MARINE GEOLOGY OF THE CONTINENTAL MARGIN OFF SOUTHERN OREGON Joseph John Spigai DUDLEY KNOX LIBRARY I POSTGRADUATE SCHOOL MONTEREY, CALIFORNIA 93943-5002 Marine Geology of the Continental Margin Off Southern Oregon by Joseph John Spigai A THESIS submitted to Oregon State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy June 1971 GRADUATE SCHOOB CALIF. 93940" APPROVED: Associate Professor of Oceanography in charge of major Chairman of the Department of Oceanography- Dean of the Graduate School Date thesis is presented Typed by Gwendolyn Hansen for Joseph John Spigai ACKNOWLEDGMENTS I would like to express sincere gratitude to Dr. LaVerne D. Kulm, my major professor, for his judicious advice and patient guidance during the entire course of this study. His generosity in giving so freely of his time and in critically reviewing this manuscript is greatly appreciated. Discussions with many people stimulated my thinking and helped me to formulate many of the ideas presented here. In this regard, thanks are due William Bales, and Drs. Richard Couch, Gerald A. Fowler, Michael S. Longuet-Higgins, and Eli A. Silver. Many fellow graduate students unselfishly contributed their knowledge and assistance to this work; thanks go especially to Drs. David W. Allen, John R. Duncan, Jr. , and Gary B. Griggs, and to David M. Chambers, Roger H. Neudeck, Robert E. Peterson, James B. Phipps, Robert C. Roush, and Kenneth F. Scheidegger. I am indebted to Dr. Gerald A. Fowler for providing the identifi- cation and correlation of the rock fauna, and to Dr. G. Ross Heath for his assistance in the analysis and interpretation of the clay minerals. I am also grateful to Miss Anastasia Sotiropoulos for her excellent work in drafting many of the figures. Appreciation is extended to the staff and students of the Department of Oceanography, Oregon State Univer- sity, and to the personnel of the R/V Yaquina for their invaluable assistance in the collection of data at sea. I am very grateful to my wife, Frances, for her enduring patience and for the unlimited help and encouragement she provided me during the course of the study. To my parents, whose sacrifices made my education possible, I am ever grateful. The opportunity to undertake this graduate study was made pos- sible by the United States Navy, under its Postgraduate Education Program; personal financial support was also provided by the Navy. The research pertaining to this study was made possible by grants from the United States Geological Survey made to Oregon State University (Contracts 14-08-0001-107 66, -11941, and -1287). Initial support for the research and ship time was provided by the Office of Naval Research (ContractNonr 1286 (10)). TABLE OF CONTENTS Page INTRODUCTION 1 SUBMARINE PHYSIOGRAPHY 5 General Physiographic Features 5 Physiographic Provinces of the Southern Oregon Continental Margin 9 Continental Shelf 12 Upper Continental Slope 13 Klamath Plateau 13 Cape Blanco Bench 15 Middle Bench 1 6 Lower Continental Slope 16 Rogue Submarine Canyon 17 REGIONAL GEOLOGY 22 General Geology of Southwestern Oregon and Northern California 22 Coastal Features and the History of Sea Level Fluctuations 31 SAMPLING AND ANALYTICAL PROCEDURES 33 Sample Collection and Processing 33 Laboratory Analyses 36 STRUCTURAL FEATURES OF THE SOUTHERN OREGON MARGIN 40 Regional Setting 40 Gravity 40 Magnetics 44 Heat Flow 47 Seismicity 48 Interpretation of the Tectonic Pattern 50 Continental Margin Structure 51 General Structural Trends 51 Continental Shelf 55 Upper Slope and Benches 58 Page Lower Slope 65 Rogue Submarine Canyon 67 Continental Margin Structure in Relation to the Regional Setting 7 0 STRATIGRAPHY 74 Unconsolidated Sediments 74 Planktonic Foraminifera-Radiolaria Abundance 74 Mazama Ash 78 Radiocarbon Dating 79 Faunal Stratigraphy of Consolidated Sediments 79 Rates of Sediment Accumulation 81 SEDIMENTOLOGY 85 Unconsolidated Sediments 85 Classification and Distribution of Sediment Types 85 Olive Gray Lutite 88 Gray Lutite 95 Sand-Silt Layers 96 Minor Sediment Types 97 The Present Sediment Pattern 99 Textural Relationships 100 Mineralogy 102 Light Minerals 106 Heavy Minerals 109 Clay Minerals 114 Organic Carbon 119 Consolidated Sediments 122 Classification and Distribution of Rock Types 122 PROCESSES OF SEDIMENTATION 128 A Proposed Model for Modern Sediment Transport on the Southern Oregon Continental Margin 128 The Initial Regime of Sedimentation 128 The Concept of Lutum Transport 133 The Oceanographic Regime 134 Turbid Layer Formation and Transport Across the Shelf 141 Transport Down the Slope 145 Deposition of Lutum on the Lower Slope 150 The Late Pleistocene and Holocene Sedimentation Patterns 153 SUMMARY AND CONCLUSIONS Development of the Structural Framework Pre-Tertiary History Tertiary History Quaternary History of Sedimentation Pleistocene Holocene and Modern BIBLIOGRAPHY APPENDICES Legend for Appendices Appendix 1. Piston core and rock dredge station locations Appendix 2. Radiocarbon age determinations from selected cores Appendix 3. Coarse-fraction composition of sedi- ment samples Appendix 4. Textural analyses of sediment samples Appendix 5. Light mineral composition of the sand fraction in selected sediment samples Appendix 6. Heavy mineral composition of the sand fraction in selected sediment samples Appendix 7. Quantitative analyses of the major clay minerals from X-ray diffraction records of selected samples 211 Appendix 8. Mineralogy of the major rock types from dredge hauls 212 Page 157 157 157 158 162 162 164 168 183 183 184 185 186 198 207 208 LIST OF FIGURES Figure Page 1. Oregon continental margin and adjacent deep-sea areas. 3 2. Bathymetry of the southern Oregon continental margin. 7 3. Precision Depth Recorder track lines in the area of investigation. 8 4. Physiographic provinces of the southern Oregon continental margin. 10 5. Selected bathymetric profiles of the continental margin. 