................................................................................................................ 82 Petrographic micrograph of the denticulated margins of a naturally weathered hornblende grain. Image captured with a Pixera imaging system. Figure 51. ........................................................................................................................ 84 Denticulation amplitude vs. horizon plot for each of the pedons in this study. These plots show the change in the mean denticulation amplitude measurement with change in horizon. Error bars represent 1 standard deviation. Figure 52. ........................................................................................................................ 87 Horizon vs. Quartz/Feldspar Ratio. Numbers were used as horizon designations instead of traditional horizon nomenclature because each pedon has various horizon orders. This plot shows the change in the quartz/ feldspar ratio with depth. Figure 53. ........................................................................................................................ 98 Scanning electron micrograph of a weathered heavy mineral grain from the A horizon of the Kalkaska pedon with very little weathering product. This product is an aluminosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 54. ........................................................................................................................ 98 Energy-dispersive spectrograph of the weathering product indicated the by circle in Figure 53. The high Si content with A1 and 0 led to the determination that it is an aluminosilicate. The Au is the signature of the coating used to prepare the grain. Figure 55. ........................................................................................................................ 99 Scanning electron micrograph of a weathered heavy mineral grain from the A horizon of the Kalkaska pedon with organic material present. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 56. ........................................................................................................................ 99 Energy-dispersive spectrograph of the organic material indicated by the circle in Figure 55 showing high C content. The Au is the signature of the coating used to prepare the grain. Figure 57. ........................................................................................................................ 100 Scanning electron micrograph of a weathered heavy mineral grain from the A horizon of the Kalkaska pedon with tube-like organic material present. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. xiv Figure 58. ........................................................................................................................ 100 Scanning electron micrograph of tube-like organic material indicated by the circle in Figure 57. Figure 59. ........................................................................................................................ 101 Energy-dispersive spectrograph of tube-like organic material indicated by the circle in Figure 57 showing high C content. The Au is the signature of the coating used to prepare the grain. Figure 60. ........................................................................................................................ 102 Scanning electron micrograph of a weathered heavy mineral grain from the E horizon of the Kalkaska pedon with more weathering product than the A horizon. This product is an aluminosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 61. ........................................................................................................................ 102 Energy-dispersive spectrograph of the weathering product indicated the by circle in Figure 60. The high Si content with Al and 0 led to the determination that it is an aluminosilicate. The Au is the signature of the coating used to prepare the grain. Figure 62. ........................................................................................................................ 103 Scanning electron micrograph of a weathered heavy mineral grain from the Bhs horizon of the Kalkaska pedon with more weathering product and organic material than the E horizon. This weathering product is an aluminosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 63. ........................................................................................................................ 103 Energy-dispersive spectrograph of the weathering product indicated the by circle in Figure 62. The high Si content with Al and 0 led to the determination that it is an aluminosilicate. The Au is the signature of the coating used to prepare the grain. Figure 64. ........................................................................................................................ 104 Scanning electron micrograph of a weathered heavy mineral grain from the B51 horizon of the Kalkaska pedon with an extensive coating of weathering product. This product is an aluminosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 65. ........................................................................................................................ 104 Energy-dispersive spectrograph of the weathering product indicated by the circle in Figure 64. The high Si content with Al and 0 led to the determination that it is an aluminosilicate. The Au is the signature of the coating used to prepare the grain. XV Figure 66. ........................................................................................................................ 105 Scanning electron micrograph of a weathered heavy mineral grain from the B81 horizon of the Kalkaska pedon with string-like features that were determined to not be organic due to the EDS having the same signature as the weathering product. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 67. ........................................................................................................................ 105 Scanning electron micrograph of string-like features indicated by the circle in Figure 66. Figure 68. ........................................................................................................................ 106 Energy-dispersive spectrograph of the string-like features indicated the by circle in Figure 66. The features were determined to not be organic due to the EDS having the same signature as the weathering product. The Au is the signature of the coating used to prepare the grain. xvi KEY TO SYMBOLS OR ABBREVIATIONS um .................. micrometer AAO ............... acidified ammonium-oxalate AAS ................ atomic absorption spectrometer CD .................. sodium citrate-dithionite EDS ................ energy-dispersive spectrograph F ed and AL; ..... iron and aluminum extracted with sodium citrate-dithionite Fe0 and A10 ..... iron and aluminum extracted with acidified ammonium-oxalate Fep and Alp ..... iron and aluminum extracted with sodium pyrophosphate PP ................... sodium pyrophosphate SEM ............... scanning electron micrograph vs. ................... versus yr B.P. ............ years before present xvii INTRODUCTION Homblende Etching as a Relative Age Indicator Homblende etching holds great promise as a relative weathering indicator. Homblende, a member of the amphibole class of minerals, is common and abundant on the earth’s surface (Blackburn and Dennen, 1988). Natural weathering of hornblende corrodes the hornblende in a crystallographically controlled manner, resulting in etch pits onthe surface of the mineral and a “sawtooth” or “denticulated” margin on which the “teeth” are parallel to the c-axis (e.g., Bemer et al., 1980; Bemer and Schott, 1982; Velbel, 1987, 1989) (Figure 1). Progressive weathering causes these features to gradually and systematically increase in depth and size from a well-defined starting point (Locke, 1979, 1986; Birkeland, 1978), the Antarctic (Ugolini and Bull, 1965; Everett, 1971; Locke, 1979, 1986). Using this knowledge, hornblende etching has been applied / 20pm Figure 1. Scanning electron micrograph showing naturally weathered hornblende with etch pits, denticulated margin, and “teeth” parallel to the c-axis. Arrow indicates c-axis direction. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. as a relative age indicator of soils in the Rocky Mountains of the United States (Hall and Heiny, 1983; Hall and Martin, 1986; Hall and Michaud, 1988), the Canadian Arctic (Everett, 1971), and Italy (Andrews and Miller, 1980). This was accomplished by statistically comparing the depth of hornblende denticulation, or denticulation amplitude, in soils and sediments of different ages (Locke, 1979). Homblende from the older sites exhibited a greater degree of etching, as measured by greater amplitude, than the hornblende from the younger sites (Locke, 1979). Thus, it was concluded that hornblende could be used to determine relative age. In the previously mentioned studies, the researchers assumed that time of exposure, or duration of weathering, was the primary factor that influenced denticulation amplitude. These studies were based on the idea that etch features enlarge over time without taking into consideration other factors that may affect this enlargement. Factors, other than time, that may significantly contribute to this weathering process include: climate, amount of water present, pH of soil, presence and abundance of clay, and presence of organic matter (White and Brantley, 1995). Previous work has only emphasized the relationship between hornblende denticulation and the duration of weathering. However, hornblende denticulation amplitude is a cumulative record of weathering, which reflects both the weathering intensity and the weathering duration. Thus, denticulation amplitude measurements might serve as a relative weathering indicator among soils of the same age but differing degrees of development. Quartz/Feldspar Weathering Ratio The quartz/feldspar ratio is a measure used to determine the degree of mineral weathering based on the persistence of quartz compared to that of feldspar (e. g., Birkeland, 1974; Muhs 1982; Bockheim et al., 1996). From Goldich’s (1938) study, we know that a sediment will become depleted in feldspar with respect to quartz as weathering continues because feldspar is less persistent than quartz. Since the hornblende denticulation amplitude may also indicate the degree of mineral weathering, these two measurements can provide information regarding the degree of mineral weathering. Soil Development Many researchers have assumed that mineral weathering and soil development occur simultaneously. Due to this assumption, or poor word choice, the terms “weathering” and “soil development” have often been used interchangeably in the literature (e. g., Ruhe, 1956; Locke, 1979; Andrews and Miller, 1980). However, weathering and soil development are two different but inter-related processes (e. g., F anning and Fanning, 1989). Weathering is the alteration of rock and minerals, at or near the Earth’s surface, due to environmental conditions (Bland and Rolls, 1998). On the other hand, soil development is the amount of pedologic change, including horizonation, that has taken place in a parent material (Birkeland, 1999) and is a function of climate, biota, topography, parent material, and time (Jenny, 1941). It is possible that soil development and weathering will correspond to each other, and that a measure of one may proxy for a measure of the other, but the two measures are not necessarily providing the same information about the soil. If hornblende denticulation amplitude can produce the same trend with regards to relative weathering as soil development does with regards to relative profile development, hornblende denticulation amplitude could be used as a proxy for soil development. Research Questions Soil development is a function of several factors; these include climate, biota, topography, parent material, and time (Jenny, 1941). The influence of any one factor is studied by examining a sequence of soils in which all but one of these soil-forming factors are held as constant as possible, so that only one factor varies in the sequence (Birkeland, 1999). Using this approach, the degree of hornblende weathering, as measured by denticulation amplitude, has been used to determine relative age of soils on glacial parent materials in studies where time or soil age is the only soil-forming factor allowed to vary (e.g., Hall and Heiny, 1983; Locke, 1979, 1986; Birkeland, 1978). However, the influence of soil-forming factors other than time on hornblende weathering remains to be examined. As Fanning and Fanning (1989) point out, soil-development involves a variety of processes, of which weathering is only one, and some aspects of soil development can proceed without progress in mineral weathering, and vice versa. Thus, a number of questions relating hornblende weathering to soil development remain. These include: 1) is hornblende weathering (as one type of mineral weathering) related to or independent of soil development? and 2) is hornblende denticulation amplitude comparable to other measures of mineral weathering? Purpose of Study This research was designed to examine the relationship between mineral weathering in general, homblende weathering in particular, and soil development. In order to examine these relationships, four soils, similar in age but at various stages of development, from northern lower Michigan were sampled for this study. Denticulation amplitude measurements were taken from each horizon in every pedon. A plot of horizon vs. denticulation amplitude was then constructed to determine if any correlations could be made. A quartz/feldspar ratio was also determined for each horizon, and a horizon vs. quartz/ feldspar ratio plot was then constructed and examined to determine if correlations could be drawn from these plots. If it is determined that correlations may be drawn between hornblende denticulation amplitude, or quartz/feldspar ratios, and soil development, it would provide researchers with other analytical tools that might be useful in situations where traditional analysis do not yield conclusive results. CHAPTER 2 Homblende Weathering General Description of Homblende Homblende is the most common mineral species of the amphibole group. It is a ferromagnesian mineral in the inosilicate mineral class (Blackburn and Dennen, 1988), having the general composition of [(Na,K)0-lCa2(Mg,Fe2+,Fe3+,Al)5Sim,5Alo,5-2022(OH)2] (Brantley and Chen, 1995). Inosilicates are characterized by silica tetrahedra that are linked in double chains by sharing alternately two and three oxygen atoms per tetrahedron (Brantley and Chen, 1995) Blackburn and Dennen (1988) describe hornblende as black to dark green, and it generally appears columnar, bladed, or fibrous, but can also be granular massive. The crystallography of hornblende is monoclinic, 2/m. Its crystals are perfect prismatic {110} at 56° and 124°, with good cleavage also at 56° and 124°. Homblende has subconchoidal to uneven fracture, with a hardness of 5 - 6, and a specific gravity of 2.9 - 3.3. Homblende Distribution Homblende is widely distributed in igneous rocks and is often associated with quartz, feldspar, and pyroxene. It is also a common component of metamorphosed basaltic rocks (Blackburn and Dennen, 1988). Goldich (193 8) found hornblende to be more persistent than olivine, pyroxene, or calcic plagioclase during weathering, but less persistent than micas, alkali feldspars, or quartz. Because it is common in both igneous and metamorphic rocks, it is abundant in transported material as well. Several studies have found that it is the most abundant of the heavy minerals in soils formed in a variety of parent materials, including hornblende schist in central Texas (Stahnke, 1965, as cited in Allen and Hajek, 1989), loess-mantled till in southwestern Iowa (Ruhe, 1956), glacial till in southwestern Montana (Hall and Martin, 1986), glacial till in Ontario, Canada (Dreimanis et al., 1957), glacial till in the northeastern United States (Johnson and Chu, 1983, as cited in Allen and Hajek, 1989), soils in the Great Lakes region (Dworkin et al., 1985), glacial till in south-central Michigan (Haile-Mariam and Mokma, 1996), and sandy outwash material of northern Michigan (Haile-Mariam and Mokma, 1995). Because hornblende is common and widespread, better understanding of its weathering is potentially widely applicable. Weathering Environment Soil environments that are well drained and leached are more conducive to weathering than environments that are poorly drained (Cremeens et al., 1992; Haile— Mariam and Mokma, 1996,1995). Bemer (1978) found that flushing water through rock, sediment, or soil removed dissolved weathering products so that they are no longer in contact with the mineral, thus increasing the dissolution rate. He went on to state that with continuous flushing, a limit could be reached where the flushing rate is no longer a controlling factor, but the mineral reactivity controls the rate. This occurs when sufficient water passes through the rock or soil to remove any weathering product, at which point the factor that controls the dissolution rate is the reactivity of the mineral. In the unsaturated zone of a soil, flushing controls the dissolution rate due to an insufficient amount of water. Therefore, the most influential climatic factors on hornblende etching are effective precipitation and infiltration of unfrozen water (Locke, 1979). Homblende Weathering Rates Weathering rates have been determined for naturally weathered hornblende by studying the mean maximum etching depth, or MMED, (Locke, 1979, 1986) and etch pit distribution (Cremeens et al., 1992; MacInnis and Brantley, 1993). These rates were generally found to be 101 — 104 times slower than the experimentally determined rates (MacInnis and Brantley, 1993; White et al., 1996). The general discrepancy between natural weathering rates, and experimentally determined rates, is discussed in detail by Velbel (1993a). Brantley et al. (1993) provide possible explanations for the difference between weathering rates determined in the laboratory and those determined by naturally weathered hornblende, such as: lack of understanding of reaction mechanisms, experimental error, inaccuracies in reactive surface area estimation, inaccuracies in wetted surface area estimation, temperature variability, and biological processes. Another possible reason for the discrepancy in these rates is the changing dissolution rates in the natural environment due to rate constants that decrease with time (White et al., 1996; Hall and Martin, 1986). Murakami et al. (1998) also found a discrepancy between natural rates and laboratory determined rates of weathering in anorthite due to the decreasing weathering rate in nature. However, they concluded that these rates decreased due to the saturation effect caused by secondary minerals between grains. Not only are weathering rates slower in nature relative to laboratory experiments, but they also differ between the upper part of the profile where the rate is rapid, and the lower part of the profile where the rate is slow (Markos, 1975). Several possible factors that could affect the dissolution rate of minerals in the natural environment include: precipitation of dissolved silica, iron precipitates, and organic and inorganic coatings on grain surfaces (Dove, 1995). Nugent et a1. (1998) found thin, hydrous, patchy natural coatings of amorphous and crystalline aluminosilicate on feldspar grains that were only visible under atomic force microscopy. They concluded that these coatings were partially inhibiting dissolution and thus contributing to the discrepancy between natural and laboratory determined weathering rates. However, Velbel (1993b) argued that examination of naturally weathered silicate minerals reveals etch pits, which are features produced when chemical reactions at the interface control the weathering rate. The presence of etch pits on surfaces beneath clay coatings or oxide and/or hydroxide products indicate diffusion through a surface layer is not rate-limiting. Similarly, Bemer and