11 6. Generalized bathymetry and selected profiles across the Rogue Submarine Canyon. 19 7. Geographic and drainage features of southwestern Oregon and northern California. 23 8. Geologic map of southwestern Oregon and northern California. 26 9. Location of piston core and dredge samples. 34 10. Flow chart for core processing and sample analysis. 35 11. Major tectonic features and earthquake epicenters in the northeast Pacific. 41 12. Free -air gravity anomaly map west of the southern Oregon-northern California coast. 43 13. Magnetic anomaly patterns off southern Oregon and northern California. 45 14. Fault-plane solutions of earthquakes and possible displacements from first motion studies in the northeast Pacific. 49 15. Track line map of continuous seismic profiles. 52 Figure Page 16. Major structural features of the southern Oregon margin. 53 17. Sparker profiles across the southern Oregon shelf and upper slope. 57 18. EDO subbottom profiles across the southern Oregon shelf edge and base of slope. 59 19. Sparker profile across Cape Blanco Bench. 60 20. EDO subbottom profile across the upper continental slope. 62 21. Sparker profile across the entire southern Oregon slope. 63 22. Sparker profile across the upper continental slope illustrating slumping. 64 23. Reflection profiles of the southern Oregon and northern California continental margins. 66 24. North-south Sparker profile across the Upper Rogue Canyon. 68 25. Summary of lithologic and stratigraphic information from piston cores from the southern Oregon margin and adjacent Blanco Valley. 7 6 26. Textural classification of unconsolidated sediments. 87 27. Coarse-fraction constituents of unconsolidated sediments. 89 28. Lithology of piston cores from the southern Oregon continental shelf. 90 29. Lithology of piston cores from the upper continental slope and benches. 91 30. Lithology of piston cores from the lower slope. 92 31. Lithology of piston cores from the Rogue Submarine Canyon. 93 Figure Page 32. Distribution of surface sediment types on the southern Oregon margin. 101 33. Phi Mean Diameter versus Phi Deviation. 103 34. Phi Mean Diameter versus Phi Skewness according to depositional environment. 104 35. Phi Mean Diameter versus Phi Skewness according to sediment type. 105 36. Light mineral composition of selected sediments and sedimentary rocks. 107 37. Major drainage basins of northwestern United States. 110 38. Clay mineral composition of selected sediments from the continental margin and major Oregon rivers. 115 39. Rock types from the southern Oregon margin. 123 40. Schematic model of modern sediment transport processes on the southern Oregon margin. 129 41. Schematic model of the oceanographic conditions over the southern Oregon margin in winter and summer. 135 42. Idealized cross -section of the late Pleistocene sediment distribution on the southern Oregon margin. 154 43. Idealized cross -section of the Holocene and modern sediment distribution on the southern Oregon margin. 156 LIST OF TABLES Table Page 1. Age, correlation and paleo-depth of Foraminifera from selected consolidated sediments on the southern Oregon continental margin. 80 2. Sedimentation rates from the southern Oregon continental margin, 82 3. A comparison of the light mineral composition of selected margin sediments with southwestern Oregon sandstones. 108 4. Comparison of the heavy mineral suites and pyroxene/ amphibole ratios of southern Oregon margin environ- ments with continental drainage basins. 112 5. Total carbon, organic carbon and calcium carbonate in selected margin sediments. 120 MARINE GEOLOGY OF THE CONTINENTAL MARGIN OFF SOUTHERN OREGON INTRODUCTION Investigations of the world's continental margins have expanded in recent years owing to increased political awareness, economic need, and a more sophisticated technology. More important, however, is the fact that both interest and understanding of continental margins have increased due to a deeper scientific insight into the basic processes of margin development and due to an awareness of the importance of con- tinental margins in the evolution of both continents and ocean basins. Recent theories concerning sea-floor spreading and plate tectonics (e. g. Vine, 1966; Isacks, e_t aL , 1968; Le Pichon, 1968) have had a major influence on the interpretation of both the geology of ocean basins and the geologic structure and evolution of the adjoining con- tinents. Because of these theories, new attention has been focused on the continental margins as the critical zones of transition between interacting continental and oceanic plates. An understanding of the tectonic history of continental margins in the light of sea-floor spread- ing must now be an essential prerequisite to help explain the processes of growth and development of ocean basins and continental masses. The continental margin off the state of Oregon has been the sub- ject of detailed investigation since 1962. The early work on the geo- morphology of the Oregon margin by Byrne (1962, 1963a, b) was followed by more detailed studies (Kulm and Byrne, 19 66; Byrne, e_t al . , 1966) as well as a number of other works. Geophysical studies have also been carried out on, or adjacent to, the Oregon margin (Dehlinger, etal., 1967, 1968; Emilia, etal., 1968; McManus, 1965; Shor, etal., 1968; Silver, 1969a, b). The Department of Oceanography, Oregon State University, in a joint effort with the United States Geological Survey (USGS), has sought to gain further and more detailed information about the Oregon con- tinental margin since 1967. Under this program an intensive and systematic study of the sediments and structure of this area is being carried out in an effort to better understand its geologic history, the dynamics of sediment movement in the water column and on the sea floor, and its possible economic mineral potential. Several published and unpublished reports