Schott (1982) found that cation-depleted layers on soil grains were too thin to be considered diffusion-limiting, and therefore weathering is controlled by surface chemistry and not by diffusion through a surface layer. However, precipitates and alteration products could slow dissolution rates by blocking or altering flow paths, and thus decrease the amount of water that reaches the mineral (Eggleton, 1986, as cited by Brantley and Chen, 1995; Velbel, 1993b; Murakami et al., 1998). Homblende Weathering Features Natural weathering of hornblende forms etch pits on the surface of the grain. These dissolution voids may be seen using a petrographic microscope. Velbel (1989) noted that hornblende weathering progresses through a sequence of features that indicate the extent of weathering. The earliest feature of weathering (under oxidizing conditions) is the presence of ferruginous material along grain boundaries or in fractures within the grain observed in thin section. This chemical precipitation of goethite, gibbsite, or kaolinite produces a boxwork structure. The more extensively weathered hornblende exhibits crystallographically controlled etch pits, which form parallel to the c-axis in the direction of cleavage (Figure 2). This feature is formed by the selective weathering at crystal defects and dislocations (Bemer et al., 1980; Velbel, 1993b). More advanced weathering produces a sharp denticulated margin, also known as cockscomb termination, hacksaw termination (e.g., Bemer et al., 1980; Bemer and Schott, 1982; Velbel, 1987, 1989, 1993b), or “sawtooth” termination (Velbel, 1989) (Figure 3). These denticulations result from side-by-side coalescence of lenticular, or lens-shaped, etch pits (Bemer et al., 1980; Bemer and Schott, 1982; Velbel, 1993b). Velbel (1989) goes on to state that even more advanced weathering produces hornblende remnants that occupy dissolution cavities bounded by ferruginous boxworks (Figure 4). These are void spaces that lack weathering product, either in the cavity or on the hornblende surface. If the parent hornblende is removed through dissolution, a “negative pseudomorph” or alveoporomorph (terminology of Delvigne, 1998) is left (Figures 5 and 6). An interesting feature of hornblende weathering observed by Velbel (1989) was pendants of ferruginous material that were separated from the hornblende remnants by void space (Figure 4). This separation suggests that the pendants formed by dissolution-reprecipitation, or neoformation. Homblende etch pits progress through a sequence of stages (Bemer et al., 1980; Bemer and Schott, 1982; Cremeens et al., 1992). These stages begin with the formation 10 of one or more lens-shaped etch pits that enlarge with weathering (Figure 7). The etch pits then coalesce either end-to-end or side-by-side. End-to-end coalescence results in the formation of longitudinal grooves or striations (Figure 8). It is the side-by-side coalescence of etch pits that form the denticulated margin (Figure 3). These tooth-like denticulations are unweathered hornblende, which are adjacent to pits that formed by the dissolution of the hornblende. Etch pits are crystallographically controlled dissolution voids, produced by interface-limited weathering mechanisms (Velbel, 1993b, and references contained therein). Bemer and Schott (1982) put forth the idea that this mechanism consists of H” attacking the mineral at a point of excess energy, primarily at a dislocation outcrop, which is accompanied by Fe oxidation in the outermost atomic layers. This selective attack also occurs at crystal defects (Bemer et al., 1980). However, this does not penetrate inward over time, but produces lens-shaped etch pits (Bemer and Schott, 1982). 200nm Figure 2. Scanning electron micrograph of crystallographically controlled etch pits, which form parallel to the c-axis in the direction of cleavage. Arrow indicates c-axis direction. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 11 ti, ‘ 20pm Figure 3. Petrographic micrograph of “sawtooth” termination on a naturally weathered hornblende grain. Image captured with a Pixera digital imaging system. Figure 4. Scanning electron micrograph of weathered hornblende showing ferruginous microboxwork (B) separated from denticulated hornblende (D) by void space with pendants (P) projecting from the microboxwork into the void space (Figure 6 in Velbel, 1989). Figure 5. Photomicro graph showing hornblende has been completely removed, leaving only ferruginous boxwork (B) and pendants (P), or “negative pseudomorph” (Figure 9 in Velbel, 1989). Figure 6. Scanning electron micrograph showing ferruginous boxwork (B) with neoformation pendants (P) (Figure 8 in Velbel, 1989). Figure 7. Scanning electron micrograph of enlarged lens-shaped etch pits in naturally weathered hornblende. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 8. Scanning electron micrograph of end-to-end coalescence of etch pits forming grooves or striations in naturally weathered hornblende (Figure 3 in Velbel, 1989). Cremeens et al. (1992) suggested that etch pits uniform in size and shape might be associated with strain due to exsolution lamellae. They also concluded irregularly shaped and spaced pits might be associated with microcracks and fractures. Homblende Weathering in Soils The weathering of hornblende can produce different weathering features and products depending on environmental conditions. Etch pit size and shape could be an indicator of the environment in which they formed. Cremeens et a1. (1992) studied the etch pits in orthoclase and pyriboles. They observed that in poorly drained soils the minerals had small pits but a large number of them. This suggested a slower growth rate and presumably less flushing with fresh water. Greater coalescence of pits was observed in well drained soils. These pits were larger, irregularly shaped, and elongated, suggesting a greater grth rate, which they attributed to increased flushing of fresh water, which would remove the weathering products and reduce contact time of the primary mineral with its weathering products. In a study conducted in Maryland, the weathering products of hornblende, such as Fe oxides and oxyhydroxides and clay minerals, were dependant upon the intensity of leaching in the profile (Cleaves, 1974). Cleaves (1974) also found that in a poorly drained, structured saprolite developed on a quartz-plagioclase-hornblende gneiss, hornblende weathered to montrnorillonite. It weathered to ferric oxides in a well drained, structured saprolite developed on the same quartz-plagioclase-hornblende gneiss parent material. He also noted that the weathering products of hornblende in a well drained saprolite (which developed on an epidote amphibolite) included kaolinite, halloysite, 15 goethite and other ferric oxides, and isotropic amorphous material. Velbel (1989) found that hornblende weathered to goethite, gibbsite, and kaolinite in an intensely leached weathering profile on a plagioclase-hornblende gneiss of the Carroll Knob mafic complex of North Carolina. Other weathering products of hornblende include: iron-rich montmorillonite (Walker et al., 1967), vermiculite, chlorite (Stephen, 1952, as cited by Huang, 1989), beidellite (Goldich, 1938), and smectite (Rice et al., 1985). In the soil, hornblende is most commonly found in the sand and silt size fraction where it constitutes a significant portion of the heavy minerals (Ruhe et al., 1966; Locke, 1979; Allen and Hajek, 1989). Haile-Mariam and Mokma (1995, 1996) studied the mineralogy of two fine-loamy hydrosequences in south-central Michigan, and two sandy Spodosol hydrosequences also in Michigan. In all the pedons studied, they found that the percent of hornblende increased with depth, as the percent of quartz decreased with depth. They attributed this trend to the persistence of quartz compared to other minerals. In addition to mineral depletion trends, hornblende etching trends have been observed in soils. Locke (1979) conducted a study on the etching of fine sand sized hornblende collected from a coarse-grained glacial deposit from Baffin Island, Canada. He found that etching decreased logarithmically with increasing depth in the profile. He also noted that the rate of etching at any give depth decreased logarithmically with increasing soil age. He proposed that precipitation and unfrozen water infiltration influenced the etching rate as a function of depth and age. Based on these observations, Locke (1979) felt that hornblende etching showed great promise as a relative dating technique. 16 In a later study, Locke (1986), found that soils of similar age exhibit trends where hornblende etching was greatest near the surface and decreased with depth. This trend was interpreted to be due to a decrease in soil wetting events at depth. He also noted that hornblende etching showed significant differences in soils of similar age that formed under different climatic regimes. This difference in hornblende etching was attributed to differential rainfall. This precipitation theory was supported by Hall and Horn (1993), who studied soils of two chronosequences, one each from the Wind River Range in western Wyoming and the Tobacco Root Range in southwestern Montana, U.S.A. They found faster weathering rates in the soils at higher elevations due to increased orographic precipitation. 17 CHAPTER 3 Quartz and F eldspar Weathering Mineral Persistence Many studies have been conducted to determine the stability of minerals by comparing the mineral composition of the fresh parent material to its weathering product (e.g., Goldich, 1938; White et al., 1996; Birkeland, 1999). This comparison allowed researchers to quantify the relative depletion of a mineral from a system. Less stable minerals are preferentially depleted. Using relative depletion data, weathering sequences have been developed that rank minerals according to their stability (Goldich, 1938; Pettijohn, 1941, as cited in Bland and Rolls, 1998; Ollier and Pain, 1996). Samuel Goldich (193 8) developed the weathering sequence that is still the standard today. Goldich (193 8) developed what he called a mineral-stability series in weathering by conducting extensive analyses on fresh and weathered materials from several sites. He studied a fresh amphibolite from the Black Hills of South Dakota and found it to be 77 percent hornblende while its residual material was only 7 percent hornblende. He concluded that homblende was undergoing extensive alteration to beidellite or related clay minerals. Goldich (193 8) also studied the Morton Gneiss in Minnesota. He concluded that the heavy minerals of this granite gneiss had been largely destroyed and that the remaining minerals represented the relatively resistant or more stable minerals. The least stable minerals were plagioclase, hornblende, epidote, titanite, and apatite, whereas ilmenite-magnetite, biotite, microcline, and orthoclase were more resistant. Quartz and 18 zircon were the most stable minerals, with even these showing signs of being altered. Additionally, Goldich (193 8) analyzed the Medford diabase from eastern Massachusetts. He found an absolute loss of heavy minerals as a result of weathering. He also noted the relative abundances of heavy minerals in the fresh and weathered diabase suggested that the rate of weathering differed for various minerals. From this diabase, Goldich ranked the minerals from least to most stable: chlorite, pyroxene, epidote, hornblende, biotite, apatite, and magnetite-ilmenite. From this classic study, Goldich (1938) proposed a mineral-stability series for weathering in which the minerals were ranked from least stable to most stable as: olivine, augite, hornblende, biotite, potassium feldspar, muscovite, and quartz (Figure 9). The plagioclase series is separate from the mafic series with sodic plagioclase being more stable than calcic plagioclase. As a group, the mafic minerals are less stable than felsic minerals. MINERAL-STABILITY SERIES IN WEATHERING Olivine Calcic plagioclase Augite Calcic-alkalic plagioclase Homblende Alkali-calcic plagioclase Alkalic plagioclase Biotite Potash feldspar Muscovite Quartz Figure 9. Mineral-Stability Series in Weathering proposed by Goldich (1938, Table 18). 19 Despite its historical precedence, Velbel (1999) points out that the term “stability series” is a misnomer. The series proposed by Goldich (1938) and the many that followed (e.g., Pettijohn, 1941, as cited in Bland and Rolls, 1998; Dryden and Dryden, 1946; Nickel, 1973; Morton, 1984; Bateman and Catt, 1985; Allen and Hajek, 1989; White et al., 1996; Velbel, 1999) rank the persistence of different minerals. These rankings are due to the kinetics of the reactions, not the thermodynamic stability of the mineral. A more accurate term for these series should be “mineral persistence series”. Quartz and F eldspar Two relatively abundant minerals that are commonly compared, and which were chosen for this study, are quartz and feldspar. A soil will become depleted in feldspar, with respect to quartz, because feldspar is less resistant to weathering than quartz (e. g., Goldich, 1938; Ollier and Pain, 1996; Birkeland, 1999). Quartz, with the formula of SiOz, is a member of the silica group of tectosilicates. The basic building block of silicate minerals is the tetrahedron, where a Si4+ is surrounded by four 02' ions. The tectosilicates are those silicate minerals in which all oxygens are shared with adjacent tetrahedra to form a network with the net charge of zero (Blackburn and Dennen, 1988). Blackburn and Dennen (1988) provide a summary of crystallography, physical properties, and occurrence of quartz. Quartz forms hexagonal prisms with pyramidal termination and can have either lefi- or right-handed forms. It is usually found massive or as crystals. Quartz has no cleavage and no parting. It has irregular to conchoidal fracture. Quartz has a hardness of 7 and a specific gravity of 2.65. Quartz is one of the most abundant minerals and is an important constituent in many 20 igneous, metamorphic, and sedimentary rocks. It is a major component in some rocks, such as sandstone and quartzite. Quartz is often the sole constituent in veins and is almost universally found in soils and sediments. Ollier and Pain (1996) note that quartz is often the last mineral to crystallize from a magma, and due to this lower temperature of crystallization, it is fairly stable at earth surface temperatures. Although quartz is extremely resistant to chemical weathering, physical weathering by wind or water can round quartz grains. Due to its resistance to chemical weathering, quartz can be a standard by which the weathering of other grains can be measured. Ollier and Pain (1996) state that due to its persistence, the proportion of quartz increases up profile because as other minerals are removed higher in the profile, quartz remains. Feldspars are aluminosilicate minerals, with the alkali feldspars having the general formula (K,Na)(AlSi3Og) and the plagioclase feldspars having the general formula (Na,Ca)(Al,Si)4Og (Blackburn and Dennen, 1988). Blackburn and Dennen (1988) provide a summary of the feldspars. The distinction between the alkali feldspars and the plagioclase feldspars is not always clear-cut due to solid solution. However, the alkali feldspars are monoclinic, or, if triclinic, have more K than Ca. The plagioclase feldspars are triclinic with less K than Na or Ca. Due to the solid solution, a wide range of chemical composition is permissible. Twinning is very common among the feldspars, occuning as simple or repeated twins. The feldspars make up almost 60% of igneous and metamorphic rocks, which constitute over 90% of the earth’s crust. Feldspars are found in the light fraction of soils and have a specific gravity of 2.6. Ollier and Pain (1996) note that feldspars alter rapidly due to the pronounced cleavage that probably allows rapid penetration of water. Feldspars often alter to kaolin, but other secondary minerals, 21 such as mica and allophane, can also be formed. Quartz/Feldspar Ratio Quartz/ feldspar ratios examine the relationship between a resistant mineral and a non-resistant mineral. In the soil environment, feldspar is more quickly leached than quartz (Nesbitt et al., 1997). Nesbitt et al. (1997) describe this relationship in a simplified model that has plagioclase and K-feldspar leached at approximately the same rate, which is 100 times faster than quartz. When dissolution begins, the soil quickly reaches silica saturation with respect to quartz. This saturation limit causes quartz to stop dissolving, however feldspar and plagioclase continue dissolving. Generally, silica saturation with respect to K-feldspar is reached next allowing only plagioclase to continue dissolving. They also note that the soil environment is critical in these reactions because the amount of organic acids present determines the amount of mineral dissolution. Studies have shown that the feldspar content of a soil decreases in the upper horizons due to an increasing of weathering intensity (e. g., Nesbitt et al., 1997; Allen and Haj ek, 1989). Birkeland (1974) stated that this observation is one way of determining weathering rates. Rate determination is accomplished by establishing the ratio of a resistant mineral (e. g. quartz) to a nonresistant mineral (e. g. feldspar) in the parent material and comparing that to the ratio of the same minerals in a horizon of known age. This gives the rate at which the less resistant mineral is depleted. This is the technique Ruhe (1956) used in Iowa when estimating that the Wisconsin surface had weathered 6800 years, the soil of the late-Sangamon surface at least 13,000 years, and the soil 22 related to the Yarmouth-Sangamon surface 105 years or more. Relative age may also be determined by using quartz/ feldspar ratios (Dorronsoro and Alonso, 1994). Dorronsoro and Alonso (1994) concluded soils of different ages would have different quartz/ feldspar ratios under conditions that hold other soil forming factors constant. The soil with the larger ratios, indicating more depletion of feldspar, is the older soil, thus providing relative age. Quartz/feldspar ratios also serve as a relative weathering indicator (Ruhe, 1956; Nesbitt and Markovics, 1997). Ruhe (1956) noted that a comparison of the weathering ratios for each site in his study indicated the intensity of weathering was greatest in the older surface. The older surface had ratios with higher values, which suggested there were greater amounts of resistant minerals relative to less resistant minerals. These conclusions indicate the quartz/feldspar ratio could be used as a relative weathering indicator. Quartz/ feldspar ratios have also been used to determine parent material uniformity due to the stability of quartz and abundance of both quartz and feldspar (Wilding et al., 1977; Osei, 1992). Lithological discontinuities can be detected using quartz/feldspar ratios (Price et al., 1975; Osei, 1992). Price et a1. (1975) were able to distinguish the boundary between an overlying loess and the underlying residuum by constructing a depth vs. quartz/ feldspar ratio plot for three profiles in their research area. They noted there was a fairly uniform quartz/feldspar ratio to a depth of about 100cm at which point there was a large increase in the ratio. They concluded this large increase indicated a lithological discontinuity. 