describe preliminary findings of this program (Kulm, etal., 1968a, b; Clifton, 1968; Chambers, 1968; Mackay, 19 69; Bales and Kulm, 19 69; Kulm and Bales, 19 69; Spigai and Kulm, 1969; Roush, 1970; Neudeck, 1970). The present study of the continental margin off southern Oregon (Figure 1) is part of the overall USGS-sponsored program. Emphasis in this study has been placed on the sediments and structure of the con- tinental slope. Data concerning the continental shelf, derived from unpublished research, will be used as supplementary information to provide a more comprehensive picture of the margin as a whole. Figure 1. Oregon continental margin and adjacent deep-sea areas. Area of investigation is indicated by the horizontal rulings The objectives of this study are: 1. to determine the structure of the continental slope, its rela- tion to the structure of the entire margin, and its influence on the present geomorphology of the slope; 2. to delineate the major bathymetric features of the slope and determine their influence on the sedimentary processes acting on the slope, both past and present; and 3. to develop a three-dimensional picture of the surface and subsurface sediment distribution pattern, and relate this pattern to the dynamic sedimentary processes operating in the water column over the continental slope and on the sea floor below. SUBMARINE PHYSIOGRAPHY General Physiographic Features The geomorphology of the Oregon continental margin was first described in detail by- Byrne (1962, 1963a, b). The continental mar- gin off Oregon was called a continental terrace by Byrne and includes only the continental shelf and slope. The Oregon shelf is narrower, steeper and has its outer edge (the shelf break) in deeper water than the world's "average" shelf as given by Shepard (1963). It is also con- siderably wider in the north than it is in the south. The continental slope extends from the shelf break to a depth of approximately 3000 m, and is also wider, and has a gentler gradient, in the north than in the south. In general terms, the northern Oregon margin is marked by the appearance of the Astoria Canyon and Fan system, a ridge -trough com- plex below 800 m, and a prominent bench below 500 m, while the cen- tral margin is characterized by submarine banks such as Heceta, Stonewall, and Coquille (Maloney, 19 65). The southern margin is noted for the numerous slope benches at various depths, the Rogue Canyon, and numerous small submarine valleys. In describing the southern Oregon margin, Byrne (1963b) noted the appearance of "terraces" at various depths as well as the numerous 6 submarine valleys, including the Rogue Canyon. He also noted the apparent antecedent nature of the "terraces" in relation to the sub- marine valleys, implying that the valleys have subsequently dissected the pre-existing "terraces. " Byrne (1963b) compiled a bathymetric map in fathoms of the southern Oregon margin utilizing United States Coast and Geodetic Survey (USC&GS) smooth sheets of surveys done mostly in the mid- and late-1920's. More recently (1968), the USC&GS issued a bathymetric chart of the same region (USC&GS Chart 1308N- 17). This chart was reconstructed from the data of the 1920 surveys and recontoured in meters. It gives a somewhat more detailed picture of the submarine physiography. The southern half of this chart, which includes only the area of this study from Cape Blanco to the Oregon- California border (42°00'N to 42°50'N) is used as the bathymetric base map (Figure 2). Twenty-five hundred kilometers of bathymetric profiles were made on four Yaquina cruises (6706, 6708, 6711 and 6802) conducted during the course of this study, utilizing a Precision Depth Recorder. They have been used to supplement the existing bathymetric information (Figure 3). Navigational control was obtained using dual Loran A receivers and standard radar and navigational procedures which gave a maximum positioning error of ± 2 km. BLANCO — 42° 30' VALLEY PORT ORFORD 0 5 0 10 w m n i ^3 KILOMETERS CONTOURS \ IN METERS 42° 30'- Figure 3. Precision Depth Recorder track lines in the area of investigation. Physiographic Provinces of the Southern Oregon Continental Margin An analysis and comparison of previous work with the newer bathymetric data described above and combined with sub-bottom pro- files has enabled the writer to subdivide the southern Oregon continental margin into a number of discrete physiographic provinces (Figure 4). The provinces can be grouped as follows: Continental Shelf Upper Continental Slope Klamath Plateau Upper and Lower Plateau Slope Cape Blanco Bench Middle Bench Lower Continental Slope Lower Bench Minor Submarine Valleys Rogue Submarine Canyon As can be seen from Figure 4, the boundaries have been generalized somewhat to approximate the underlying structural limits rather than follow the exact bathymetric contours, which only indicate the boundaries of the present dissected surface. A number of selected bathymetric profiles across the margin, which show the relation of the various physiographic provinces with depth, are shown in Figure 5. 10 BLANCO —42° 30' VALLEY PHYSIOGRAPHIC PROVINCES 'CAPE BLANCO PORT ORFORD KILOMETERS CONTINENTAL ER TEAU I SLOPE SHELF 42° 30' - 124° 30 Figure 4. Physiographic provinces of the southern Oregon continental margin. 11 KILOMETERS rAPf BLANCO PORT ORFORD 42° 03' \ BROOKINGS Figure 5. Selected bathymetric profiles of the continental margin. 12 Continental Shelf The continental shelf off southern Oregon extends to depths vary- ing between 120 m off the Rogue River to 200 m off Port Orford. Directly off Cape Blanco the shelf break is difficult to determine because the Cape Blanco Bench abuts the shelf edge and results in a gradually decreasing gradient of approximately 1 ° down to a depth of nearly 500 m before the first sharp break in slope (Figure 5, profile at 42°50'N)„ In the area of investigation the slope width varies from 15 to 30 km; it is narrowest off Port Orford and Cape Sebastian and widest near the California border and just north of the Rogue River mouth. Chambers (1968) has described the southern Oregon shelf between 42°20'N and 42°40'N„ He particularly noted the meandering contours in this area, which indicates a thin sediment cover, and the exposure of bedrock such as the Rogue River Reefs. Similar outcrops or shoals, such as the Blanco Reef, occur southwest of Cape Blanco. Chambers also noted the occurrence of submerged terraces which are especially well defined at latitude 42°21'N. Here terraces occur at 35, 60, 70, 85, 100, 120 and 145 meters (± 5 m) and were most likely formed during stillstands of sea level which interrupted the last major trans- gression of the sea. 13 Upper Continental Slope With the exception of the continental shelf, all of the physio- graphic provinces recognized in this study lie within the boundaries of the southern Oregon continental slope between 42°00'N and 42°50'N. The slope is widest just off Cape Blanco and also south of Cape Sebastian (Figure 5). The steepness of the slope ranges from an average of 2° on the upper slope (above 1000 m) to an average of 5° on the lower slope (below 1000 m) because of the presence of benches. The upper slope is characterized by the presence of benches, plateau slopes, and the Upper Rogue Canyon. It extends from the shelf break to depths varying from 750 to 1250 m. The lower boundary is the lowest extent of the gently inclined benches and/or the beginning of the steeper lower slope. The benches are distinctive not only because they occupy the greatest area of the upper slope, but also because their gentler gradient suggests a different structural origin from the steep escarpments above and below (Figure 5). Klamath Plateau The largest bench on the southern Oregon margin lies between 500 and 750 m within the southern part of the upper slope between Cape Sebastian and the California border (Figure 5, profiles at 42°00'N to 42°17'N). This feature is the continuation of a large marginal plateau 14 which extends unto the upper slope off northern California as far south as 41°10'N. It has been named the Klamath Plateau by Silver (1969a) and this name is used here. The profiles of Figure 5 illustrate the moderate steepness (1°30') and relief of the Klamath Plateau and show how its width narrows from 25 km at 42°00'N to 1 0 km west of Cape Sebastian. Silver (1 969a, p. 1 1 ) noted a maximum width of 30 km for the Plateau on the northern California margin and also found that the Plateau slopes gently to the south. The relatively flat topography and bench-like appearance of the Plateau terminates abruptly with the appearance of a topographic high, or a series of highs, which mark its western edge (Figure 5, profiles at 42°03'N, 42°10'N, and42°17'N). Silver ( 19 69a) and Kulm and Bales (19 69 ) suggest that such topographic highs on the Klamath Plateau represent anticlinal folds. These folds have dammed, or ponded, the sediments behind them and have tended to smooth any underlying topo- graphic irregularities. Other benches on the upper slope probably have a similar origin; this underlying structural control may in fact be more extensive and continuous than the boundaries of the Klamath Plateau (Figure 4) would indicate. The Klamath Plateau, the Cape Blanco Bench (Figure 4; Figure 5, profile at 42°50'N) and the large bench areas on the upper slope north of Coquille Bank (Byrne, 1963b) may actually represent a single bench controlled by a single anticlinal fold system. The folds may even be contemporaneous in origin. The 15 ponded sediments have formed a single bench which has since been partially obscured by slumping and dissected by submarine valleys. Immediately above and below the Klamath Plateau are the Upper and Lower Plateau Slope respectively (Figure 5, profile at 42°03'N). The Upper Plateau Slope forms a transitional zone between the con- tinental shelf and the Klamath Plateau, while the Lower Plateau Slope forms a transition between the Plateau and Middle Bench. Cape Blanco Bench The Cape Blanco Bench is named after Cape Blanco on the Oregon coast (Figures 2, 4, and 5). Although there is an almost con- tinuous slope of about 1° from the shoreline at Cape Blanco down to a depth of about 500 m, the Cape Blanco Bench extends from about 250 to 500 m. As was previously noted, the Cape Blanco Bench and the Klamath Plateau may have formed as parts of a single bench system, the connecting segment between the two having since been obscured by the Upper Rogue Canyon and by slumping on the upper slope. The extremely dissected nature of the Cape Blanco Bench dif- ferentiates it from the Klamath Plateau. This can readily be seen in the profiles at 42°45'N and 42°43'N (Figure 5). The dissected nature of the Bench appears to be the result of a number of small north-south trending channels which are restricted to this portion of the Bench and may be structurally controlled. 16 Middle Bench The Middle Bench is so named because of its occurrence midway down the continental slope. It is exposed in the southern part of the slope below 42°05'N at depths between 1000 and 1250 m and occurs again at approximately the same depth at 42°3 6'N (Figure 5). It is apparent that the two exposures of the Middle Bench are part of a con- tinuous system, representing the second in a series of at least three bench systems which extend the length of the southern Oregon margin and probably beyond. From physiographic data alone, however, it is not possible to determine whether these three systems also represent three different structural episodes. The southerly exposure of the Middle Bench marks the transition of the upper to the lower slope on the southernmost Oregon margin. Lower Continental Slope The lower continental slope extends from a depth of 1000 to 1250 m down