23 CHAPTER 4 Spodosols Podzolization Podzolization is the pedogenic process by which Fe, Al, and organic carbon are translocated down the soil profile from the O, A and E horizons to an illuvial horizon. The driving force behind this process is the percolation of acidic water through the soil profile. Infiltrating water becomes acidic as it passes through the decaying leaf litter of the O horizon. This process produces Podzol or Spodosol profiles (DeConinck, 1980; Buurman and van Reeuwijk, 1984; Schaetzl and Isard, 1991; Mokma and Evans, 2000; Schaetzl, 2002). The podzolization process can be summarized in the following four steps: (1) organic material on the surface decays and fulvic acids are produced which are carried by the water as it percolates through the profile, (2) organic matter then complexes with and mobilizes F e3 + and Al3+ cations, (3) these mobile cations are translocated from the O, A and E horizons to the B horizon where (4) the Fe and A1 are deposited (Mokma and Evans, 2000). The fulvic acids, percolating through the soil with the water, acidify the upper horizons. This acidic environment causes intense weathering of the minerals, which release cations; of particular interest are Fe and Al. The Fe and A1 are then translocated down the profile (Mokma and Evans, 2000). There are two current theories to explain the translocation of Fe and Al during podzolization. One theory states that the fulvic acids chelate the Fe and A1 to form organometallic complexes that are then translocated 24 in percolating water. Once the complexes reach the water table, or a less permeable layer, microbes break down the fulvic acid thus causing the precipitation of the Fe and Al (DeConinck, 1980; Buurman and van Reeuwijk, 1984). The other theory is the “proto- imogolite” theory. This theory was developed by researchers who observed that Fe and Al exist in amorphous, inorganic compounds such as imogolite and allophane. According to this view, the Fe and A1 are translocated into the B horizon as proto- imogolite first, and then organic matter precipitates onto the Fe and Al (Farmer et al., 1980; Farmer, 1982; Anderson et al., 1982). Some aspect of each theory, and possibly a combination of the two, might better explain the translocation of Fe and A1 during podzolization rather than either theory independently (Wang et al., 1986; J akobsen, 1991; Schaetzl, 2002). Factors Contributing to Podzolization Several factors contribute to the rate of podzolization. These factors include: (1) climate, especially snowmelt infiltration, (2) parent material texture, (3) parent material composition, and (4) type of vegetation in the region. Climate affects the rate of podzolization by determining not only the amount of precipitation, but just as importantly, it determines whether or not that precipitation infiltrates. Schaetzl and Isard (1996) observed that regions that received little snow during the winter had a slower rate of podzolization than regions that received enough snow to accumulate a deep winter-long snowpack. The soil under deep snowpack is insulated and does not freeze often, unlike soil that has little snow accumulation (Isard and Schaetzl, 1998). The insulated soil is impacted by a high amount of infiltration of the 25 snowmelt during spring, due to the quantity of meltwater available and the permeability of the non-frozen soil. The frozen soil does not allow for infiltration during spring until it has thawed, by which time there may be little or no snowmelt remaining to enter the profile (Schaetzl and Isard, 1996). Because long, continuous influxes of meltwater contribute to the rate of podzolization, climate is an important factor. Parent material texture is a significant control on podzolization. The best expressed Spodosol profiles are in coarse-textured soils (Gardner and Whiteside, 1952; Messenger et al., 1972; Vance et al., 1986; Schaetzl and Isard, 1991). The coarse texture results in more rapid permeability and higher infiltration rate of the acidic water. Podzolization is inhibited by high clay contents. Mokma and Evans (2000) stated that clay might slow or even stop the translocation of Fe and Al by adsorbing the Fe and Al cations, thus making them immobile. Due to this, the E and B horizons will not thicken quickly and the B horizons will redden slowly. Another way clay interrupts the process is by weathering so quickly in the acidic water that it releases more cations than the system can handle. The few chelating acid molecules get overwhelmed and immobilized, stopping the translocation. For these reasons, clay in the soil parent material must be minimal or weathered extensively by acidic water prior to podzolization. In the best expressed Spodosols, the parent material is sandy and composed primarily of quartz with lesser amounts orthoclase, plagioclase, pyroxene, amphibole, mica, and opaque minerals (Haile-Mariam and Mokma, 1995; Mokma and Evans, 2000). Carbonate minerals are notably few in the parent material. Any carbonate minerals present will react with the percolating water to neutralize the acid and slow the rate of podzolization (Mokma and Evans, 2000). Therefore, most of the carbonates must be 26 weathered and leached from the system for podzolization to progress. Vegetation also contributes to the rate of podzolization. Schaetzl (2002) noted that certain types of coniferous vegetation, such as pine and hemlock, accelerate podzolization. The leaf-litter under conifers is composed of needles, which are more acid than the leaf-litter under broadleaf trees. Decay of the conifer needles produces more of the acids needed for podzolization. For this reason, Spodosols are associated with coniferous and coniferous-deciduous forests. The Spodosol Profile Spodosols have four major horizons (Mokma and Evans, 2000). The O and/or A horizon is dark brown to black due to the high organic content. The process of podzolization leaves a gray to ashy white eluvial zone, or E horizon, from which Fe and Al have been removed. The B horizon is below the A horizon and above the B horizon, which is a dark reddish brown illuvial zone where the Fe and A1 have accumulated. The illuvial zone is a spodic (Bs or Bhs) horizon. There may be several B subhorizons; those directly below the B horizon are usually reddest and darkest becoming more yellow down profile. The underlying C horizon usually has an even more yellow hue than the lowest B horizon. As a Spodosol develops, the E horizon becomes whiter while the Bs or Bhs horizon continues to darken and thicken. The POD index is a numerical index of Spodosol development based on this pathway of development (Schaetzl and Mokma, 1988). This number is calculated using the differences of Munsell hues between the eluvial and illuvial horizons. Due to the requirement of an eluvial horizon being present 27 in the soil, a POD index cannot be calculated for a pedon lacking an E horizon. If both eluvial and illuvial horizons are present, redder hues in the B horizons, and more B subhorizons, will result in a higher POD index. A higher number indicates a more strongly developed Spodosol. Extractable Iron and Aluminum in Spodosols Iron and aluminum exist in soils in various pedogenic forms. When identifying a spodic horizon, the Fe and Al content must be determined because the horizon is characterized by the accumulation of translocated Fe and A1 (McKeague and Day, 1965; Soil Survey Staff, 1999). Three extractants are commonly used in this determination. Sodium citrate-dithionite (CD) is used because it primarily extracts “free” Fe, denoted by Fed, from pedogenic iron oxides, and “free” Al, denoted by Ald, from aluminum oxides (Birkeland, 1999; Weisenbom, 2001). The term “free” is used to describe compounds having a single species of coordinating cation (Jackson et al., 1986; Weisenbom, 2001). The term is also used for an iron oxide that can be reduced and dissolved by a dithionite treatment (Anonymous, 1996; Weisenbom, 2001). Acidified ammonium-oxalate (AAO) extracts “active” Fe, denoted by Fee, and “active” Al, denoted by A10, from noncrystalline hydrous oxides. The term “active” is given to oxides that are small in size, have a high surface area, and a high degree of reactivity (Loeppert and Inskeep, 1996; Weisenbom, 2001). AAO also extracts Al from noncrystalline, or amorphous, aluminosilicates such as allophane and imogolite and other weathering products (Farmer et al., 1983). F eO selects Fe mainly from ferrihydrite (F65H0804H20), and is such a good test that its value may be multiplied by 1.7 to estimate the amount of ferrihydrite present (Parfitt and 28 Childs, 1988; Birkeland, 1999). The third extractant is sodium pyrophosphate (PP). It extracts Fe, denoted by Fep, and Al, denoted by Alp, from organically-bound complexes and, to a lesser degree, noncrystalline hydrous oxides. It has been reported and long assumed that these three extractants do not extract Fe and Al from crystalline minerals (Mehra and Jackson, 1960; Ross and Wang, 1993; McKeague et al., 1971). In addition to analyzing the extraction data directly, various calculations and formulations may be made from the raw data. These calculations indicate the form of Fe and Al that is found in each horizon (e. g., crystalline or noncrystalline, complexed or not). The ratio of AAO / CD, or “activity ratio” (Blume and Schwertmann, 1969), measures the amount of pedogenic Fe that is amorphous, plus fen'ihydrite (Birkeland, 1999). It approximates the relative crystallinity of “free” oxides. A low value of Feo/ F ed indicates Fe oxides with a high degree of crystallization or aging (Blume and Schwertmann, 1969; Weisenbom, 2001), while a high value from this ratio indicates Fe is in a more “active” noncrystalline form that is complexed either organically or inorganically (Weisenbom, 2001 ). The molar ratio of Fep / A1p is the difference between organically complexed, noncrystalline Fe and A1 (Barrett, 1995; Gustafsson et al., 1995). Ratio values less than one signify there is more organically complexed, noncrystalline Al than Fe in that horizon (Weisenbom, 2001), while a ratio of greater than 10.5 could indicate that imogolite is not likely to form due to insufficient Al (Huang, 1991). The noncrystalline, organically complexed forms of Fe and Al versus the noncrystalline, inorganic complexed forms of Fe and Al in a horizon may be estimated by (Fe0 — Fep) / Fep and (A10 — Alp) / Alp (Wang et al., 1986; Wang and Kodama, 1986; 29 Schaetzl, 1992). Higher calculated values signify that inorganic complexes contribute more than organic complexes to the state of Fe and Al in the soil. Conversely, lower calculated values emphasize the role of organic complexes over inorganic complexes. The relative amount of crystalline “free” oxides compared to noncrystalline “free” oxides may be determined by the difference between the CD value and the AAO value (McKeague et al., 1971; Parfitt and Childs, 1988). When the CD values are larger than the AAO values, crystalline oxides are present (Parfitt and Childs, 1988). Thus, this calculation can, under favorable circumstances, be used to estimate the Fe in goethite, lepidocrocite, and hematite (Parfitt and Childs, 1988; McKeague et al., 1971). The value is negative when the AAO extracts a large amount of “active” noncrystalline oxides and organically bound noncrystalline forms, and which exceeds the CD extracted “free” value. Thus, a low, or negative, value indicates that the “free” oxides are predominantly in noncrystalline form (McKeague et al., 1971). 30 CHAPTER 5 Study Area Site Location and Description The study area for this research was located in the northern Michigan county of Kalkaska (Figure 10). Three of the four sites chosen for this study were within Blue Lake Township; the fourth site was in neighboring Bear Lake Township (Figure 11). All of the sites were located on the sandur east of the Outer Port Huron moraine (Figure 12). Three soil series are present at the four sampling sites; these are the Grayling, Rubicon, and Kalkaska series. Site Selection Criteria The four sites used in this research were chosen based on several criteria intended to constrain the soil-forming factors, which are: climate, organics or biota, relief or topography, parent material, and time (Jenny 1941, as cited in Birkeland, 1999). Two of these factors (climate and organics) are active (dynamic, in that they can change with time) in the soil forming process and really cannot be constrained. Even though these factors cannot be fully constrained, their effects were constrained, and between—site variation in these factors was held to a minimum by choosing sites that were in close proximity to one another. The passive factors (relief, parent material, and time) were accounted for by holding them as constant as possible among the sites. Due to the higher elevations of the moraine, this region is within a major lake effect snowbelt (Schaetzl, 2002). As previously stated, in order to minimize the effects 31 Kalkaska County Figure 10. A map of counties in lower Michigan with Kalkaska County highlighted. Silt‘ l . ‘ T ' r? t ' 3 Rapid ' . T Cdld I Sitcl' "1 iii 1 . w . . ' atwat r River I I . .31“ We MW .- if - ..Tb bwnshib » . . at, r at. I- i l- i A 51103 i I i; i i I - i l l .. MW'Qurokaalka TTYTY ii“? Township Township 3m 4‘ VII “..1 ‘ntrN‘mMnme: t i . . i t .. i .‘t. I Boardman I Orange: 2 bliver ' ; i : ‘ prnshrp 7 Township”? Tawnship ' ” Springfield barlleld , T Tovtnship . frownship I . . N Figure 11. A map of Kalkaska County townships with research sites indicated. Sites 1, 2, and 3 are in Blue Lake Township, while Site 4 is in Bear Lake Township. 32 Z Figure 12. Map of the morainic system (shown in black) of Michigan with a portion of the Outer Port Huron moraine and the study area for this research in north-central Michigan is indicated in red. (This map is public domain, per a personal communication with Dr. Randy Schaetzl.) 33 of climate, the distance between sites was kept to a minimum. The soil-forming factor of vegetation could not be fully constrained. Jack pine (Pinus banksiana) was primarily found on the Grayling, red (P. resinosa) and white pine (P. strobus) were found on the Rubicon, and hardwood, such as sugar maple (Acer saccharum) and yellow birch (Betula allegheniensis), was found on the Kalkaska sites. The type of vegetation present could be due to the soil’s capacity to hold water, which increases with more strongly developed soils (Shetron, 1974; Mokma and Evans, 2000). It could also be due to the history of fire (Schaetzl, 1994, 2002; Barrett, 1997). Vegetation-soil co-occurrences in the region are widespread (Mokma and Vance, 1989). The vegetation patterns are temporal and shift, so it cannot be definitively stated whether vegetation is acting as a dependent or independent variable in soil development in this region (Schaetzl, 2002). Relief was constrained by choosing sites where the ground slope at the pit was 0- 3%. Due to the relief, all the sites are well drained with no water table. Differences in the parent material were constrained by choosing sites that were all located on the sandur east of the Outer Port Huron moraine (Figure 12). This moraine was formed by the Lake Michigan Lobe, whose ice was mainly from the Superior east area of the Canadian Shield and flowed southwestward across a region north of Lake Huron, then into southern Michigan (Figure 13) (Dworkin et al., 1985). The sediments in this area are almost all glaciofluvial sand or gravel with some glaciolacustrine silt and clay; till is very rare (Blewett et al., 1993). All of the sediments were deposited by glacial meltwater during the retreat of the Port Huron readvance. Some of the sediments may have been reworked by meteoric runoff and post-glacial streams (Schaetzl, 2002). 34 . .r K Canadian Sh' Id Lake Superior L Superior ' ._ / . . “A, ._ 9 Grenville Mi I \5 a / \ / // Lake Erie / \ / ' i'l\ l by/ Figure 13. Map showing the ice movement over the Great Lakes region during the Wisconsin glacial stage. The arrows indicate ice flow and the dashed lines indicate ice lobe boundaries. The Outer Port Huron moraine was deposited by the Lake Michigan Lobe, which flowed southward from the Superior east area of the Canadian Shield (modified from Dworkin et al., 1985). 35 By choosing sites on the same deposit and in close proximity to one another, age was also constrained. The Outer Port Huron moraine contains some of the youngest Late Wisconsin deposits in southern Michigan (Blewett, 1990). The age of the morainic system was estimated at 12,960 i 350 yr B.P. (Blewett et al., 1993; Schaetzl, 2002). Based on the above-mentioned criteria, four soils from the Spodosol order were chosen (Figure 11). Site 1 is a soil from the Kalkaska series, which is a strongly developed Spodosol. Sites 2 and 3 are from the Rubicon series, which is less deve10ped than the Kalkaska. Site 2 is a strong Rubicon, while site 3 is a weak Rubicon. Site 4 is from the Grayling series and is the least developed of the soils sampled. 36 CHAPTER 6 Methods Sample Preparation The samples that were used in this research were gathered from four soil pits dug in Kalkaska County (Figures 10 and 11). The pits were dug with a backhoe, and the soils classified based on field observations. Approximately 4kg of soil were taken from each identified pedogenic horizon. The samples were air dried and sieved through an 8mm sieve to remove gravel and larger organics such as twigs and leaves. The samples were then split to obtain 1kg of sample. A soil splitter was used to ensure the sample used in testing was representative of the whole. The grain size distribution was determined for each pedon by Laura Beeman using particle size analysis, following standard laboratory procedures (Soil Survey Division Staff, 1993) (Schaetzl, 2002). The soil pH for each horizon was measured with a model #720A Orion pH meter using a 2:1 slurry of distilled water to soil. The clays were dispersed by soaking and shaking the sample overnight in a solution of sodium hexametaphosphate ([NaPO3]6-Na2CO3) and distilled water. The sample was then wet sieved through a 53 um sieve. The samples were then placed in a bath of 3% sodium hypochlorite (NaOCl) to remove organic matter, rinsed with distilled water, and dried in an oven at 105°C. The samples were then sieved to obtain the fine sand, or 125 um-250um, size fraction; this was the size fraction used for this research because the heavy minerals are concentrated in this size fraction (following Locke, 1979; Hall and Horn, 1993). 37 Homblende Denticulation Amplitude Determination The heavy fraction of the fine sand, which includes hornblende, was separated from the cleaned sample by heavy liquid separation. Approximately 42g of each sample were suspended in approximately 50mL of sodium polytungstate (3Na2WO4-9WO30HZO), which had been mixed with distilled water (following the manufacturer’s instructions) to obtain a specific gravity of 2.95. Homblende, having a specific gravity of 2.9 — 3.3 (Blackburn and Dennen, 1988), sank to the bottom of the separatory funnel and was easily extracted. The denticulation amplitude was then determined for 300 hornblende grains from each horizon. This quantity was chosen because of the high confidence limit it provides while allowing a large number of samples to be analyzed in a timely manner (Van der Plas and Tobi, 1965; Pettijohn et al., 1973). The hornblende grains were placed on a glass slide and viewed under a petrographic microscope at 200x. Immersion oil with a refractive index of 1.5 was used to aid in the viewing of the grains at this magnification (hornblende has a refractive index of 1.679 - 1.698 (Klein and Hurlbut, 1999)). To measure the denticulation amplitude, a graded ocular was used in which one scale division equals 4pm. Homblende Classification Scheme Based on amplitude, the grains were assigned a classification that was modeled after the classification scheme of Locke (1979) (Figure 14). Grains that showed no apparent etching were assigned Class 0 (Figure 15). Class 1 indicates etching was present but was less than 4pm in depth (Figure 16). Class 2 indicates the grains with 38 etching greater than 4pm but less than 8pm (Figure 17). Class 3 had etching greater than 8pm but less than 12pm, and so on (Figures 18 — 23). For statistical analysis the mean value of the depth of etching in each class was used. Class 0 was assigned a value of 0. Class 1 was assigned a value of 2pm; in class 2 the value was 6pm; and in class 3, 10pm; and so on. These values were then used to calculate a weighted average of denticulation amplitude for each horizon. This was accomplished by multiplying the class value times the number of grains in that class, summing these products, then dividing that sum by the total number of grains counted, which was 300 per horizon. Unexpected Etching Features Unexpected etching features were observed while examining grains from the A and E horizons of the Kalkaska, strong Rubicon, and weak Rubicon pedons. On the average, 1 - 3 grains from each of these horizons had dissolution holes through the middle of them (Figure 24). These holes may reflect hornblende dissolution around inclusions, and/or removal of inclusions. Other grains fi'om these same horizons, except for the E horizon of the weak Rubicon pedon, had dissolution features that caused the grains to appear as though they were fracturing (Figure 25). Dissolution at multiple dislocations arrayed in a plane could also produce this corrosion form: side-to-side coalescences of en echelon etch pits is known to form such features (Bemer and Schott, 1982; Velbel, 1989). Several grains exhibited both the dissolution hole and fracture feature (Figure 26). A method for measuring these features was not taken into consideration when designing this study simply because of the lack of knowledge at the time with regard to such features. The research was designed for measurements to be 39 Figure 14. Diagrammatic etching scale showing progress of etching on broken (left) and rounded (right) faces. Depth of etching is shown by indentation in grain margin and is highlighted by shadow (Figure 2 in Locke, 1979). 40 20pm Figure 15. Class 0. Petrographic micrograph of a hornblende grain with no apparent etching. Image captured with a Pixera digital imaging system. 20pm Figure 16. Class 1. Petrographic micrograph of a hornblende grain with etching between 0pm — 4pm. Image captured with a Pixera digital imaging system. 41 —' 20pm Figure 17. Class 2. Petrographic micrograph of a hornblende grain with etching between 4pm — 8pm. Image captured with a Pixera digital imaging system. 20pm Figure 18. Class 3. Petrographic micrograph of a hornblende grain with etching between 8pm — 12pm. Image captured with a Pixera digital imaging system. 42 20pm Figure 19. Class 4. Petrographic micrograph of a hornblende grain with etching between 12pm - 16pm. Image captured with a Pixera digital imaging system. 20pm Figure 20. Class 5. Petrographic micrograph of a hornblende grain with etching between 16pm - 20pm. Image captured with a Pixera digital imaging system. 43 20pm Figure 21. Class 6. Petrographic micrograph of a hornblende grain with etching between 20pm — 24pm. Image captured with a Pixera digital imaging system. 20pm Figure 22. Class 7. Petrographic micrograph of a hornblende grain with etching between 24pm — 28pm. Image captured with a Pixera digital imaging system. 44 20pm ' Figure 23. Class 8. Petrographic micrograph of a hornblende grain with etching between 28pm - 32pm. Image captured with a Pixera digital imaging system. 45 20pm Figure 24. Petrographic micrographs of weathered hornblende grains with holes. Arrows point to holes. Images captured with a Pixera digital imaging system. 46 20pm Figure 25. Petrographic micrographs of weathered hornblende grains with fractures. Images captured with a Pixera digital imaging system. 47 20pm Figure 26. Petrographic micrographs of weathered hornblende grains with both holes and fractures. Images captured with a Pixera digital imaging system. 48 taken of the denticulated edge of the hornblende in order to determine amplitude and subsequent classification. However, when the amplitude of the denticulated edge was measured and the classification was determined, these grains were most likely placed in a class that underestimated the extent of their weathering. These grains were assigned, on the average, to classes 1-3 based on denticulation amplitude of the grain margin; yet, the amount of dissolution that had to occur in order to produce holes and fracture features is greater than the classification 1-3 would indicate. However, the decision was made to classify these grains based on the original research design. Due to the small number of grains encountered with either of these features, it is doubtful that the results are skewed by virtue of using this classification scheme. Perhaps future studies should take these dissolution features into consideration when designing the research in order to utilize or develop a classification scheme that would better accommodate grains with these unusual features. Quartz/Feldspar Ratio Determination The degree of mineral weathering was determined by quantifying the quartz/ feldspar ratios in each of the horizons. Approximately 25g of the cleaned, fine sand size fraction of the sample was used from each horizon to prepare thin sections. The samples were impregnated with epoxy and allowed to harden to produce a cube that was then sliced, mounted on microscope slides, and ground to a thickness of approximately 30pm. The thin sections were stained with Alizarin red to help distinguish plagioclase feldspar from quartz. However, both plagioclase and K-feldspar were counted as feldspar for the ratio determination. Point counting was performed on these thin sections to obtain 49 quartz/ feldspar ratios. Only quartz and feldspar grains were counted, other minerals and void spaces were disregarded. A total of 300 grains were counted per horizon. Again, the quantity of 300 was chosen for the optimum combination of accuracy and speed (Van der Plas and Tobi, 1965; Pettijohn et al., 1973). Iron and Aluminum Extractions The three extracts that were used in this research are: sodium citrate-dithionite (CD), acidified ammonium-oxalate (AAO), and sodium pyrophosphate (PP). Extractions were performed on the illuvial and eluvial horizons only; the C horizons were not analyzed. The methods, which were used to perform these extractions, were from Weisenborn (2001), which were modifications of methods by Ross and Wang (1993) and Loeppert and Inskeep (1996). The modifications were necessary to obtain a sufficient amount of supernatant to yield detectable and measurable amount of Fe and Al when analyzed by atomic absorption spectrophotometer (AAS). The supematants were analyzed for extractable Fe and Al on the flame apparatus of the Perkin-Elmer 5100 PC AAS in the Geochemistry Laboratory at Michigan State University. Weisenborn (2001) developed the method for this analysis by modifying the methods from the Soil Survey Laboratory Staff (1996) and Anonymous (1989). The results were reported in percentages of the analyte cation. Laboratory-Produced Etching Features Two points around which this research was designed involve chemical analysis of soils to determine Fe and A1 accumulation, and measurement of hornblende denticulation 50 amplitude. It was thought that CD could be employed in both of these analyses. CD could be one of the extractants used to determine Fe and Al accumulation, and it could be used to “clean up” the hornblende grains to prepare them for visual examination (Jackson et al., 1986). Jackson et al. (1986) describe CD as a good method of Fe removal because it is a strong reducing agent, and has a chelating agent to sequester F e2+ and F e3 + ions. The CD system uses sodium citrate as a chelating agent, sodium dithionite (N a23204) as a reducing agent, and sodium bicarbonate (NaHCO3, pH 7.3) as a buffer. During the design of this thesis, the question arose as to the degree to which CD, a strong reducing agent, would attack the hornblende, a ferromagnesian mineral. If CD would reduce and remove any crystalline Fe, it could duplicate or enhance the natural etch-features that were going to be measured visually. Using a petrographic microscope, it is possible to discern and measure features that are approximately 1pm in length. Due to this level of detail, and the importance to this study of accurate amplitude measurements, the question of laboratory-produced features on hornblende grains from soils could not be ignored. In order to answer this question fully it was decided that CD, along with the other laboratory pre—treatments, would be tested on hornblende grains to determine whether or not they would produce any artificial alteration features. Homblende from the Michigan State University mineralogy teaching collection was used as the control to test the different pre-treatments. The hornblende was first ground with a mortar and pestle then sieved to obtain a fine sand grain size fraction of 125 um - 250nm. The sample was then divided into thirds for treatment. The first third of the sample was not treated. The remainder of the sample was exposed to a clay dispersant ((NaP03)6 + NazCO3) in order to follow the standard sample cleaning process 51 (Soil Survey Division Staff, 1993). The sample was slowly shaken over night in the solution then rinsed with distilled water and dried in an oven at 85°C. The sample was then split in half for the remaining two treatments. One half was treated with sodium hypochlorite (NaOCl) to simulate the removal of organic matter. Bleach, with 6% available NaOCl, was mixed with distilled water to a concentration of 3% available NaOCl. The hornblende grains were placed in the bleach for approximately 8 hours, rinsed with distilled water, and then dried. The final portion of the sample was treated in sodium citrate-dithionite (Na3C6H507 + IOHZO + Na2S204). The hornblende grains were placed in the solution and shaken for 12 hours then placed in the refrigerator for 6 hours. The grains were then rinsed with distilled water and dried. A J EOL J SM-6400V scanning electron microscope (SEM) was used to view the hornblende grains to search for alteration features caused by the pre-treatrnents. The control hornblende grains, which were untreated, showed smooth freshly exposed surfaces and no etching features (Figures 27 and 28). The grains that were treated with sodium hypochlorite also had smooth surfaces (Figures 29 and 30). There were no signs of alteration or weathering features on these grains. The grains that were treated with sodium citrate-dithionite all contained etch pits and laboratory-produced alteration features (Figures 31 and 32), some of them very large (Figure 33). The edges of the grains were not sharp, but showed signs of alteration (Figures 34 and 35). The surfaces on these grains were not smooth, but showed dissolution features (Figures 36 and 37). The hornblende grains treated with CD were the only grains to show any sign of laboratory-produced alteration. 52 200nm Figure 27. Scanning electron micrograph of an untreated hornblende grain. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 28. Scanning electron micrograph of the indicated area on Figure 27 showing smooth surfaces with no etching features. 53 200nm Figure 29. Scanning electron micrograph of a hornblende grain treated with sodium hypochlorite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. . ,L‘ . ”flirting ’ 392'?“ .‘l 20pm Figure 30. Scanning electron micrograph of the indicated area on Figure 29 showing smooth surfaces with no etching features. 54 Area 1 Area 2 Figure 31. Scanning electron micrograph of a hornblende grain treated with sodium citrate-dithionite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. Figure 32. Scanning electron micrograph of the indicated ‘Area 1’ on Figure 31 showing laboratory-produced alteration features. 55 20pm Figure 33. Scanning electron micrograph of the indicated ‘Area 2’ on Figure 31 showing large laboratory-produced alteration feature. 56 ”I. {13' ii 0 I I, II“ All“ {MN 9m i it iiilll'l-"" i .jll 200nm Figure 34. Scanning electron micrograph of a hornblende grain treated with sodium citrate-dithionite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. ,_. /.v ’.' I: I”. ,, err-2y: Figure 35. Scanning electron micrograph of the indicated area on Figure 34 showing laboratory-produced alteration at the edges. 57 Figure 36. Scanning electron micrograph of a hornblende grain treated with sodium citrate—dithionite. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 5 mm Figure 37. Scanning electron micrograph of the indicated area on Figure 36 showing laboratory-produced alteration on surfaces. 58 The laboratory-produced alteration features raised two major concerns about the results of this study. The first concern was whether the laboratory-produced features are discemable at the petrographic microscope scale. This was a problem because the petrographic microscope would be used to classify the hornblende according to a scheme based on the denticulation amplitude. If the laboratory-produced features were mistaken for natural features and then measured as such, the grain would be classified incorrectly. Due to these laboratory-produced alteration features, I was also concerned was about the validity of the data from the atomic absorption spectrophotometry (AAS). These data could be skewed by mineral dissolution, which would release Fe and Al into the supernatant for measurement. The values obtained would not accurately reflect the “free” Fe and A1 but would also measure the Fe and Al from dissolved hornblende. It was clear from Figures 31 - 37 that hornblende dissolved in the presence of CD and released Fe and Al into solution. It had to be determined how much Fe and Al was released per a given volume of hornblende, and whether or not it was a large enough amount to skew the results from the AAS. In order to estimate how much Fe and Al was released, a representative hornblende formula was adopted. The hornblende that is found in Michigan soils was glacially transported from a source area in Canada (Dworkin et al., 1985). There are several studies that provide chemical analyses of hornblende from the Grenville Orogen of Ontario, Canada, which is part of the source area for the sediments in this study (Cureton et al., 1997; Cosca et al., 1991). Averaging the values provided, a representative hornblende formula was derived: K01 INaO.42C31.83 (Mgr .47P‘efiostCfiz.06'1‘i0.07M110.02/3\10.75)Vi (A11.69Si6.31)iv 022F0.02(0H)1.98 59 From this average formula, the molecular weight of hornblende was calculated to be 906.28 g/mol. With a specific gravity of 3.1 g/cm3 (Klein and Hurlbut, 1999), the molar volume of hornblende is 292.35 cm3/mol. Iron in this formula is 149.68 g/mol, or 16.5% of the total weight. It occupies a volume of 48.28 cm3/mol. Aluminum is 65.84 g/mol, or 7.3% of the total weight, and occupies 21.29 cm3/mol. The volume of hornblende that can be dissolved in the presence of CD, during the time allotted in the methodology, had to be determined. Stereo pairs were made using pictures from the SEM to estimate the depth of the laboratory-produced etch pits (Figures 38 — 41). The stereo pairs were made by taking two Polaroid pictures of the feature of interest. The feature was centered on the screen, and then the first picture was taken. The stub was tilted 3 degrees then re-centered, and another picture was taken. By placing the pictures side-by-side and using stereoscopes, the depth of etch pits was estimated. An average depth for all etch pits viewed was 0.5 pm — 1pm. Several assumptions were made before continuing with the calculations. First, it was assumed, for the sake of calculation, that the grains were all spherical. This makes the volume of the grain much easier to calculate. Also, due to the fact that the natural soil grains of this thesis study are from glacial deposits that have been reworked by meltwater, they tend to have a rounded, sub-spherical shape. The second assumption is that the grains are all 200nm in diameter. This number was chosen because it is an intermediate value for the fine sand size fraction (125 um — 250nm) used in the study. The final assumption is that a shell lum thick was dissolved from each grain. One micrometer is the largest value in the range of 0.5 pm — 1.0um, which was observed as the average depth for laboratory-produced features. By using 1pm, the amount of 60 l» l" I l-' I'l .‘V . ‘ ”a...“ m‘ . E01. 1 SHTWMI , ‘ , 1 r 1:; U I", KB ,. in 1:71 LI .4: mm Figure 38. Stereo pair 1. Polaroid pictures of scanning electron images were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 61 Figure 39. Stereo pair 2. Polaroid pictures of scanning electron images were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 62 I.” "j- [Ir I’ll Figure 40. Stereo pair 3. Polaroid pictures of scanning electron micrographs were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 63 Figure 41. Stereo pair 4. Polaroid pictures of scanning electron micrographs were used to estimate the depth of the laboratory-produced etch pits on the hornblende grains that were treated with sodium citrate-dithionite. The feature in the left picture is centered and straight; the right picture is tilted 3 degrees. Images were taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. material added to solution was over-estimated to give an upper limiting value. In order to estimate the amount of Fe and Al added to solution by the dissolution of hornblende, several calculations were made. The molecular weight and the assumptions stated above were used to calculate the total volume of a 200nm diameter sphere of hornblende to be 2.35x10'6cm3 with a mass of 7.30x10'6g. When a 1 pm shell was subtracted to simulate dissolution, the volume of the sphere dropped to 2.32x10‘6cm3 with a new mass of 7.19x10'6g. The volume loss was 3.6x10'80m3 and the mass loss was 1.12x10'7g. This was a 1.53% loss of both volume and mass. In order to determine if this volume and weight loss would affect AAS values from the soils, the percentage of hornblende in a soil sample also needed to be determined. This percentage would be the portion of the soil that would dissolve in CD, thus adding Fe and Al to solution. The percent of hornblende in the soil samples collected for this thesis study was determined by performing grain counts (Table 1). Using this percent value, calculations were made to determine the amount of Fe and Al that could be added to solution from hornblende for each horizon. These calculated values were then compared to the actual values of Fe and Al that were obtained by AAS (Table 1). The amount of Fe and Al that could be in solution due to CD dissolution of hornblende was negligible. According to calculations, only 3.6x10'8cm3 or 1.12x10'7g would be lost when a 1pm shell of hornblende dissolved from a 200nm sphere. This would release 1.8x10'8g of Fe and 8.0x10'9g of Al per 7.30x10'6g of hornblende. The amount of Fe and Al that was measured using AAS was significantly higher than the amount of Fe and Al that could be added by the dissolution of hornblende. The AAS measurements were one order of magnitude larger than the calculated values for 13 of the 65 Table l. Pedon Data for Laboratory-Produced Alteration Features Pedon Soil Sample Homblende Fe (%) in Fe (%) Al (%) in Al (%) i‘lUtiLuu weight (%) in fine dry sample due to CD dry sample due to CD (g) grain sample AA value dissolution AA value dissolution Kalkaska A 1.54 no data 0.067 no data 0.015 no data E 1.59 2.5 0.052 0.010 _ 0.013 0.004 Bhs 1.52 3.7 0.748 0.014 0.145 0.006 le 1.52 2.2 0.160 0.008 0.183 0.004 B82 1.53 1.5 0.054 0.006 . 0.059 0.003 Bw 1.50 no data 0.045 no data 0.045 no data Strong Rubicon A 1.56 no data 0.085 no data 0.027 no data E 1.52 2.2 0.080 0.008 r 10.017 0.004 ' le 1.50 6.5 0.696 0.025 0.264 0.011 ‘ Bs2 1.53 4.7 0.328 0.018 ,_ 0.179 70.008 Weak Rubicon A 1.60 no data 0.040 no data 0.017 no data E 1.50 3.3 0.036 0.012 0.008 0005 B51 1.59 3.3 0.286 0.013 0.176 0.006 B82 1.64 4.0 0.049 0.017 0.065 0.007 3 Grayling A 1.58 4.3 0.059 0.017 """ii‘jb‘fimwmdi'dtfimi Bs 1.52 5.3 0.108 0.020 0.111 ,. 0.009 .5 n 66 Blue: Al comparisons Yellow: Fe comparisons Orange: value below valid detection limit , comparisons. It was two orders of magnitude larger for 6 of the comparisons. In the remaining 6 comparisons, the AAS measurements were higher than the calculated values, but not significantly higher. Thus, I concluded that the Fe and Al from the sesquioxide coatings on the soil particles were the dominant contributor to the measured values from the AAS. The original concern that needed to be addressed was whether the laboratory- produced features are discemable under a petrographic microscope. This concern, in the context of this thesis study, became moot. The grains used for denticulation amplitude measurements were not cleaned of sesquioxide coating prior to examination. That was deemed unnecessary because the coatings were thin enough that they did not affect the measurements taken. However, for future studies, these laboratory-produced features are a source of potential error that should be addressed. It has been reported that CD does not extract Fe and Al from crystalline minerals (Mehra and Jackson, 1960; McKeague et al., 1971; Ross and Wang, 1993). Ross and Wang (1993) stated that this treatment removes the noncrystalline iron oxides, organically bound Fe and Al, as well as the finely divided hematite, goethite, lepidocrocite, and ferrihydrite. Mehra and Jackson (1960) tested CD against three other methods of iron oxide removal to determine its effectiveness. The other three methods were: (1) sodium sulfide-oxalic acid (NaZS- NH4Cl-oxalic acid), at a pH of 3.5-10 (Truog et al., 1937), (2) sodium dithionite in acid system (Na2SZO4-HZO), at a pH of 3.5 (Deb, 1950), and (3) zinc-ammonium oxalate (Zn- oxalic acid-NH4 oxalate), at a pH of 3.6 (Haldane, 1956). It was shown that these three methods attack iron-rich silicate minerals. In their conclusion, Mehra and Jackson (1960) said of CD extractant, “Above all, the treatment has almost no destructive effect on iron 67 silicate clay mineral.” My study, however, showed that CD did produce alteration features on hornblende grains that were discemable and measurable under a petrographic microscope. Therefore, it is not recommended to clean hornblende samples with CD when the objective is to measure denticulation amplitude or other etching features. 