to 3000 to 3050 m where it meets the abyssal plain. However, south of Cape Blanco Bench the lower slope may begin at depths as shallow as 750 m. Between Cape Blanco and the Oregon-California border the lower slope has an inclination from 4° 20' to 8° 10'. In general the lower slope is marked by the existence of a num- ber of small submarine valleys which extend from the upper reaches 17 of the lower slope down to the abyssal plain. There are approximately eleven such submarine valley systems, excluding the Rogue Submarine Canyon. The two most prominent valley systems are called the Blanco Seavalley at 42°45'N, and the Brookings Seavalley at 42°05'N (Figures 2 and 4). Most of the valley systems drain only the lower slope, extending from a depth of about 1000 to 1250 m. They terminate on the sea floor at a depth of about 3000 m. Their average gradient is about 4°. The Lower Bench is also exposed on the lower slope between 42°05'N and 42° 15'N at a depth between 2250 and 2500 m (Figure 5). The topography shown in Figure 5 suggests that this bench may also be part of a continuous system at this level, the third such system identified on the southern Oregon slope. Exposures of this bench fur- ther to the north on the lower slope may be absent due to sedimentation. Byrne (1963b) cites other exposures of benches, or in his terminology "terraces, " at various depths, such as those occurring at 1600 to 1700 m at 42°50'N. Rogue Submarine Canyon Byrne (1963b) recognized the Rogue Submarine Valley and placed its terminus at a depth of approximately 1100 to 1200 m. However, the present study shows that this depth only marks what is now termed the Upper Rogue Canyon, and that this valley system is continuous to 18 the base of the slope to a depth of 3050 m„ The Upper Rogue Canyon heads on the continental shelf at a depth of 145 meters and is located some 12 miles northwest of the mouth of the Rogue River. From here it continues to a depth of about 1500 m in a generally east-west direc- tion. The Lower Rogue Canyon extends from this depth down to 3050 m at the base of the slope (Figure 2), and has a general northeast- southwest trend. As can be seen from Figure 2, the Lower Rogue Can- yon is actually a continuation of a northeast-southwest-trending swale which lies to the north of the Canyon and enters it at approximately 1500 m. The writer believes that the Canyon system is continuous, and although it may not be structurally continuous in its entirety, it is a single valley system from shelf break to base of slope. Figure 6 shows the major contours of the Rogue Canyon system together with a series of north-south profiles across the axis and a longitudinal profile down- axis. Structural control on the Upper Rogue Canyon is suggested in the profiles C-C, F-F', G-G1, and I-I', which show a notch or bench- like feature; this characteristic is not apparent on the profiles of the Lower Canyon. The north side of the Upper Canyon slopes 8°, the south side 15°; the steepness of both the northwest and southeast walls of the Lower Canyon averages 12°. The entrance of the swale, which marks the beginning of the Lower Rogue Canyon, is a gradual one. It occurs between 1500 and 1750 m (profiles J-J' and K-K1, Figure 6), 19 CONTOURS '*« -42° 30 DEPTHS 8 CONTOURS IN METERS VE X20 Figure 6. Generalized bathymetry and selected profiles across the Rogue Submarine Canyon. 20 but a noticeable change in the axial profile is evident in profiles K-K' and L-L,1. The axial slope of the Rogue Canyon varies over its extent, with the axis divided into five segments (Figure 6). The steepness of the axis varies down canyon from only 1°50' at segment AB to a very steep 14° at EF, the lowest segment. The steepness of the Upper Canyon is similar to that for the upper slope (2°, or 5° if the benches are . ignored) which indicates that the Upper Canyon may be more or less in equilibrium, or at grade with the surrounding upper slope. This would tend to indicate either that the lower slope is accreting or up -building faster than the Lower Rogue Canyon, or that the Lower Rogue Canyon may in fact be down-cutting. These generalizations do not take into account the added effects of uplift or subsidence. For example, during a period of uplift the Lower Canyon would down-cut faster and thus have a steeper gradient than the surrounding slope. The axis configuration of the Rogue Canyon may also be an important factor in determining whether a particular segment is eroding or accreting. The axis is distinctly V-shaped in the upper reaches of the Canyon, particularly between segments AB and BC. This suggests an erosional character, but this is not particularly borne out by the axial slope of these segments. The Upper Rogue Canyon may then owe its V-shape to a structural origin. The Lower Rogue Canyon has a generally more rounded shape. This indicates that sediment may be 21 accumulating in the axis of the Lower Canyon, although the steepness seemingly contradicts this. In short, physiographic evidence alone may not be sufficient to determine the character of the Rogue Canyon; how- ever, sedimentological and structural evidence to be presented later indicates that the Upper Rogue is stable or slightly erosional, while the Lower Rogue Canyon appears to be accumulating sediment. 22 REGIONAL GEOLOGY General Geology of Southwestern Oregon and Northern California The geologic history of the continental margin off southern Oregon is directly or indirectly related to the adjacent geologic provinces of the continent which lies landward of the margin. It is pertinent, therefore, to review the geology not only of the possible sediment source areas, but also of those areas whose geology and structural history may be related to that of the margin. All or part of the southern Oregon Coast Ranges, the Klamath Mountains, the Cas- cade Mountains, and the northern California Coast Ranges lie within southwestern Oregon and northern California (Figure 7). Oregon's Coast Ranges are a series of coastal hills composed of thick sequences of Tertiary volcanic and sedimentary rocks. They are divided into northern and southern ranges which together stretch for about 250 miles along western Oregon. The southern Coast Ranges consist mainly of Eocene sedimentary and volcanic rocks, with younger Mio-Pliocene and Quaternary formations present only near Coos Bay and Cape Blanco. The Siuslaw, Umpqua, Coos, and Coquille are the main rivers which drain this portion of the Coast Ranges. The middle fork of the Coquille River is considered to be the province's southern boundary (Figure 7). Structurally, the southern Coast Ranges consist 23 I24c ^V/> 123° NORTHERN BOUNDARY OF THE KLAMATH MT. PROVINCE, INCLUDING OUTLIERS I22c t o UJ UJ o < 62 u) was made with a standard Emery settling tube (Emery, 1938) as modified by Poole (1957). The per- centages of sand, silt and clay were determined for each sample; in addition, standard grain size parameters (Inman, 1952) were computed for each sample using a CDC 3300 computer (Appendix 4). Coarse - fraction analyses were made on approximately 350 samples to determine the compositional variations of the major sediment types within each core (Appendix 3). A split of at least 300 grains were identified and counted using a binocular microscope following a grid pattern. A count of the Radiolarian and planktonic Foraminiferan faunas was made on all samples where it was possible to obtain a combined count of at least 100 specimens for the two groups. Heavy mineral analyses were made on 33 selected samples. The heavy minerals were concentrated using tetrabromethane (specific gravity 2. 96). The concentrate was mounted on a glass slide in Aroclor (R. I. 1. 66) and at least 200-300 grains were identified and counted (Appendix 6). The majority of the mineral grains had a median 37 diameter between 62 and 125 u, but no attempt was made to determine the exact size variations within mineral groups or among various samples. The light mineralogy of twelve selected samples was analyzed using the feldspar staining techniques of Bailey and Stevens (I960); the tabulated results are given in Appendix 5. Organic carbon and calcium carbonate determinations were made on 1 6 selected samples representing the various sedimentary environ- ments using a LECO Automatic Carbon Analyzer. The method outlined by Peterson (1969, p. 44) was followed in preparing the samples; the formulas used in computing the percent total carbon, organic carbon and calcium carbonate were also taken from Peterson (1969). These values are recorded in Table 5 (see Sedimentology ), together with those samples taken from cores 6711-2 and 6706-5, which were analyzed by Peterson (1969). The clay mineralogy of 22 samples was determined using X-ray diffraction techniques; the samples were treated and prepared for examination according to the methods outlined by Heath (19 68) except for the use of the silver plug sample holder and the sample spinner described below. The treated suspensate from the < 62 (a fraction of each sample was allowed to settle until only the < 2 (jl fraction remained in suspension. Only this fraction was collected and used in subsequent analyses. In order to minimize differential segregation of the clay mineral particles, approximately 1-1.5 milliliter of the clay suspension 38 was applied by means of a replicating pipette to the surface of a porous silver plug which served as the sample holder. This procedure pro- duced a thin highly-oriented aggregate. A solution of 60% glycerol and 40% 0. 3 molar magnesium chloride was sprayed on the oriented sample, then dried for at least one -half hour and analyzed. This latter pro- cedure enhanced detection of the montmorillonite clays by expanding their lattices. All the clay mineral samples were analyzed on a Phillips -Norelco diffractometer with a Geiger-Muller counting tube. Samples were scanned from 2° to 14° 29 at a goniometer speed of 1/2° 20 /minute, and from 24° to 26° 20 at a goniometer speed of 1/ 8° 29 /minute. The resulting diffractograms were produced on a linear scale strip chart recorder running at 1/2 inch/minute. A sample spinner was employed which spun samples in the plane of reflection. The presence of montmorillonite, illite and chlorite was deter- mined from the diffractograms by identifying the peaks at 16. 8 A, 9. 95 A, and 7. 05 A, respectively. Using the criteria of Biscaye (1964), and comparing the 3. 54 A chlorite peak with the 3. 58 A kaolinite peak, it was concluded that no kaolinite was present in any of the samples and that this latter peak represented only chlorite. The semi-quantita- tive method of Biscaye (1965) for determining the relative proportions of montmorillonite, illite and chlorite by means of weighting factors was employed to interpret the diffractograms. The resulting percentages 39 for each of the three major clay minerals in each sample are provided in Appendix 7. Persistent minor peaks on several of the diffracto- grams were identified as amphiboles. 