68 CHAPTER 7 Results Soil Classification The samples from the four sites were analyzed in the laboratory for various physical properties (Table 2) and chemical properties (Table 3). Using these data, the pedons were then classified based on Keys to Soil Taxonomy (Soil Survey Staff, 1998). Site 1, Kalkaska, was classified as a sandy, mixed, frigid Typic Haplorthod (Schaetzl, 2002) (Figures 42 and 43). This site was the most strongly expressed soil studied, and the only pedon with a Bhs horizon. Site 2, strong Rubicon, is a sandy, mixed, frigid Entic Haplorthod (Schaetzl, 2002) (Figures 44 and 45). Site 3, weak Rubicon, is a sandy, mixed, frigid Typic Haplorthod (Schaetzl, 2002) (Figures 46 and 47). Site 4, Grayling, was classified as a sandy, mixed, fiigid Entic Haplorthod (Schaetzl, 2002) (Figures 48 and 49). This was the least expressed soil studied, and it lacked an E horizon. Based on the laboratory tests, the Grayling soil classified as a Rubicon. The site, however, was mapped and field classified as a Grayling. So, in order to minimize confusion, I will continue to refer to site 4 as a Grayling soil. There was a trend in soil development that increased from southeast to northwest with the most strongly developed soil, the Kalkaska, being the farthest northwest. That site is in close proximity to Starvation Lake, where local residents report it is not uncommon to have 60+ cm of snow on the ground during winter. The snOWpack decreases slightly to the southeast where the lesser- developed soils are located (Schaetzl, 2002). 69 Table 2 . Physical Properties of Pedons. Horizon Depth Munsell Grain Size T C (cm) Color C f IVC le s [Ms lFs WFSIS [C (moist) per cent by weight Kalkaska POD Index of 1.5 Oi 0-3 A 3-16 7.5YR2/0 0.0 0.7 13.0 52.8 18.5 3.3 11.6 0.0 sand E 16-24 5YR4/3 0.0 0.5 15.5 54.1 14.8 2.6 12.4 0.0 sand Bhs 24-31 5YR2.5/2 0.0 1.7 12.1 50.2 17.1 2.2 13.2 3.5 c.sand le 31-58 7.5YR4/6 0.0 0.8 11.0 58.0 21.9 3.4 4.9 0.0 sand Bs2 58-85 5YR 5/8 4.0 1.9 16.0 62.5 16.2 0.8 2.5 0.0 sand Bw 85—128 10YR 4/6 4.0 0.8 7.3 59.3 29.3 1.5 1.8 0.0 sand 2E/Bt 128-160+ 10YR 4/6(E) 8.0 1.1 14.1 58.7 21.6 2.6 2.0 0.0 sand 7.5YR 4/4(Bt) Strong Rubicon POD Index of 0 Oi 0-5 A 5-9 7.5YR 3/2 2.0 1.4 14.3 52.4 19.0 3.2 9.7 0.0 sand E 9-18 7.5YR4/2 2.0 1.3 14.5 52.4 18.8 3.3 8.2 1.4 sand le 18-28 5YR4/6 6.0 1.8 16.0 50.2 18.0 4.5 8.8 0.7 sand BS2 28-48 7.5YR 4/6 6.0 2.3 11.7 51.6 22.2 5.7 6.5 0.0 sand BC 48-79 10YR 5/6 8.0 2.6 16.5 60.4 17.9 0.7 2.0 0.0 sand C 79—170 10YR 6/4 8.0 1.4 14.3 59.4 21.1 2.4 1.4 0.0 sand 2C 170+ 10YR 6/3 0.0 0.8 10.0 69.5 17.9 1.1 0.7 0.0 sand gay C f - Coarse fragments >2mm VC 3 - Very coarse sand 1.0mm - 2.0mm C s — Coarse sand 0.5mm - 1.0mm M s - Medium sand 0.25mm - 0.5mm F s - Fine sand 0.125mm - 0.25mm VF s - Very fine sand 0.053mm - 0.125mm S - Silt 2pm - 50pm C - Clay <2um T C - Texture Class c. sand - coarse sand 70 Table 2 (continued). Physical Properties of Pedons. Horizon Depth Munsell Grain Size T C (cm) Color C f IVC sIC 5 IM 5 IF 8 IVF sIS IC (moist) per cent by weight Weak Rubicon POD Index of 2 Oi 0-4 A 4-9 N 2/0 0.0 0.7 19.7 64.2 11.6 0.6 3.3 0.0 sand E 9-20 7.5YR 4/2 0.0 0.7 17.5 62.0 13.7 1.2 3.7 1.2 sand le 20-41 5YR 3/4 0.0 1.0 14.0 63.7 15.4 1.2 4.8 0.0 sand 882 41-62 10YR 4/6 0.0 0.2 10.3 71.5 16.1 0.7 0.1 1.0 sand BC 62-126 10YR 5/4 1.0 0.1 3.0 65.0 30.7 0.8 0.5 0.0 sand C 126-165+ 10YR 6/4 1.0 0.2 2.6 49.5 46.0 1.4 0.0 0.4 sand Grayling POD Index of 0 Oi 0-7 A 7-11 7.5YR 3/2 0.0 0.4 12.1 63.8 13.9 0.8 9.0 0.0 sand Bs 11-36 7.5YR 4/6 0.0 0.2 14.7 64.7 15.8 1.0 3.6 0.0 sand BC 36-62 10YR 5/6 0.0 1.0 15.3 67.0 15.7 0.4 0.7 0.0 sand C 62-116 10YR 5/4 2.0 0.2 14.7 72.4 11.9 0.4 0.4 0.0 sand 2C 116-140 10YR 5/3 6.0 1.8 16.4 59.0 20.4 1.3 1.1 0.0 sand 3C 140-165+ 10YR 5/3 0.0 0.3 11.5 62.8 23.5 1.3 0.6 0.0 sand Key C f - Coarse fragments >2mm VC 8 - Very coarse sand 1.0mm - 2.0mm C s - Coarse sand 0.5mm - 1.0mm M s - Medium sand 0.25mm - 0.5mm F s - Fine sand 0.125mm - 0.25mm VF s - Very fine sand 0.053mm - 0.125mm S - Silt 2pm - 50pm C - Clay <2um T C - Texture Class 71 Table 3. Chemical Properties of Pedons. Horizon Depth pH Fed Fep Fe0 Ald Alp (cm) per cent of dry soil Kalkaska Oi 0-3 A 3-16 4.7 0.07 0.02 0.01 0.01 0.01 0.01 E 16-24 4.2 0.05 0.01 0.01 0.01 0.01 0.01 Bhs 24-31 4.0 0.75 0.17 0.18 0.15 0.14 0.14 851 31-58 6.1 0.16 0.07 0.04 0.18 0.20 0.22 B52 58-85 6.6 0.05 0.03 0.02 0.06 0.07 0.10 Bw 85-128 6.7 0.05 0.01 0.01 0.04 0.06 0.08 2E/Bt 128-160+ 6.9 Best Expressed Rubicon Oi 0-5 A 5-9 5.4 0.09 0.03 0.02 0.03 0.03 0.03 E 9—1 8 6.1 0.08 0.02 0.03 0.02 0.03 0.03 le 18-28 5.7 0.70 0.26 0.17 0.26 0.30 0.30 B32 28-48 6.4 0.33 0.08 0.08 0.18 0.20 0.28 BC 48-79 7.1 C 79-170 7.2 2C 170+ 7.6 Least Ex ressed Rubicon Oi 0-4 A 4-9 4.0 0.04 0.01 0.01 0.02 0.02 0.02 E 9-20 6.1 0.04 0.01 0.01 0.01 0.01 0.01 le 20-41 6.1 0.29 0.06 0.06 0.18 0.18 0.24 B52 41-62 6.9 0.05 0.02 0.01 0.06 0.07 0.13 BC 62-126 7.2 C 126-165+ 7.5 Gra ' 0-7 7-1 1 1 1-36 36-62 62-116 116-140 140-165+ 72 Figure 42. The Kalkaska pedon is the most strongly expressed Spodosol in the development sequence of this study. Photo taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. 73 Site Name: Kalkaska Pit Location: Sand Lake Road Slope at pit (%): 0 Parent material: Outwash, coarse at depth Drainage class: Well drained Water table (cm): none Landform type: Outwash plain Elevation: 394m Erosion: none Evidence of bioturbation: none Vegetation at site: Sugar maple, Red pine, Yellow birch Map: Starvation Lake Quadrangle Horizon Horizon Moist Color Structure Consistence Depth (cm) (Munsell) grade Oi 0 - 3 A 3 — 16 7.5YR 2/0 weak granular moist v friable E 16 -— 24 5YR 4/3 weak granular moist v friable Bhs 24 — 31 5YR 2.5/2 medium subangular blocky moist fi'iable B51 31 - 58 7.5YR 4/6 medium subangular blocky moist friable 882 58 — 85 7.5YR 5/8 weak subangular blocky moist v friable Bw 85 — 128 10YR 4/6 weak subangular blocky moist v friable 2E/Bt 128 — 160+ 10YR 5/3 E weak subangular blocky moist v fiiable 7.5YR 4/4 Bt Figure 43. Selected information from the Soil Profile Description Sheet of the Kalkaska pedon. 74 Figure 44. The strong Rubicon pedon is the second best expressed Spodosol in the development sequence of this study. Picture taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. 75 Site Name: Strong Rubicon Pit Location: N. Blue Lake Road across from Hassley Road (private) Slope at pit (%): 0 Parent material: Outwash Drainage class: Excessively well drained Water table (cm): none Landform type: Outwash Plain Elevation: 384m Erosion: none Evidence of bioturbation: none Map: Starvation Lake Quadrangle Vegetation at site: White pine stumps, reindeer moss, White birch, Pin cherry, Aspen Horizon Horizon Moist Color Structure Consistence Depth (cm) (Munsell) grade Oi 0 - 5 A 5 — 9 7.5YR 3/2 weak granular moist v friable E 9 — 18 7.5YR 4/2 weak subangular blocky moist v friable le 18 — 28 5YR 4/6 weak subangular blocky moist v friable Bs2 28 - 48 7.5YR 4/6 weak subangular blocky moist v friable BC 48 — 79 10YR 5/6 8 grain subangular blocky moist friable C 79 — 170 10YR 6/4 3 grain subangular blocky moist v friable 2C 170+ 10YR 6/3 5 grain subangular blocky moist v friable Figure 45. Selected information from the Soil Profile Description Sheet of the strong Rubicon pedon. 76 Figure 46. The weak Rubicon pedon is the third best expressed Spodosol in the development sequence of this study. Photo taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. 77 Site Name: WeakrRubicon Pit Map: Fredrick Quadrangle Location: Deward Road on west side Slope at pit (%): 0 Landform type: Terrace Parent material: Outwash Elevation: 355m Drainage class: Excessively well drained Erosion: none Water table (cm): none Evidence of bioturbation: none Vegetation at site: Red we, Jack pine, White pine, White spruce, Red maple, Pin cherry Blueberry, reindeer moss Horizon Horizon Moist Color Structure Consistence Depth (cm) (Munsell) grade Oi 0 - 4 A 4 -— 9 N 2/0 weak granular moist v friable E 9 — 20 7.5YR 4/2 weak medium granular moist v friable B81 20 — 41 5YR 3/4 weak subangular blocky moist friable Bs2 41 — 62 10YR 4/6 8 grain subangular blocky moist v friable BC 62 — 126 10YR 5/4 5 grain subangular blocky moist v friable C 126 — 165+ 10YR 6/4 s grain subangular blocky moist v friable Figure 47. Selected information from the Soil Profile Description Sheet of the weak Rubicon pedon. 78 Figure 48. The Grayling pedon is the least strongly expressed Spodosol in the development sequence of this study. Photo taken by Dr. Randy Schaetzl, Professor, Dept. of Geography, Michigan State University. 79 Site Name: GraylingPit Location: Goose Creek Road across from Berkley Road Slope at pit (%): 0 Parent material: Outwash Drainage class: Excessively well drained Water table (cm): none Vegetation at site: Jack pine, blueberry, planted Red pine, reindeer moss Landform type: Terrace in Goose Creek Valley Elevation: 351m Erosion: none Evidence of bioturbation: none Map: Lake Margrethe Quadrangle Horizon Horizon Moist Color Structure Consistence Depth (cm) (Munsell) grade Oi 0-7 A 7 — 11 7.5YR 3/2 weak subangular blocky moist loose Bs 11 — 36 7.5YR 4/6 weak subangular blocky moist v fiiable BC 36 — 62 10YR 5/6 s grain subangular blocky moist loose C 62 — 116 10YR 5/4 3 grain subangular blocky moist loose 2C 116 — 140 10YR 5/3 s grain subangular blocky moist loose 3C 140 — 156+ 10YR 5/3 s grain subangular blocky moist loose Deep 3C 280 — 300+ 10YR 5/3 8 grain subangular blocky moist loose Figure 49. Selected information from the Soil Profile Description Sheet of the Grayling pedon. 80 Homblende Denticulation Amplitude Three hundred hornblende grains from each horizon were examined under a petrographic microscope to determine the denticulation amplitude along the margin of the grain perpendicular to the c-axis (Figure 50). A weighted average of denticulation amplitude and the standard deviation were then calculated for each horizon (Table 4, raw data in Appendix A). The weighted average for each horizon was calculated by multiplying the number of grains in each class of that horizon by the mean amplitude measurement for that class, adding these values together, then dividing the sum by 300, which is the number of grains measured in each horizon. These values were used to create a horizon vs. denticulation amplitude plot for each site (Figure 51). All four sites exhibited the same trend within each profile. The denticulation amplitude increased up profile; denticulation amplitudes were minimal in the C horizons, but were larger in the shallower soil horizons. The range of denticulation amplitude, which is the difference between the highest weighted average and the lowest weighted average in each profile, increased as the strength of soil development increased. The Grayling pedon had a range of 0.39pm - 3.24pm or 2.85pm, and the weak Rubicon was 0.57pm - 5.79pm or 5.22pm. The strong Rubicon had a range of 1.15 pm - 7.16pm or 6.01 pm, while the Kalkaska pedon was 1.31 um-8.21um or 6.90pm. The average denticulation amplitude measurement for each individual pedon was calculated by averaging the weighted horizon values (Table 4). The average denticulation amplitude for the Grayling was 1.53pm, the weak Rubicon was 2.56pm, the strong Rubicon was 3.09pm, and the Kalkaska was 3.70pm. All four plots exhibit a break in slope directly above the uppermost B horizon. In the Grayling the break was between the A and 851 horizons. 81 20pm Figure 50. Petrographic micrograph of the denticulated margins of a naturally weathered hornblende grain. Image captured with a Pixera imaging system. 82 Table 4. Denticulation Amplitude Results Soil & Weighted Standard Maximum Minimum Pedon Pedon Pedon Horizon Amplitude Deviation Amplitude Amplitude Weighted Highest Lowest Average Measure Measure Average Average Average Kalkaska 3.70 8.21 1.31 A 8.21 4.10 26 2 E 7.28 4.06 22 2 Bhs 3.39 2.47 14 2 le 2.79 1.92 14 2 882 2.67‘ 1.82 14 2 Bw 2.20 1.39 10 O 2E/Bt 1.74 2.00 14 0 Deep 2C 1.31 1.43 10 0 Strong Rubicon 3.09 7.16 1.15 A 7.16 4.53 30 2 E 5.84 4.00 26 2 B51 2.43 1.44 14 2 B32 1.94 0.87 6 0 BC 1.77 1.15 10 0 C 1.36 1.20 6 0 2C 1.15 1.61 10 0 Weak Rubicon 2.56 5.79 0.57 A 5.79 4.17 26 2 E 4.73 2.99 14 2 B5] 2.28 1.26 14 2 Bs2 2.04 0.63 6 0 BC 1.65 0.99 6 0 C 0.89 1.15 6 0 Deg) C 0.57 1.07 6 0 Gr yling 1.53 3.24 0.39 A 3.24 2.20 14 2 Bs 2.23 0.97 6 0 BC 1.75 1.15 6 0 C 1.42 1.38 10 0 2C 1.14 1.57 10 0 3C 0.53 1.12 6 0 Deep 3C 0.39 0.93 6 0 83 I IL._-o fl Strong Rubicon Horizons 012345678910111213 012345678910111213 Mean Denticulation Amplitude (urn) Mean Denticulation Amplitude (um) Weak Rublcon Horizons Horizons Y T 3456789101112” 012345678910111213 0 1 2 Mean Denticulation Amplitude (um) Mean Denticulation Amplitude (um) Figure 51. Denticulation amplitude vs. horizon plot for each of the pedons in this study. These plots show the change in the mean denticulation amplitude measurement with change in horizon. Error bars represent 1 standard deviation. 84 In the other three plots the break was below the E and above the uppermost B horizon. Quartz/Feldspar Ratios Horizon vs. quartz/feldspar ratio plots were constructed for each site (Table 5 and Figure 52, raw data in Appendix B). The quartz/feldspar values for the deepest horizons for all sites plot at roughly the same point. The two more weakly developed soils, the Grayling and the weak Rubicon, had similar trends, with no systematic vertical change in the quartz/ feldspar ratios. This differs from the two more strongly developed soils, the strong Rubicon and the Kalkaska, which plotted similarly to one another exhibiting an irregular but unmistakable increase in the quartz/feldspar ratios up profile. Extractable Iron and Aluminum Three extractants were used in this study to determine where Fe and Al were enriched or depleted, and to determine the form of the Fe and Al in each horizon. Fed data, which indicate the presence of organically complexed Fe as well as “free” Fe, was at maximum in the horizon immediately below the B horizon in the Kalkaska pedon and both Rubicon pedons (Table 3). The Grayling pedon does not exhibit an E horizon, but the Fed is at maximum abundance in the Bs horizon, which is the uppermost B horizon. Ald data mimic the Fed trend in each profile, except in the Kalkaska pedon. Ald is at maximum abundance in the B51 horizon, which is the second B horizon, not the uppermost B horizon. The PP extractant (F ep, Alp), which extracts Fe and A1 from organic complexes, shows a maximum abundance of both Fe and Al in the uppermost B horizon for all pedons (Table 3). 85 Table 5. Quartz / Feldspar Ratios Soil & Quartz / Feldspar Horizon Ratios Kalkaska A 15.67 E 16.65 Bhs 8.38 881 11.00 B52 5.98 Bw 9.34 2E/Bt 7.33 Deep 2C 6.32 Strong Rubicon A 24.00 E 11.50 851 16.65 852 10.54 BC 16.65 C 9.71 2C 6.14 Weak Rubicon A 5.67 E 7.82 851 6.32 852 4.45 BC 6.69 C 4.56 Deep C 7.11 Ms A 5.25 B5 6.50 BC 6.14 C 8.38 2C 10.54 3C 8.68 Deep 3C 7.57 86 Quartz/Felds par Ratios l “g 2 ‘1: 5 a 3 E a ,2 4 + Kalkaska “g, H -I— Strong Rubicon g 5 + Weak Rubicon 3 r + Grayling £2 6 O N - '8 7 I: 8 _ 0 5 10 15 20 25 30 Quartz/Felds par Ratio Figure 52. Horizon vs. Quartz/Feldspar Ratio. Numbers were used as horizon designations instead of traditional horizon nomenclature because each pedon has various horizon orders. This plot shows the change in the quartz/feldspar ratio with depth. 87 F 60 and A10 data, which are a measure of the “active” Fe and A1 of noncrystalline hydrous oxides, are at maximum in the uppermost B horizon in each pedon, except the Kalkaska pedon where A10 is at maximum in the B51 horizon (Table 3). Various calculations were made with the extraction data (Table 6). These calculations provide more information as to the nature of the Fe and Al in each horizon (Appendix C). The activity ratio is calculated by AAO / CD (F e0 / Fed) and indicates the amount of crystalline Fe oxides in the soil. A low ratio indicates a large amount of crystalline Fe oxides. The A horizon of the Kalkaska pedon had the lowest ratio, which was calculated to be 0.14. The other ratios were moderately high between 0.20 and 0.40, with the highest ratio being 0.40 from the 852 horizon of the Kalkaska pedon. The calculation that indicates whether the “free” oxides are in a crystalline form or noncrystalline form is CD — AAO (Fed — F eo and Ald -— A10) (Table 6). Low values from this calculation indicate the oxides are primarily in a noncrystalline form. The values for Fed — FeO are moderately low and range from 0.03 to 0.57. The values for Ald - Al0 are negative, except for one positive value of 0.01 in the Bhs horizon of the Kalkaska pedon. The molar ratio of Fep / Alp signifies the difference between Fe and Al in the organically complexed, noncrystalline form (Table 6). Ratio values less than 1.0 signify that there is more organically complexed, noncrystalline Al than Fe in the horizon. The Kalkaska pedon has values greater than 1.0 in the A, E, and Bhs horizons, while the BSI , BS2 and Bw horizons have values less than 1.0. The strong Rubicon has a value equal to 1.0 in the A horizon, while the E, B51, and 852 horizons are less than 1.0. However, the weak Rubicon has a value equal to 1.0 in the B horizon, while the A, B51, and 852 88 Table 6. Fe and Al Extraction Calculations Horizon FeO/Fed Fed-Fe0 Ald-Alo Few/Alp (Fe(,-Fep)/Fep (Al(,-Al,,)/Alp percent of dry soil Kalkaska A 0.14 0.06 0 2.00 -0.50 0 E 0.20 0.04 0 1.00 0 0 Bhs 0.24 0.57 0.01 1.20 0.06 0 B5] 0.25 0.12 -0.04 0.35 -0.43 0.10 852 0.40 0.03 -0.04 0.43 -0.33 0.43 Bw 0.20 0.04 -0.04 0.17 0 0.33 Strong Rubicon A 0.22 0.07 0 1.00 -0.33 0 E 0.38 0.05 -0.01 0.67 0.50 0 B51 0.24 0.53 —0.04 0.86 -0.35 0 Bs2 0.24 0.25 -0.01 0.04 0 0.40 Weak Rubicon A 0.25 0.03 0 0.50 0 0 E 0.25 0.03 0 1.00 0 0 851 0.21 0.23 -0.06 0.33 0 0.33 852 0.20 0.04 -0.07 0.29 -0.50 0.86 Grayling A 0.33 0.06 -0.01 1.67 -0.60 0 Bs 0.27 0.07 -0.06 0.54 -0.57 0.31 89 horizons are less than 1.0. The Grayling pedon has a value greater than 1.0 in the A horizon and a value less than 1.0 in the B5 horizon. The calculations of (FeO — Fep) / Fep and (A10 — Alp) / Alp indicate whether the Fe and A1 are organically bound or inorganically bound in a noncrystalline form (Table 6). Higher values suggest the inorganically bound complexes contribute more to this measurement, while lower values indicate the organically bound complexes are more dominant. In the Kalkaska pedon, the Fe values are low and range from —0.5 to 0.06. While the Al values in the Kalkaska are zero in the A, E, and Bhs horizons, in the remaining three horizons they range from 0.1 to 0.43. The strong Rubicon has Fe values that are either zero or negative, except in the E horizon where the value is 0.5. The Al values for this pedon are zero in the A, E, and B51 horizons, while the B52 is 0.4. The weak Rubicon has Fe values of zero, except in the B52 horizon where the value is —0.5. The Al values for this horizon are zero except for the B51 and B52 horizons where the values are 0.33 and 0.86 respectively. The Grayling pedon has Fe values that are negative and Al values that are zero and 0.31. 