40 STRUCTURAL FEATURES OF THE SOUTHERN OREGON MARGIN Regional Setting The structural and tectonic pattern in the northeast Pacific adjacent to Oregon and California has been examined and interpreted by McManus (1965), Wilson (1965a, b), Vine and Wilson (1965), and most recently by Tobin and Sykes (1968), Menard and Atwater (1968, 1969), McKenzie and Morgan (1969), Silver (1969a, b), and Atwater and Menard (1970). The two fracture zones which lie directly west of southern Oregon and northern California are the Blanco Fracture Zone (Figures 1 and 11 A) and the Mendocino Fracture Zone (Figure 11A), respectively. The former is a transform fault (Wilson, 1965b) which separates the Juan de Fuca Ridge and the Cascadia Basin on the north- west from the Gorda Ridge to the southeast (Figure 1). The ocean area immediately north of the Mendocino Fracture Zone is one of active sea- floor spreading at present (Vine, 1966) and this undoubtedly has had an effect on the continental margin off northern California and southern Oregon. Gravity Gravity data for the southern Oregon margin and the adjacent ocean basin has been analyzed by Dehlinger (1969) and Dehlinger, 41 m o ID in 1/5 a. iii (TZ (TZ LUO uj° (- |-»- (£|- Z •"< o< LU bJ u QD o Q-o n 1 -J 1 _i UJ < O 42 e_t al . (1967, 1968, 1970), The area west of southern Oregon has near- zero average free-air anomalies (Figure 12) which indicates that the area is essentially in isostatic equilibrium (Dehlinger, e_t al_ . , 1970). In addition, the Blanco Fracture Zone does not appear to affect the bathymetry of the southern Oregon margin (McManus, 1965) nor the pattern of gravity anomalies (Figure 12). On the other hand, the southern Oregon margin does exhibit a pattern of three distinct gravity anomalies (Couch, 1970). A high nega- tive anomaly (-80 to -100 mgal) is located over the base of the southern Oregon continental slope (Figure 12). This anomaly may be due to the dip of the Moho beneath the continent and the overlying wedge of sedi- ments at the base of the slope which appear to thicken northward of 42°00'N (Couch, 1970). A positive anomaly of +1 to 410 mgal is situated near the edge of the continental shelf and suggests the presence of a structural high. Another negative anomaly (-40 to -60 mgal) is situated over the continental shelf south of 42°30'N and is especially well developed over the northern California shelf at about 41°00'N (Figure 12). Couch (1970) suggests that this anomaly could represent a large depression such as a basin or down-faulted structure which may be filled with sediment. A small positive anomaly of +20 mgal is present immediately southwest of Cape Blanco over the shelf and upper slope and may be related to a structural high. A distinct break in the gravity anomaly pattern is present on 43 44°r if I CAPE BLANCO 124' 123" 44c 43* g; CAPE MENDOCINO 42" 41' 40* I27« 126* 125' 124' 123* Figure 12. Free-air gravity anomaly map west of the southern Oregon-northern California coast. After Dehlinger, et al. , (1970). Shaded pattern denotes continental slope. Arrow (A) indicates the position of the surface trace of the west-southwest unconformity on the shelf. 44 the southern Oregon shelf (Figure 12, arrow A). The location and trend of this break corresponds to that of a west-southwest unconformity which Bales and Kulm (1969) have recognized as the dividing line between two regions of differing structural character. Magnetics The pattern of magnetic anomalies off the southern Oregon and northern California continental margins has been interpreted by Emilia, et al. ( 1968) and Silver ( 1969a, b). Silver ( 1969a) notes the presence of magnetic anomaly 3 (5 million years, Vine, 1966) near the base of the continental slope off northern California (Figure 13, D). He cites this as evidence for underthrusting of the continental margin in late Cenozoic time. Emilia, et al . (1968) have noted the anomalous magnetic character of the crust off northern Oregon and the presence of high susceptibility material which dips in toward the continent near the Mendocino Frac- ture Zone (Figure 13, A). Only minor negative and positive anomalies are observed over the continental shelf and slope off southern Oregon. The small 100 to 150 gamma negative anomaly is the only distinctive negative feature (Figure 13, A), while only small positive anomalies are seen at the base of the continental slope (Figure 13, B). It is not clear whether these small positive anomalies are a continuation of anomaly 2 (Emilia, et al. 1968; Figure 13, B and C) or whether they represent Magnetic anomaly patterns off southern Oregon and northern California. A. Total magnetic field. B. Positive magnetic anomalies. C. Magnetic anomalies in the northeast Pacific. Anomaly 3 (arrow) is the westernmost positive anomaly adjacent to the continental margin. D. Anomaly 3 as mapped on the continental slope. E. Water-depth contour map off Oregon (700 m contour interval). Shaded portion indicates shelf. Numbers indicate stations where source depths were cal- culated. F. Source depths to magnetic anomalies. Contours in 2 km intervals; source depths to nearest hundred m. A, B, E, and F after Emilia et aj.. (1968); (modified after Raff, 1966); D after Silver (1969b). 46 anomaly 3 noted by Silver (Figure 13, C and D). It is also not clear whether anomaly 2 of Emilia, et al . (1968), and anomaly 3 noted by Silver are the same. The source depth of the magnetic anomalies (Figure 13, E and F) on the southern Oregon and northern California continental slope vary between 3. 4 and 4. 5 km (Emilia, et al. , 1968), which agrees with similar calculations made by Silver (1969a) for anomaly 3. This would place the anomalies, depending upon water depth, within the second layer (consolidated sediments and basalts). According to Vine and Wilson (1965), the bulk of the magnetization rests in this thin (1 to 2 km) layer of basalts (layer 2) which overlies the main oceanic crust. Deeper anomalies have been noted by Emilia, et al. at the base of the slope off the Eel River in northern California (10. 1 km), and on the northern Oregon margin (12. 5 and 14. 6 km). The sources for these latter anomalies are deeper within the oceanic crustal layer. Figure 13, B indicates the distinct difference between the mag- netic character of the northern and southern Oregon continental shelf. Large positive anomalies are observed over the northern Oregon shelf, probably due to the presence of coastal volcanics which extend beneath the shelf of this area (Emilia, ejt al . , 1968). In contrast, the southern Oregon shelf is nearly barren of such anomalies which may reflect the low magnetic susceptibility of the probable metamorphic rocks beneath the southern shelf. Kulm, et al. (1968a) have analyzed magnetic data 47 of a more detailed nature for the continental shelf adjacent to the Rogue River and have interpreted the anomalies as possible evidence for placer deposits. Heat Flow Dehlinger, e_t al. (1970) report the heat flow values from 76 sta- tions measured in the northeast Pacific. Those stations north of the 2 Mendocino escarpment have an average heat flow value of 2. 