90 CHAPTER 8 Discussion Homblende Denticulation Amplitude The weighted average, maximum, and minimum denticulation amplitude of hornblende were determined for each horizon within each pedon (Table 4). Homblende grains were more weathered, as determined by the denticulation amplitude measurements, at the top of the soil profile than at the bottom of the profile in all four sites. The average was smallest in the deepest C horizon sampled and largest in the A horizon for every pedon. The Grayling C horizon average was 0.39pm compared to its A horizon of 3.24pm. The weak Rubicon averages were 0.57pm in the C horizon and 5.79pm in the A horizon. The strong Rubicon averages were 1.15pm and 6.01pm in C and A horizons respectively, while comparable Kalkaska data were 1.31pm in the C horizon compared to 8.21pm in the A horizon. Standard deviations were calculated on the weighted averages for each horizon within each pedon (Table 4). The standard deviation, which indicates the spread in the data, was largest in the A horizon and became progressively smaller with depth in the profile for each pedon. This trend is to be expected since the more extensively weathered hornblende is in the upper horizons. While the homblende denticulation amplitude showed an obvious trend within each profile, the extent of weathering varied between the profiles. This variability directly corresponded with the strength of soil development. The more strongly developed soils had the more extensively weathered hornblende. This was determined by 91 comparing the minimum, maximum, and average value of denticulation amplitude for each profile, which was calculated from the weighted averages for each horizon (Table 4). The minimtun and maximum values of the pedons followed a progression that increased with increasing strength of soil deve10pment. The lowest weighted amplitude average for the Grayling, weak Rubicon, strong Rubicon, and Kalkaska were 0.39pm, 0.57pm, 1.15pm, and 1.31pm respectively. The highest weighted averages were 3.24pm, 5.79pm, 7.16pm, and 8.21pm respectively. The profile average denticulation amplitude also corresponded with strength of soil development (Table 4). The average denticulation amplitude for the Grayling was 1.53um, the weak Rubicon was 2.56pm, the strong Rubicon was 3.09pm, and the Kalkaska was 3.70pm. The ranges of denticulation amplitude, as well as the average amplitudes, all directly correspond with the strength of soil development. Horizon vs. Homblende Denticulation Amplitude A plot of horizon vs. weighted denticulation amplitude average was constructed for the soil profiles (Fig. 51). All four plots exhibit a break in slope directly above the uppermost B horizon. In the Grayling the break was between the A and B51 horizons. In the other three plots, the break was below the E and above the uppermost B horizon. A possible explanation for this break in slope is the presence of two acid systems at work in the soil. Ugolini et a1. (1977) suggested that soluble organic acids control the pH and the anionic balance in the upper portion of the soil, whereas the lower portion is dominated by a bicarbonate system. They found mobile fulvic acids were reduced by 60 to 70 percent within the B horizons. The uppermost B horizon delineated the lower boundary 92 of the biopedological compartment that starts at the top of the canopy. The lower compartment extended downward from the uppermost B horizon, and was in contact with groundwater. The pH increased at the boundary between these two systems, which could account for the varying degree of etching. In the Grayling soil there was a distinct increase in pH from 4.4 to 6.1 between the A and Bs horizons (Table 3), which was the same point at which the horizon vs. denticulation amplitude plot had a distinct break in slope (Figure 51). The pH increased from 4.0 to 6.1 between the A and E horizons in the weak Rubicon soil (Table 3). However, this was not where the horizon vs. denticulation amplitude plot had a break in slope (Figure 51). The break for this plot was between the E and 351 horizons, where the pH held constant at 6.1. The strong Rubicon actually had a decrease in pH from 6.1 to 5.7 (Table 3) between the E and B51 horizons (Figure 51) where the plot had a break in slope, and then the pH did increase to 6.4 in the B52 horizon. The Kalkaska horizon vs. denticulation amplitude plot exhibited a similar trend to that of the strong Rubicon. In the Kalkaska the pH decreased from 4.2 to 4.0 (Table 3) between the E and Bhs horizons where there was a break in slope (Figure 51). However, the pH did increase to 6.1 in the B51 horizon. There was an increase in pH down profile in all four pedons. Only in the Grayling pedon did the increase in pH correspond with the break in slope of the plot. This pH increase did not, however, directly correspond with the break in slope for the other three pedons; the pH increase occurred one horizon below the break in slope. The data did not show strong, convincing evidence that the boundary between the two 93 systems of acid was the cause for the break in slope on each of the horizon vs. denticulation amplitude plots, however it may be influential. Influence of pH on Dissolution Rates It was observed that the pH in the upper section of the pedon, above the distinct change in pH, was between 4 and 5, while the pH in the lowest sample taken was between 6 and 7.5 (Table 3). The difference in pH between the upper and lower intervals was roughly 2 units in all four sites. The hornblende in the upper profile, with the lower pH, was 6 to 10 times more deeply etched than the hornblende in the lower profile, with the higher pH. All of the grains have been at their present location for the same amount of time, so the hornblende in the upper section etched 6 to 10 times faster than the hornblende in the lower section. These empirical data were compared with experimental data from literature. Sverdrup (1990, as cited in Brantley and Chen, 1995) determined a dissolution rate for hornblende, having the composition of Ca1_7Mg3,5Fel_3A11,4Si6,9Ozz(OH)2, at various values of pH. At a pH of 4.4, a dissolution rate of 1.58x10'l4 moles inosilicate/cmz/sec was measured. At a pH of 6.5, the dissolution rate was 2.00x10’IS moles inosilicate/cmz/sec. Given these two rates, dissolution at a pH of 4.4 is 7.9 times faster than dissolution at a pH of 6.5. This value lies within the range observed in the empirical data where dissolution rates were 6 to 10 times faster at a pH of 4 to 5 than at a pH of 6 to 7.5. This consistency supports the idea that the variation in hornblende etching in the upper profile is more strongly influenced by variation in pH than in age. 94 Factors that Influence pH It has been concluded by many researchers that pH does influence the dissolution rate of primary minerals (Brantley and Chen, 1995, and all references therein). However, there are many factors that influence the pH of a soil, and thus ultimately influence the dissolution rate of primary minerals. The range in the dissolution rate of hornblende observed in the empirical data could be explained by natural conditions that vary among the pedons, such as organic matter and the presence of clay. Organic matter influences the soil environment in a number of ways. The accumulation of trace metals in the soils has been attributed to the scavenging of organic humic and fulvic acids based on the observed association between organic-rich soils and high metal values (Hoffman and Fletcher, 1981). It is accepted that organic matter strongly influences the mobility of Fe and Al by complexation and thus providing a means of transport for the Fe and Al ions (e. g., Antweiler and Drever, 1983; Mokma and Evans, 2000; Chadwick and Graham, 2000; Ollier and Pain, 1996; Birkeland, 1999). Several researchers have studied the effects of organic matter on pH (e. g., Antweiler and Drever, 1983; Mast and Drever, 1986; Birkeland, 1999). Antweiler and Drever (1983) found that the presence of organic matter could lower the pH of the pore water, and thus influence the dissolution rates of minerals. In their study they noted that the pH of percolating water ranged from 4.3 to 6.5. Laboratory experiments were conducted on samples from their study area, in the absence of organic compounds, and yielded a pH range of 7.3 to 9.5. In the laboratory samples the Fe and Al concentrations were below detection limits because Fe and A1 are not soluble in that range of pH values. 95 They also noted that organic matter breaks down to form several different acids that react differently in the soil environment. Acids that were found in their study include: humic/fulvic, oxalate, acetate, and formate acids. Propionate, succinate, and malonate were also tentatively identified. Therefore, they concluded that the breakdown of primary minerals is strongly influenced by the presence of organic matter. Kalkaska Heavy Mineral Grains and Clay Clay may also have an influence on the dissolution rate of primary minerals. Not only does clay inhibit the pH from lowering in soils (Bloom, 2000; Churchman, 2000), pedogenic clay may attach to grains, fill pore space, and alter water flow patterns (Chadwick and Graham, 2000; Goldberg et al., 2000; Radcliffe and Rasmussen, 2000). Understanding that clay may have an influence on the dissolution rate of primary minerals, heavy mineral grains from the Kalkaska A, E, Bhs, and B51 horizons were viewed in a scanning electron microscope (SEM) in order to View any clay that might be present. The Kalkaska pedon was chosen because of the very distinct pH change from 4.0 to 6.1 that occurs between the Bhs and B51 horizons (Table 3). Roughly 15 grains from each horizon were rinsed to remove dust, and heavy liquid separated to make the selection process easier. In order to observe the clay coatings, or coverings, the clay and sesquioxides were not removed. Energy-dispersive spectroscopy (EDS) was used to identify the weathering product. The grains above the change in pH were markedly different from the grains below the pH change. These differences included: organic material, weathering product chemistry, and weathering product abundance. In horizon A, with a pH of 4.7, there is 96 very little weathering product present (Figure 53), however, the weathering product that is present was identified as an aluminosilicate, most likely kaolin (A12SizOs(OH)4) (Figure 54). The determination was based on elements identified by EDS. Organics are also present on these grains as determined by EDS analyses that show a high C peak (Figures 55 and 56). Tube-like features (Figures 57 and 58) were also scanned with EDS (Figure 59) and determined to be organic material due to the high C content. In the E horizon, with a pH of 4.2, there is more clay present than in the A horizon, but less organic matter (Figure 60). The weathering product in this horizon was identified as an aluminosilicate (Figure 61). The grains from Bhs horizon, with a pH of 4.0, have notably more clay than the B horizon, but the grains are not completely coated (Figure 62). This weathering product was also an altuninosilicate (Figure 63). There is also more organic matter in this horizon than in the E horizon (Figure 62). The le horizon has a pH of 6.1 and is the uppermost horizon below the distinct change in pH. In this horizon, the grains are covered extensively with clay (Figure 64), which was determined to be an aluminosilicate (Figure 65). String-like features were observed in this horizon (Figures 66 and 67), however EDS did not show them to contain carbon (Figure 68). They all had the same EDS signature as the weathering product, which was identified as an aluminosilicate. The identity of these features is unknown. Several observations were made among the horizons. Tungsten was detected with the aluminosilicate in the Bhs and B51 horizons (Figures 63, 65, 68). It most likely sorbed to the clay from the heavy liquid, which was sodium polytungstate, during the laboratory procedure to separate the heavy minerals from the sample. Also, upon examining the EDS data, it was noticed that Al increased in abundance with depth 97 200nm Figure 53. Scanning electron micrograph of a weathered heavy mineral gain from the A horizon of the Kalkaska pedon with very little weathering product. This product is an aluminosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 4000 ”flacan 15.360 Figure 54. Energy-dispersive spectrogaph of the weathering product indicated the by circle in Figure 53. The high Si content with Al and 0 led to the determination that it is an aluminosilicate. The Au is the signature of the coating used to prepare the gain. 98 100nm Figure 55. Scanning electron microgaph of a weathered heavy mineral gain from the A horizon of the Kalkaska pedon with organic material present. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 5000 INJCOO O "‘ 9“ nu Ca Ti 90 n F- I K , a“ A J Figure 56. Energy-dispersive spectrogaph of the organic material indicated by the circle in Figure 55 showing high C content. The Au is the signature of the coating used to prepare the gain. 99 200nm Figure 57. Scanning electron microgaph of a weathered heavy mineral gain from the A horizon of the Kalkaska pedon with tube-like organic material present. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 50pm Figure 58. Scanning electron microgaph of tube-like organic material indicated by the circle in Figure 57. 100 3000 ‘ an amazon m m , N "a u Ca 3 1J Bu Fe Qu I “' ‘“ “ W ., [L ““ 0 l A Au I.) luv .000 14.400 Figure 59. Energy-dispersive spectrogaph of tube-like organic material indicated by the circle in Figure 57 showing high C content. The Au is the signature of the coating used to prepare the gain. 101 200nm Figure 60. Scanning electron microgaph of a weathered heavy mineral grain from the B horizon of the Kalkaska pedon with more weathering product than the A horizon. This product is an aluminosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 7000 Inacon le 16.875 Figure 61. Energy-dispersive spectrogaph of the weathering product indicated the by circle in Figure 60. The high Si content with A1 and 0 led to the determination that it is an aluminosilicate. The Au is the signature of the coating used to prepare the gain. 102 Organic A material 200nm Figure 62. Scanning electron microgaph of a weathered heavy mineral grain from the Bhs horizon of the Kalkaska pedon with more weathering product and organic material than the B horizon. This weathering product is an aluminosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. 4000I I it unJCOn 4.. I0 . 000 Rev 1:. .520 Figure 63. Energy-dispersive spectrogaph of the weathering product indicated the by circle in Figure 62. The high Si content with A1 and 0 led to the determination that it is an aluminosilicate. The Au is the signature of the coating used to prepare the gain. 103 200m Figure 64. Scanning electron microgaph of a weathered heavy mineral gain from the 351 horizon of the Kalkaska pedon with an extensive coating of weathering product. This product is an alurrrinosilicate. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. I n 3 = o n ktV 15.360 Figure 65. Energy-dispersive spectrogaph of the weathering product indicated by the circle in Figure 64. The high Si content with A1 and 0 led to the determination that it is an aluminosilicate. The Au is the sigrature of the coating used to prepare the gain. 104 200nm Figure 66. Scanning electron microgaph of a weathered heavy mineral grain from the BSI horizon of the Kalkaska pedon with string-like features that were determined to not be organic due to the EDS having the same signature as the weathering product. Image taken at the Center for Advanced Microscopy at Michigan State University by Leslie Mikesell and Ewa Danielewicz. ilill|j :‘ (.11 Jill" In I ' iimj U“ ‘;H'r an Figure 67. Scanning electron microgaph of string-like features indicated by the circle in Figure 66. 105 3000 Si Or'DCOf‘! RI I? o I"°°° 15.360 Figure 68. Energy-dispersive spectrograph of the string-like features indicated the by circle in Figure 66. The features were determined to not be organic due to the EDS having the same signature as the weathering product. The Au is the signature of the coating used to prepare the grain. 106 (Figures 54, 61, 63, 65). This is to be expected since Al is more soluble at low pH values, allowing it to move more deeply in the profile (Mokma and Evans, 2000). The mobility of the A1 allowed aluminosilicates to form lower in the pedon. A marked increase of pedogenic clay on gains, which was observed between the E and Bhs horizons, could indicate altered water flow paths that would affect the dissolution rate of primary mineral (e. g., Dove, 1995; Nugent etal., 1998). The gain size distribution data support this qualitatively observed increase in clay in the Bhs horizon (Table 2). The Bhs horizon of the Kalkaska pedon is 3.5 percent clay, while the other horizons have 0.0 percent clay. It is at this same point, between the E and Bhs horizons of the Kalkaska pedon, that the break in slope is observed in the horizon vs. denticulation amplitude plot (Fig. 51); meaning the dissolution rate above the Bhs horizon is faster than the rate in the Bhs horizon and below. Determining the exact effect clay has on the dissolution rate of primary minerals is beyond the scope of this study, however, the presence of clay does seem to influence the dissolution rate of primary minerals in soils. Quartz/Feldspar Ratios as a Weathering Indicator The quartz/ feldspar ratio was chosen to determine relative weathering. For this study, a horizon vs. quartz/feldspar ratio plot was constructed for each of the four pedons (Figure 52). A base, or initial, value was determined by calculating a quartz/feldspar ratio for the deepest sample taken. The deepest horizons for all sites plot at roughly the same point. This indicated that the parent material for the four sites is fairly uniform with respect to the quartz and feldspar content. The divergence in the plots from that point was assumedly due to soil development. The two less developed soils, the Grayling and 107 the weak Rubicon, plotted similarly, and had near vertical trends; the quartz/feldspar ratios were relatively uniform up-profile. Upper horizons in these two pedons were not depleted in feldspar with respect to quartz, indicating they were not extensively weathered. The two more strongly developed soils, the strong Rubicon and the Kalkaska, also plotted similarly to one another. They both had higher ratios up-profile than the first two plots, and they both had many breaks in slope producing a zigzag pattern. The larger ratio value was caused by a depletion of feldspar with respect to quartz, indicating that these two pedons were more extensively weathered than the Grayling and weak Rubicon. Grain Size and Quartz/Feldspar Ratios The multiple breaks in slope, or zigzag pattern, on the horizon vs. quartz/feldspar ratio plot may be due to the gain size used for point counting. Point counts were performed on thin sections that contained only the fine sand size fraction (125 um- 250pm) of the soil. During the design of this research, it was decided to use the fine sand size fraction throughout the study because several studies show that heavy minerals, which include hornblende, are concentrated here (e.g., Locke, 1979; Hall and Horn, 1993). Quartz and feldspar are present in this size fraction, however, they are light minerals and their abundance may not be accurately represented in this size fraction alone. The point counting did produce data with an obvious overall trend, but the zigzag pattern may be due to choosing a narrow range and small gain size for point counting. Feldspar grains may break apart along cleavage planes when weathering, producing more and smaller gains than the original. These smaller gains would be concentrated in this size fraction and would not be an accurate depiction of the mineral’s abundance in the 108 whole soil. Quartz, on the other hand, is more persistent than feldspar and may not break up into smaller gains during weathering. This could lead to skewed ratios. The trend of the plot is an accurate depiction for the overall trend of the quartz/feldspar ratios in each pedon. The zigzag pattern, however, may be the result of using only the fine sand size fraction of the soil when point counting. The A horizon of the Kalkaska pedon does not have the highest quartz/feldspar value, which is different than expected. The zigzag pattern of the slope has the Kalkaska and strong Rubicon plots crisscross each other with the Rubicon having the highest value in the A horizon (Figure 52). In examining the data, it was observed that the strong Rubicon has two anomalies in the physical data (Table 2). The pH values for this pedon are more variable than the pH values of the other three pedons. The Kalkaska, weak Rubicon, and the Grayling pedons all have pH values that exhibit a clear trend; that being a trend of decreasing pH values from the A to the E or an upper B horizon, and then increasing with depth. The pH values for the strong Rubicon do not exhibit a clear trend. This pedon also is different from the other three in its gain size distribution. It is the only pedon to have coarse fragnents in the upper horizons. The other pedons have coarse fragments in their lower horizons; however, the strong Rubicon has the highest percent of coarse fragnents in those horizons. These two anomalies may indicate the composition of the parent material for the strong Rubicon is slightly different than the parent material of the other three soils. A different composition in the parent material would yield different quartz/feldspar ratios. One observation that makes this explanation questionable is that the deepest samples from all four pedons plot at roughly the same point. However, the anomalous gain size distribution alone could yield these 109 differences. The strong Rubicon has a higher abundance of the coarse fragments than the other pedons; however, the point counts were performed on the fine sand size fraction of the soil. By using a narrow range of gain sizes, this test could give erroneous data for this pedon. Due to the close proximity of the pedons to one another in the field, and the clear and strong trends in the analytical data for both the hornblende denticulation amplitude and the Fe and Al mobility, it was concluded that the quartz/feldspar ratio trends are due to the difference in gain size and not a difference in composition of the parent material. Changes in the quartz/feldspar ratios indicate bulk mineralogy change. Rai and Kittrick (1989) suggested soils are short-term entities relative to the time required for the dissolution of minerals to the extent that changes in bulk chemistry could be observed. The age of these pedons, approximately 13,000 yr B.P., may simply be too young for this index to yield conclusive trends and results. It is not known which contributes more to the ambiguity in these results, the age of the soil or the grain size fraction used in point counfing. While these plots do not contradict the order of soil development as assessed in the field, they do not support it either. This mineral weathering indicator simply distinguishes between the two least developed pedons and the two more strongly developed pedons. A clear ordering of the four pedons cannot be demonstrated using quartz/ feldspar ratios. Summary Homblende denticulation amplitude measurements and soil development both 110 indicate the same trend. They suggest the order of 5011 development among the pedons, in increasing strength, is the Grayling pedon, the weak Rubicon pedon, the strong Rubicon pedon, and the Kalkaska pedon. They both show the same clear, strong trends within each pedon and the same convincing trends among pedons. The quartz/feldspar ratios results did not exhibit strong trends due to ambiguity in the data, which was most likely caused by the young age of the soil or the gain size fraction used for point counting. These ratios did, however, indicate that weatherable silicates were less depleted from the less developed Grayling and weak Rubicon pedons, and more depleted from the more developed strong Rubicon and Kalkaska pedons. 111 CHAPTER 9 Conclusions Homblende Denticulation Amplitude The hornblende denticulation amplitude measurements were plotted and used to conclude that mineral weathering increased up profile. The plots also showed clear trends that were used to determine a relative weathering order for the pedons in this study. Comparison of Homblende Denticulation Amplitude and Soil Development The hornblende denticulation amplitude measurements and the soil development trends were similar both within and among the four pedons. There was a definite trend of the mineral weathering and soil development being less in the lower profile and progessively increasing in intensity up profile. Also, the mineral weathering trend among the pedons mimicked the order of soil development. From these trends, it can be concluded that, as these soils developed, the minerals weathered, and thus weathering and soil development are related and not independent of one another in these soils. Comparison of Homblende Denticulation Amplitude and Quartz/Feldspar Ratios Quartz/feldspar ratios resulted in trends similar to, but not as strong as, those of the denticulation amplitude measurements. The ratios decreased up profile indicating feldspar was weathering more in the upper profile than in the lower profile. This follows 112 the denticulation amplitude trend, but due to the zigzag nature of the ratio plot, the trend is not strong or clear. Comparison of Soil Development and Quartz/Feldspar Ratios Attempting to use quartz/ feldspar ratios as indicators of soil development may be inappropriate for soils as young as the pedons in this study, which are approximately 13,000 yr B.P. The ratio of quartz/feldspar relies on the depletion of feldspar as compared to quartz, which is not a rapid process. The plots made in this study showed the two least developed soils to be similar, and the two strongest developed soils to be similar to each other but different than the first two. The results did not provide a convincing relative weathering order for the pedons. Weathering Versus Soil Development The terms weathering and soil development have occasionally been used interchangeably. This study shows that mineral weathering and soil development do proceed in tandem on the soils of the study area. However, the terms mineral weathering and soil development reflect two different processes. Bland and Rolls (1998) defined weathering as, “the alteration by chemical, mechanical, and biological processes of rock and minerals, at or near the Earth’s surface, in response to environmental conditions.” Birkeland (1999) defines soil development as, “a qualitative measure of the amount of pedologic change that has taken place in the parent material.” These two processes, although usually occurring simultaneously and/or in parallel in regolith, should be named and referred to correctly. 113 Homblende as an Age Indicator Homblende denticulation amplitude, as an age indicator, may not be appropriate to use on older soils. The hornblende in older soils may have reached a steady state denticulation morphology so there would be no significant difference in the denticulation amplitude (Hall and Michaud, 1988). Steady state conditions in older soils may render this tool useless for determining relative weathering. Denticulation amplitude measurements may be useless as an age indicator also due to steady state. Age determination of a soil, or relative ages of multiple soils, would be impossible by denticulation amplitude alone once steady state has been reached. Without significant differences in amplitudes from one soil to another, an age determination cannot be made. As previously stated, the quartz/ feldspar ratio may be an inappropriate soil development indicator for younger soils because of the time requirements needed for a region to show a depletion in feldspar. However, hornblende denticulation amplitude measurements and quartz/ feldspar ratios complement each other. Where the quartz/feldspar ratio is not a good development indicator for younger soils, the denticulation amplitude measurements yield convincing conclusions. The opposite is also true, where the denticulation amplitude measurements may become useless due to steady state in older soils, the quartz/ feldspar ratios may exhibit strong trends. Homblende etching has been studied by many researchers and has been deemed a suitable tool for relative age determination (Clayton, 1979; Hall and Michaud, 1988; Hall and Heiny, 1983; Hall and Martin, 1986; Locke, 1979, 1986). This study, however, demonstrates that soils formed in deposits of the same age can have significant differences in denticulation amplitudes from one pedon to another. The differences 114 between the end members, the Grayling and Kalkaska pedons, is great enough that a researcher might conclude they were of different ages if hornblende amplitude was used as a relative age indicator in this setting. Homblende weathers to produce etch features in a predictable manner that is easily recognizable. However, by using these features alone, it is impossible to distinguish intensity of weathering from duration of weathering. Because of this fact, concluding the etch features are due to duration of weathering is an assumption at best and simply wrong at worst. 115 APPENDICIES 116 APPENDIX A Denticulation Amplitude Measurements Raw Data Table Pedon and Class 0 Class 1 ClassZ Class 3 Class4 Class 5 Class 6 Class 7 Class 8 Tot Avg Std Horizon MAO MA2 MA6 MA 10 MA 14 MA 18 MA22 MA26 MA 30 Dev Kalkaska A 0 33 132 88 35 8 3 l 0 300 8.21 4.10 E 0 61 128 76 27 5 3 0 0 300 7.28 4.06 Bhs O 216 67 14 3 0 0 0 O 300 3.39 2.47 B51 0 256 34 9 1 0 0 0 0 300 2.73 1.92 852 0 259 33 7 1 0 O 0 0 300 2.67 1.82 Bw 24 251 23 2 O 0 0 0 0 300 2.20 1.39 2E/Bt 101 177 14 7 1 0 0 0 0 300 1.74 2.00 Deep2C 129 159 11 l 0 0 0 0 0 300 1.31 1.43 Pedon Avg 3.69 Strong Rubicon A 0 67 137 61 21 8 4 1 1 300 7.16 4.53 E 0 112 120 44 19 3 l 1 0 300 5.84 4.00 BS] 0 272 25 2 1 O 0 0 0 300 2.43 1.44 B52 25 267 8 0 0 O 0 0 0 300 1.94 0.87 BC 55 236 8 1 0 0 0 O 0 300 1.77 1.15 C 110 183 7 0 0 0 0 0 0 300 1.36 1.20 2C 159 128 10 3 0 0 0 0 O 300 1.15 1.61 Pedon Avg 3.09 Weak Rubicg A 0 117 117 44 14 4 3 l O 300 5.79 4.17 E 0 139 124 30 7 0 0 0 0 300 4.73 2.99 851 0 283 14 2 1 0 0 0 0 300 2.28 1.26 B52 6 288 6 0 0 0 0 0 0 300 2.04 0.63 BC 62 233 5 0 0 0 0 0 0 300 1.65 0.99 C 175 121 4 0 0 0 0 0 0 300 0.89 1.15 DeepC 223 73 4 O 0 O 0 0 0 300 0.57 1.07 Pedon Avg 2.56 (3423:1198 A 0 219 70 10 1 0 0 0 0 300 3.24 2.20 B5 2 280 18 0 0 0 0 0 O 300 2.23 0.97 BC 59 230 11 0 0 0 0 0 0 300 1.75 1.15 C 111 178 10 1 0 0 0 0 0 300 1.42 1.38 2C 163 121 15 1 0 0 0 0 O 300 1.14 1.57 3C 233 61 6 0 0 0 0 0 0 300 0.53 1.12 Deep 3C 248 49 3 0 0 0 0 0 0 300 0.39 0.93 Pedon Avg 1.53 Key: MA - Mean amplitude for that class Tot - Total gains measured 117 Avg - Weighted average Std Dev - Standard deviation APPENDIX B Quartz and Feldspar Raw Data Table Soil & Quartz Feldspar Total Ratios Horizon Counts Counts Counts Kalkaska A 282 18 300 1 5.67 E 283 1 7 300 16.65 Bhs 268 32 300 8.38 B81 275 25 300 11.00 B82 257 43 300 5.98 Bw 271 29 300 9. 34 2E/Bt 264 36 300 7.33 Deep 2C 259 41 300 6.32 Strong Rubicon A 288 12 300 24.00 E 276 24 300 1 1.50 B81 283 17 300 16.65 BS2 274 26 300 10. 54 BC 283 1 7 300 16.65 C 272 28 300 9.71 2C 258 42 300 6.14 Weak Rubicon A 255 45 300 5.67 E 266 34 300 7.82 BS] 259 41 300 6.32 B82 245 55 300 4.45 BC 261 39 300 6.69 C 246 54 300 4.56 DeeLC 263 37 300 7.1 1 Ms A 252 48 300 5.25 BS 260 40 300 6.50 BC 258 42 300 6.14 C 268 32 300 8.38 2C 274 26 300 10.54 3C 269 31 300 8.68 Deep 3C 265 35 300 7.57 118 APPENDIX C Extractable Iron and Aluminum Podzolization is the process by which Spodosols are developed. During this process, Fe and A1 are translocated from the O and A horizons to the B horizons. This process leaves an eluvial zone, or E horizon, which is depleted in Fe and Al, and produces an accumulation of Fe and Al in the B horizons. One measure of strength of soil development is the difference in the amount of Fe and A1 in the A and E horizons from that in the B horizons. To determine the amount of Fe and Al that has translocated, Fe and Al were chemically extracted from each of these horizons. In a well developed soil, there should be a trend that exhibits low amounts of Fe and Al in the A and E horizons and a higher amount of Fe and A1 in the B horizons, typically with the highest amount in the uppermost B horizon. Iron and aluminum are found in crystalline, complexes (organic and inorganic), and noncrystalline or amorphous forms in the soil. Three extractions were performed which were designed to test for Fe and Al in these various states. The first extractant was sodium citrate-dithionite (CD), which extracted the “free” Fe and A1. This value represents the total pedogenic crystalline, noncrystalline, and organically bound Fe and Al. The results are denoted as F ed and Ald. The second extractant was acidified ammonium-oxalate (AAO), which is good for indicating the presence of noncrystalline, or amorphous, Fe and Al. The results are denoted as Feo and A10. The third extractant was sodium pyrophosphate (PP), which is a good indicator for the presence of organically 119 bound Fe and Al. The results are denoted as F ep and Alp. These values yield a fairly accurate depiction of the movement of Fe and Al within the pedon. The F ed, which indicates the presence of “free” Fe in crystalline, noncrystalline, and organically bound complexes, has its maximum abundance in the horizon immediately below the B horizon in the Kalkaska pedon and both Rubicon pedons (Table 3). This maximum in the uppermost B horizon indicates a translocation of Fe down profile in each pedon. The Grayling pedon does not exhibit an B horizon, but the Fed is at maximum abundance in the B5 horizon showing Fe is being eluviated from the A horizon. The Ald mimics the F ed trend in each pedon except in the Kalkaska. The Ald value is highest in the B51 horizon, which is the second B horizon, not the uppermost B horizon. It is common, and almost expected, to have the maximum value for A1 occur lower in the pedon than the maximum value for Fe because of the solubility of A1 at lower pH values (Mokma and Evans, 2000). These trends show a downward movement of both “free” Fe and Al in all four pedons. The Feo, which indicates the “active” Fe in noncrystalline hydrous oxides, is at maximum abundance in the uppermost B horizon, and has a much lower abundance in the upper horizons of all four pedons (Table 3). The A10, which indicates amorphous aluminosilicates, trend is the same as the Ald trend in that it is highest in the uppermost B horizon in each pedon, except in the Kalkaska pedon where it is highest in the B51 horizon. There is a downward movement of “active” Fe and Al in all four pedons. The F ep, which indicates the organically bound, or complexed, Fe, is highest in the uppermost B horizon, and is lower in abundance in the upper horizons (Table 3). The 120 Alp trend is the same as the A10 and the Ald. The results from this extractant also suggest a downward movement of Fe and Al. In examining the data directly, as individual extractions, an accumulation of Fe and Al in the B horizons was observed (Table 3). In all four pedons, the lower values for Fe and A1 are in the A and E horizons. The higher values for Fe and A1 are in the uppermost B horizon, or in the second B horizon in the case of the Kalkaska pedon, and then the values decrease from that point with depth. Generally, Fed > Fep > Feo, which indicates that Fe exists as “free” oxides, either crystalline or noncrystalline, in all four pedons. Also, in all four pedons Alo > Alp > Ald and suggests that Al is primarily “active” noncrystalline. Various calculations may be made using these values that provide more information about the state of the Fe and Al in the horizon. The “activity ratio”, or F eo / Fed, for all four pedons yields moderately low values (Table 6). This indicates that Fe is predominantly crystalline in the pedons. The relative amount of crystalline “free” oxides compared to noncrystalline “free” oxides is calculated by CD — AAO (Fed — Fe0 or Ald — A10) (Table 6). Fed — Feo is positive for every horizon in each pedon indicating that crystalline Fe oxide is the primary form of “free” Fe. The Ald — Al0 value is either zero or negative. One exception to this is the Bhs horizon of the Kalkaska, which is still a low value at 0.01. This indicates that “free” Al is primarily noncrystalline in all four pedons. The molar value of F ep / Alp signifies whether Fe or A1 is more prevalent in the noncrystalline organic complexes (Table 6). Low ratio values indicate more of the complexes are Al than Fe. In the Kalkaska pedon, the A, E, and Bhs horizons have 121 Fep/Alp values of one or greater, signifying that Fe complexes are equal to or more prevalent than A] complexes in the upper portion of the pedon; while Pep/Alp in the B51, B52 and Bw horizons are less than one, indicating that Al complexes are more prevalent in the lower portion of the pedon. This trend could be explained by the solubility and subsequent mobility of A1 at these pH values. The A horizon of the strong Rubicon has a Fep/Alp value of one, indicating the number of Fe complexes is roughly equal to Al complexes, while the E, 851 , and 832 horizons have organic complexes mainly of Al. The weak Rubicon has Al complexes more prevalent than Fe complexes in all horizons except the B horizon, where they are roughly equal. The Grayling pedon has mainly Fe complexes in the A horizon and predominantly Al complexes in the B5 horizon. There is an overall trend among the four pedons of higher values in the upper pedon, indicating more Fe complexes than Al complexes, and decreasing values in the lower pedon, indicating Al complexes are more prevalent. The form of the noncrystalline complexes, whether organic or inorganic, can be estimated by (Fe0 — Fep) / F ep and (A10 — Alp) / A1p (Table 6). For this calculation, lower values indicate organic complexes are predominant, while higher values indicate inorganic complexes are more prevalent. The values for Fe are zero or negative, suggesting Fe is predominantly organically complexed. This is true for all horizons except in the Kalkaska Bhs horizon where it is 0.06, and in the strong Rubicon where it is 0.50. However, these values are considered low (Wang and Kodama, 1986), indicating Fe is organically complexed in these horizons as well. The values for A1 are zero or positive in all horizons. Even with the values being positive, they are still considered low ratio values, indicating Al is also primarily organically complexed. 122 To summarize, the extraction data indicates that Fe and Al in crystalline, noncrystalline, and complexed forms are all moving down the profile. In all four pedons, the Fe is predominantly in “free” crystalline form, meaning it is in Fe oxides. When it is complexed, it is organically complexed. The Al, in all four pedons, is predominantly “active” noncrystalline and organically complexed. Where “free” Al is present, it is also noncrystalline. It was also observed that the extraction data support the strength of soil development order of the pedons (Table 3 and 5). The Kalkaska pedon exhibits strong, clear trends with maximum values higher than those of any other pedon. This pedon is the only one in which the horizon of maximum Al abundance is below that of maximum Fe abundance, indicating a more strongly developed soil. The strong Rubicon pedon exhibits clear trends, and the maximum values are lower than those of the Kalkaska. The weak Rubicon pedon has trends that are not as clear as the first two pedons, and the maximum values are lower as well. The Grayling pedon does not exhibit an E horizon, indicating it is weakly developed, and so only the A and Bs horizons could be analyzed. These data do, however, show a downward movement of Fe and A1. The extraction data do correlate to the strength of soil development order. 123 REFERENCES 124 REFERENCES Allen, BL. and BF. Hajek. 1989. Mineral occurrence in soil environments. In: J .B. Dixon and SB. Weed (eds), Minerals in Soil Environments, 2nd ed. SSSA Book Series 12199-264. Anderson, H.A., Berrow, M.L., Farmer, V.C., Hepburn, A., Russell, J .D., and AD. Walker. 1982. A reassessment of podzol formation processes. J. Soil Sci. 33: 125- 136. Andrews, J.T. and G.H. Miller. 1980. Dating Quaternary deposits more than 10,000 years old. In: R.A. Cullingford, D.A. Davidson, and J. Lewin (eds), Timescales in Geomorphology. John Wiley & Sons Ltd., New York, New York: 263-287. Anonymous. 1989. Methods for determination of inorganic substances in water and fluvial sediments. In: F ishman, M.J., and LC. Friedman (eds.), Techniques of water-resources investigations of the United States Geological Survey. USDI Laboratory Analysis 5, Chapter A1, US Govt. Printing Office, Washington, DC. 595p. Anonymous. 1996. Glossary of soil science terms. Soil Sci. Soc. Am., Madison, WI. 134p. Antweiler, RC. and J .I. Drever. 1983. The weathering of a late Tertiary volcanic ash: Importance of organic solutes. Geochim. et Cosmochim. Acta 47:623-629. Barrett, LR. 1995. A stump prairie landscape in northern Michigan: Soils, forest vegetation, logging, and fire. Ph.D. dissertation. Michigan State University, East Lansing, Michigan. 206p. Barrett, LR. 1997. Podzolization under forest and stump prairie vegetation in northern Michigan. Geoderma. 78:37-58. Bateman, RM. and J.A. Catt. 1985. Modification of heavy mineral assemblages in English coversands by acid pedochemical weathering. Catena. 12:1-21. Bemer, RA. 1978. Rate control of mineral dissolution under earth surface conditions. Am. J. Sci. 278:1235—1252. Bemer, R.A., Sjoberg, E.L., Velbel, M.A., and MD. Korm. 1980. Dissolution of pyroxenes and arnphiboles during weathering. Science 207: 1205-1206. Bemer, RA. and J. Schott. 1982. Mechanism of pyroxene and amphibole weathering. II. Observations of soil gains. Am. J. Sci. 282:1214-1231. 125 Birkeland, P.W. 1974. Pedology, weathering, and geomorphological research. Oxford University Press, Inc., New York, New York, 285p. 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. 1999. Soils and geomorphology, 3rd ed. Oxford University Press, New York, New York, 430p. Blackburn, W.H. and W.H. Dennen. 1988. Principles of Mineralogy. Wm. C. Brown Publishers, Dubuque, Iowa. 413p. Bland, W. and D. Rolls. 1998. Weathering: An introduction to the scientific principles. Oxford University Press Inc., New York, New York, 271p. Blewett, W. 1990. The glacial geomorphology of the Port Huron complex in northwestern southern Michigan. Ph.D. dissertation, Department of Geogaphy, Michigan State University, East Lansing, MI. Blewett, W., Winters, HA, and R.L. Rieck. 1993. New age control on the Port Huron moraine in northern Michigan. Phys. Geog. 14:131-138. Bloom, RR. 2000. Soil pH and pH buffering. In: Summer, M.E., (ed.) Handbook of Soil Science. CRC Press, Boca Raton, Florida. pp. B-333-B-352. Blume, HP. and U. Schwertmann. 1969. Genetic evaluation of profile distribution of aluminum, iron, and manganese oxides. Soil Sci. Soc. Am. Proc. 33:438-444. Bockheim, J.G., Marshal, J.G., and HM. Kelsey. 1996. Soil-forming processes and rates on uplifted marine terraces in southwestern Oregon, USA. Geoderma. 73:39-62. Brantley, S.L., Blai, A.C., Cremeens, D.L., MacInnis, I., and R.G. Darrnody. 1993. Natural etching rates of feldspar and hornblende. Aquatic Sciences 55(4):262- 272. Brantley, S.L. and Y.Chen. 1995. Chemical weathering rates of pyroxenes and amphiboles. In: A.F. White and S.L. Brantley (eds), Chemical Weathering Rates of Silicate Minerals. Min. Soc. Am., Rev. Mineral. 31:119-172. Buurman, P. and LP. van Reeuwijk. 1984. Proto-imogolite and the process of podzol formation: A critical note. J. Soil Sci. 35:447-452. Chadwick, DA. and RC. Graham. 2000. Pedogenic processes. In: Summer, M.E., (ed.), Handbook of Soil Science. CRC Press, Boca Raton, Florida. pp. E-41-E-75. 126 Churchman, G]. 2000. The alteration and formation of soil minerals by weathering. In: Summer, M.E., (ed.), Handbook of Soil Science. CRC Press, Boca Raton, Florida. pp. F -3-F -76. Clayton, J .L. 1979. Nutrient supply to soil by rock weathering. p. 75-96. In: Impact of Intensive Harvesting on Forest Nutrient Cycling, State University of New York, College of Environmental Science and Forestry, Syracuse, New York. Cleaves, ET. 1974. Petrologic and chemical investigation of chemical weathering in mafic rocks, eastern piedmont of Maryland: Maryland Geological Survey Report of Investigations No. 25, 28p. Cosca, M.A., Essene, E.J., and JR. Bowman. 1991. Complete chemical analyses of metamorphic homblendes: Implications for normalizations, calculated HzO activities, and therrnobarometry. Cotnrib. Mineral. Petrol. 108:472-484. Cremeens, D.L., Darmody, R.G., and L.D. Norton. 1992. Etch-pit size and shape distribution on orthoclase and pyriboles in a loess catena. Geochim. et Cosmochim. 56:3423-3434. Cureton, J .S., Van der Pluijm, BA, and E.J. Essene. 1997. Nature of the Elzevir- Mazinaw domain boundary, Grenville Orogen, Ontario. Can. J. Earth Sci. 34:976- 991. Deb, B.C. 1950. The estimations of free iron oxide in soils and clays and their removal. J. Soil Sci. 1:212-220. DeConinck, F. 1980. Major mechanisms in formation of spodic horizons. Geoderma 24:101-128. Delvigne, J .E. 1998. Atlas of micromorphology of mineral alteration and weathering. Mineral. Ass. Can, Ottawa, Ontario. 494p. Dorronsoro, C. and P. Alonso. 1994. Chronosequence in Almar River fluvial-terrace soil. Soil Sci. Soc. Am. J. 58:910-925. Dove, RM. 1995. Kinetic and thermodynamic controls on silica reactivity in weathering environments. In: A.F. White, and S.L. Brantley (eds.), Chemical Weathering Rates of Silicate Minerals. Min. Soc. Am., Rev. Mineral. 31:236-290. Dreimanis, G.H., Reavely, G.H., Cook, R.J.B., Knox, KS, and F.J. Moretti. 1957. Heavy mineral studies in till of Ontario and adjacent areas. J. Sed. Pet. 27(2): 148-161. Dryden, L. and C. Dryden. 1946. Comparative rates of weathering of some heavy minerals. J. Sed. Pet. 16:91-96. 127 Dworkin, S.I., Larson, G.J., and G.W. Monaghan. 1985. Late Wisconsinan ice-flow reconstruction for the central Great Lakes region. Can. J. Earth Sci. 22:935-940. Eggleton, RA. 1986. The relation between crystal structure and silicate weathering rates. In: S.N. Colman, and DP. Dethier (eds.), Rates of Chemical Weathering of Rocks and Minerals. Academic Press, Inc., Orlando, Floridaz21-40. Everett, KR. 1971. Soils of the Meserve Glacier area, Wright Valley, South Victoria Land, Antarctica. Soil Sci. 112:425-438. Fanning, D.S. and M.C.B. Farming. 1989. Soil morphology, genesis, and classification. Wiley. New York, New York. 395p. Farmer, V.C., Russell, J .D., and ML. Berrow. 1980. Irnogolite and proto-imogolite allophane in spodic horizons: Evidence for a mobile aluminum silicate complex in podzol formation. J. Soil Sci. 31:673-684. Farmer, V.C. 1982. Significance of the presence of allophane and imogolite in podzol Bs horizons for podzolization mechanisms: A review. Soil Sci. Plant Nutr. 28:571- 578. Farmer, V.C., Russell, J .D., and B.F.L. Smith. 1983. Extraction of inorganic forms of translocated A1, Fe and Si from a podzol Bs horizon. J. Soil Sci. 34:571-576. Gardner, DR. and ER Whiteside. 1952. Zonal soils in the transition region between the Podzol and Gray-Brown Podzolic regions in Michigan. Soil Sci. Soc. Am. Proc. 16:137-144. Goldberg, S., Lebron, I., and D.L. Suarez. 2000. Soil colloidal behavior. In: Summer, M.E., (ed.) Handbook of Soil Science. CRC Press, Boca Raton, Florida. pp. B-195-B-239. Goldich, SS. 1938. A study in rock-weathering. J. Geo. 46:17-58. Gustafsson, J .P., Bhattacharya, P., Bain, D.C., Fraser, AR, and W.J. McHardy. 1995. Podzolisation mechanisms and the synthesis of imogolite in northern Scandinavia. Geoderma. 66:167-184. Haile-Mariam, S. and D.L. Mokma. 1995. Mineralogy of two sandy Spodosol hydrosequences in Michigan. Soil Survey Horizons, Winter2121-l32. Haile-Mariam, S. and D.L. Mokma. 1996. Mineralogy of two fine-loarny hydrosequences in south-central Michigan. Soil Survey Horizons, Summerz65-74. Haldane, AD. 1956. Determination of free iron oxide in soils. Soil Sci. 82:483-489. 128 Hall, RD. and J .S. Heiny. 1983. Glacial and postglacial physical stratigaphy and chronology, North Willow Creek and Cataract Creek drainage basins, eastern Tabacco Root Range, southwestern Montana, U.S.A. Arct. Alpine Res. 15(1):19 -52. Hall, RD. and RE. Martin. 1986. The etching of hornblende gains in the matrix of alpine tills and periglacial deposits. In: S.M. Colman, and DP. Dethier (eds.), Rates of Chemical Weathering of Rocks and Minerals. Academic Press, Inc., Orlando, F lorida:101-128. Hall, RD. and D. Michaud. 1988. The use of hornblende etching, clast weathering, and soils to date alpine glacial and periglacial deposits: A study from southwestern Montana. Geo. Soc. Am. Bul. 100:458-467. Hall, RD. and LL. Horn. 1993. Rates of hornblende etching in soils in glacial deposits of the northern Rocky Mountains (Wyoming-Montana, U.S.A.): Influence of climate and characteristics of the parent material. Chem. Geo. 105:17-29. Hoffman, SJ. and WK. Fletcher. 1981. Organic matter scavenging of copper, zinc, molybdenum, iron and manganese, estimated by a sodium hypochlorite extraction (pH 9.5). J. Geochem. Explor. 15:549-562. Huang, RM. 1989. Feldspars, olivines, pyroxenes, and amphiboles. In: J.B. Dixon, and SB. Weeds (eds.), Minerals in Soil Environments. SSSA Book Series 1:975- 1050. Huang, P.M. 1991. Ionic factors affecting the formation of short-ranged aluminosilicates. Soil Sci. Soc. Am. J. 55:1172-1180. Isard, SA. and R.J. Schaetzl. 1998. Effects of winter weather conditions on soil freezing in southern Michigan. Phys. Geog. 19:71-94. Jackson, M.L., Lim, CH, and L.W. Zelazny. 1986. Oxides, hydroxides, and aluminosilicates. p.101-150. In: A. Klute (ed.), Methods of Soil Analysis. Part 1. Physical and Mineralogical Methods. SSSA Book Series No. 9(1), SSSA and ASA, Madison, Wisconsin. J akobsen, B.H. 1991. Multiple processes in the formation of subarctic podzols in Greenland. Soil Sci. 152:414-426. Jenny, H. 1941. Factors of soil formation. McGraw-Hill, New York, New York, 281p. Johnson, LE. and CH. Chu. 1983. Mineralogical characterization of selected soils from northeastern United States. Pennsylvania Ago. Expr. Sta. Bull. 847p. 129 1.0.1. '. ”.4 It?! i. Klein, C. and CS. Hurlbut, Jr. 1999. Manual of mineralogy, 21St ed. John Wiley & Sons, Inc., New York, New York. 681p. Locke, W.W. III. 1979. Etching of hornblende grains in Arctic soils: An indicator of relative age and paleoclimate. Quat. Res. 11(2):197-212. Locke, W.W. III. 1986. Rates of hornblende etching in soils on glacial deposits, Baffin Island, Canada. In: S.M. Colman, and DP. Dethier (eds), Rates of Chemical Weathering of Rocks and Minerals. Academic Press, Orlando, Florida: 129-145. Loeppert, RH. and WP. Inskeep. 1996. Iron. p. 639-663. In: Bartels, J .M. (ed.), Methods of Soil Analysis. Part 3. Chemical Methods. SSSA Book Series No. 5(3), SSSA and ASA, Madison, Wisconsin. MacInnis, IN. and S.L. Brantley. 1993. Development of etch pit size distributions on dissolving minerals. Chem. Geo. 105231-49. Markos, G. 1975. Rates and kinetics of plagioclase dissolution in late Quaternary tills in Colorado and Wyoming. p. 1190. Geo. Soc. Am. Abstr. with Pro. 7(7). Annual Meeting. Mast, MA. and J .I. Drever. 1986. Effect of organic chelates on the dissolution rate of oligoclase. p. 685. Geo. Soc. Am. Abstr. with Prog. 99‘h Annual Meeting and Expo. McKeague, J .A. and J .H. Day. 1965. Dithionite- and oxalate-extractable Fe and Al as aids in differentiating various classes of soils. Can. J. Soil Sci. 46:13-22. McKeague, J .A., Brydon, J .E., and NM. Miles. 1971. Differentiation of forms of extractable iron and aluminum in soils. Soil Sci. Soc. Am. Proc. 35:33-38. Mehra, OR and ML. Jackson. 1960. Iron oxide removal from soils and clays by a dithionite-citrate system buffered with sodium bicarbonate. In: E. Ingerson (ed.), Clays and Clay Minerals. Monogaphy No. 5. Earth Science Series. Pergamon Press, New York, New York. p. 317-327. Messenger, A.S., Whiteside, ER, and AR. Wolcott. 1972. Climate, time, and organisms in relation to Podzol development in Michigan sands: I. Site descriptions and microbiological observations. Soil Sci. Soc. Am. Proc. 36:633-638. Mokma, D.L. and CV. Evans. 2000. Spodosols. In: Summer, M.E., (ed.) Handbook of Soil Science. CRC Press, Boca Raton, Florida. pp. E-307-E-321. Mokma, D.L. and GP. Vance. 1989. Forest vegetation and origin of some spodic horizons, Michigan. Geoderma 43:311-324. Not seen by this author. 130 Morton, AC. 1984. Stability of detrital heavy minerals in Tertiary sandstones from the North Sea Basin. Clay Min. 19:287-308. Muhs, DR. 1982. A soil chronosequence on Quaternary marine terraces, San Clemente Island, California. Geoderma. 28:257-283. Murakami, T., Kogure, T., Kadohara, H., and T. Ohnuki. 1998. Formation of secondary minerals and its effect on anorthite dissolution. Am. Mineral. 83:1209-1219. Nickel, E. 1973. Experimental dissolution of light and heavy minerals in comparison with weathering and intrastratal dissolution. Contributions to Sed. 1:1-68. Nesbitt, H.W., Fedo, C.M., and GM. Young. 1997. Quartz and feldspar stability, steady and non-steady-state weathering and petrogenesis of silicate sands and munds. J. Geol.105:173-191. Nebsitt, H.W. and G. Markovics. 1997. Weathering of ganodioritic crust, long-tenn storage of elements in weathering profiles, and petrogenesis of siliciclastic sediments. Geochim. et Cosmochim. Acta 61 :1653-1670. Nugent, M.A., Brantley, S.L., Pantano, CG, and RA. Maurice. 1998. The influence of natural mineral coatings on feldspar weathering. Nature 395:588-590. Ollier, C. and C. Pain. 1996. Regolith, soils and landforms. John Wiley & Sons, Ltd., West Sussex, England, 316p. Osei, BA. 1992. Sand mineralogy and related properties of some soils in south-westem Nigeria. Commun. Soil Sci. Plant Anal. 23:1815-1826. Parfitt, R.L. and CW. Childs. 1988. Estimation of forms of Fe and A1: A review, and analysis of contrasting soils by dissolution and Moessbauer methods. Aust. J. Soil Res. 26:121-144. Pettijohn, F.J. 1941. Persistence of heavy minerals and geologic age. J. Geo. 49:610-625. Pettijohn, F .J ., P.E. Potter, and R. Siever. 1973. Sand and Sandstone. Springer-Verlag, New York, New York, 618p. Price, T.W., Blevins, R.L., Bamhisel, R.I., and H.H. Bailey. 1975. Lithologic discontinuities in loessial soils of southwestern Kentucky. Soil Sci. Soc. Am. Proc. 39:94-98. Radcliffe, DE. and TC. Rasmussen. 2000. Soil water movement. In: Summer, M.E., (ed.), Handbook of Soil Science. CRC Press, Boca Raton, Florida. pp. A-87-A- 127. 131 Rai, D. and J .A. Kittrick. 1989. Mineral equilibria and the soil system. In: J .B. Dixon and SB. Weed (eds.), Minerals in Soil Environments, 2nd ed. SSSA Book Series 1:161-198. Rice, Jr., T.J., Buol, S.W., and SB. Weed. 1985. Soil-saprolite profiles derived from mafic rocks in the North Carolina piedmont: 1. Chemical, morphological, and mineralogical characteristics and transformations. Soil Sci. Soc. Am. J. 49:171- 178. Ross, G.J. and C. Wang. 1993. Extractable Al, Fe, Mn, and Si. In: Carter, M.R. (ed.), Soil Sampling and Methods of Analysis. Can. Soc. Soil. Sci., Boca Raton, Florida. p. 239-246. Ruhe, R.V. 1956. Geomorphic surfaces and the nature of soils. Soil Sci. 82(6):441-455. Ruhe, R.V., R.B. Daniels, and J .G. Cady. 1966. Landscape evolution and soil formation in southwestern Iowa. US. Dept. Agr. Tech. Bul. 1349. 242p. Schaetzl, R.J. and D.L. Mokma. 1988. A numerical index of podzol and podzolic soil development. Phys. Geog. 9:232-246. Schaetzl, R.J. and SA. Isard. 1991. The distribution of Spodosol soils in southern Michigan: A climatic interpretation. Annals Assoc. Am. Geogs. 81 :425-442. Schaetzl, R.J. 1992. Beta spodic horizons in podzolic soils (Lithic Haplorthods and Haplohumods). Pedologie. 42:271-287. Schaetzl, R.J. 1994. Changes in O horizon mass, thickness and carbon content following fire in northern hardwood stands. Vegetatio 115:41-50. Schaetzl, R.J. and SA. Isard. 1996. Regional-scale relationships between climate and strength of podzolization in the Great Lakes region, North America. Catena 28:47-69. Schaetzl, R.J. 2002. A spodosol-entisol transition in northern Michigan: Climate or vegetation? Soil Sci. Soc. Am. J. 66(4):1272-1284. Shetron, S.G. 1974. Distribution of free iron and organic carbon as related to available water in some forested sandy soils. Soil Sci. Soc. Am. Proc. 38:359-362. Soil Survey Division Staff. 1993. Soil survey manual. USDA Handbook No. 18. US Govt. Printing Office, Washington, DC. 437p. Soil Survey Laboratory Staff. 1996. Soil survey laboratory methods manual. USDA-SCS, National Soil Survey Center. Soil Survey Investigations Report. 42, Version 3.0, National Soil Survey Center, Lincoln, Nebraska. 693p. 132 Soil Survey Staff. 1998. Keys to soil taxonomy, 8th ed. Natural Resources Conservation Services, US Dept. Agri. 326p. Soil Survey Staff. 1999. Soil taxonomy: A basic system of soil classification for making and interpreting soil surveys, 2nd ed. USDA Agic. Handbook 436. US Govt. Printing Office, Washington, DC. 870p. Stahnke, CR. 1965. A study of mineral transformations and weathering processes occurring during the genesis of two soils developed from gneiss and schist in Llano Co., Texas. M.S. thesis. Texas Tech University, Lubbock. TX. Stephen, 1. 1952. A study of rock weathering with reference to the soils of the Malvem Hills. J. Soil Sci. 3:20-33. Sverdrup, H.U. 1990. The kinetics of base cation release due to chemical weathering. Lund University Press. Truog, E., Taylor, J .R., Jr., Pearson, R.W., Weeks, M.E., and R.W. Simonson. 1937. Procedure for special type of mechanical and mineralogical soil analysis. Soil Sci. Soc. Am. Proc. 1:101-112. Ugolini, RC. and C. Bull. 1965. Soil development and glacial events in Antarctica. Quatemaria 7:251-269. Ugolini, F .C., Minden, R., Dawson, H., and J. Zachara. 1977. An example of soil processes in the Abies Amabilis zone of central Cascades, Washington. Soil Sci. 124(5):291-302. Van der Plas, L. and AC. Tobi. 1965 . A chart for judging the reliability of point counting results. Am. Jour. Sci. 263287-90. Vance, G.F., Mokma, D.L., and SA. Boyd. 1986. Phenolic compounds in soils of hydrosequences and developmental sequences of Spodosols. Soil Sci. Soc. Am. J. 50:992-996. Velbel, MA. 1987. Rate-controlling factors in the weathering of some ferromagnesian silicate minerals. Transactions, 13th Congess of the International Society of Soil Science, Hamburg, Germany 621107-1118. Velbel, MA. 1989. Weathering of hornblende to ferruginous products by a dissolution- reprecipitation mechanism: Petrogaphy and stoichiometry. Clays and Clay Min. 37:515-524. Velbel, M.A. 1993a. Constancy of silicate-mineral weathering-rate ratios between natural and experimental weathering: implications for hydrologic control of differences in absolute rate. Chem. Geol. 105289-99. 133 Velbel, M.A. 1993b. Formation of protective surface layers during silicate-mineral weathering under well-leached, oxidizing conditions. Am. Mineralogist 78:405- 414. Velbel, MA. 1999. Bond strength and the relative weathering rates of simple orthosilicates. Am. J. Sci. 299:679-696. Walker, T.R., Ribbe, RH, and RM. Honea. 1967. Geochemistry of hornblende alteration in Pliocene red beds, Baja California, Mexico. Geo. Soc. Am. Bul. 78:1055-1060. Wang, C. and H. Kodama. 1986. Pedogenic imogolite in sandy Brunisols of eastern Ontario. Can. J. Soil Sci. 66:135-142. Wang, C., McKeague, J .A., and H. Kodama. 1986. Pedogenic imogolite and soil environments: Case study of Spodosols in Quebec, Canada. Soil Sci. Soc. Am. J. 50:711-718. Weisenborn, B.N. 2001. A model for fragipan evolution in Michigan soils. M.S. thesis. Michigan State University, East Lansing, Michigan. White, AF. and S.L. Brantley (eds). 1995. Reviews in Mineralogy, Volume 31: Chemical weathering rates of silicate minerals. Min. Soc. Am. Washington, DC. 583p. White, A.F., Blum, A.E., Schulz, M.S., Bullen, T.D., Harden, J ., and ML. Peterson. 1996. Chemical weathering rates of a soil chronosequence on ganitic alluvium: I. Quantification of mineralogical and surface area changes and calculation of primary silicate reaction rates. Geochim. et Cosmochim. Acta 60:2533-2550. Wilding, L.P., Smeck, NE, and LR. Drees. 1977. Silica in soils: Quartz, cristobalite, tridymite, and opal. In: Dinauer, R.C. (ed.), Minerals in Soil Environments. Soil Sci. Soc. Am., Madison, Wisconsin: 471-540. 134