4 |j.cal/cm sec. , which is one-third greater than the average for ocean ridges and nearly double the value for ocean basins. Mesecar (1968) and Korgen (19 69) have studied the sediment temperature gradients and heat flow values over the Oregon continental slope. Mesecar measured heat flow values at eight stations down the continental slope and on the adjacent abyssal plain and noted a linear increase in the value of both the sedi- ment temperature gradient and the heat flow with depth. He found the 2 mean heat flow values on the slope to be 2. 0 ucal/cm sec. and that 2 on the abyssal plain to be 3. 2 ucal/cm sec. Thus the continental slope, and the deep sea environment adjacent to it are areas where the heat flow is higher than normal, reflecting the presence of high heat sources, which in turn may suggest the presence of a low density con- vecting mantle. 48 Seismicity The seismicity of the areas adjacent to the southern Oregon mar- gin has been examined by Tobin and Sykes (1968) and Bolt, e_t al . (1968). Earthquake epicenters concentrate along the Gorda Ridge, the Mendocino Fracture Zone east of Gorda Ridge, and the Blanco Fracture Zone between Juan de Fuca and Gorda Ridges (Figure 11, B). Accord- ing to Dehlinger (1969), fault-plane solutions (Figure 14) indicate that all but one of the relative motions in this area are consistent with ten- sion in an east-west direction. Hamilton (1969) also postulates that crustal extension has been the dominant tectonic movement during most of Cenozoic time along the western United States. Bolt, et al. (1968) postulate that the fault-plane solutions on the shelf north of Cape Mendocino and in Gorda Basin are consistent with a right-lateral strike-slip which trends about N40°W. They suggest that distortion of the oceanic crust within Gorda Basin, as well as some degree of underthrusting along the continental margin may account for the east- ward component of strike-slip. Dehlinger's interpretation (1969) of tension in an east-west direc- tion negates the existence of regional compression and casts doubt on the possibility of underthrusting along the southern Oregon-northern California margin. However, Byrne, e_t al . ( 19 66) and Silver (1969a, b) have cited evidence of compressional features on the margin off central 132° 130° 128° 126° 124° 122' 49 Figure 14. Fault-plane solutions of earthquakes (numbers 1-23) and possible displacements from first motion studies (num- bers 24-34) in the northeast Pacific. Arrows indicate horizontal component of displacement; N indicates nor- mal faulting. After Dehlinger, et^ al. , (1970). 50 Oregon and northern California, respectively. Silver (19 69a) notes that the presence of anticlinal and synclinal folds on the northern California margin corroborates his view of underthrusting in that area. Compressional features on the margin may also be attributable to local vertical forces rather than to large-scale underthrusting; hence, the origin of these features is an open question at present and one which requires further analysis. Interpretation of the Tectonic Pattern The tectonic pattern of the northeast Pacific is a complex one. Silver (1969a, b), Dehlinger, et al . (1970), Bolt, et al . (1968), Menard and Atwater (1968, 1969), Atwater and Menard (1970), and others have developed various theories for the large-scale tectonic movements involving the interaction of oceanic and continental lithospheric plates. A general consensus of these workers indicates a predominantly east-west direction of movement of these plates prior to Miocene time. From Miocene time to the present, however, the movement of oceanic and continental plates (i. e. the Pacific andNorth American plates, respec- tively) with respect to one another has shifted direction to a northwest- southeast trend. This major component of movement is a right lateral strike-slip motion along the San Andreas -Queen Charlotte Fault systems (Figure 11, A) accompanied by a smaller component of eastward under- thrusting of the oceanic plate against the continent. 51 Continental Margin Structure The structure of the continental margin off southern Oregon has been investigated with the aid of detailed continuous seismic-profiling surveys conducted during 1967, 1968, and 1969. Approximately 2000 km of track lines have been made off the southern Oregon shelf and slope between Cape Blanco and the Oregon-California border (Figure 15). The profiling equipment used included an EG&G Sparker with both 5, 000 and 10, 000 joule capacity, a 20 cu. in. Bolt air gun and a 3. 5 KHz Edo transducer. Based on an analysis of these continuous seismic profiles, a map has been compiled (Figure 16) which illustrates the major structural features of the southern Oregon margin. The struc- ture of the continental shelf illustrated in Figure 16 has been derived from the interpretation of Bales and Kulm (1969). General Structural Trends The trend of both the major fault and fold systems on the margin is generally north-south, parallel or sub-parallel to the margin (Bales and Kulm, 1969). The major exceptions to this trend are the east-west faults running through the Upper Rogue Canyon and the several east- west folds on the continental shelf south of Cape Sebastian. It appears from Figure 16 that the continental shelf and upper slope are character- ized by folds, whereas the lower slope topography may owe its origin 52 Figure 15. Track line map of continuous seismic profiles. Shaded area denotes the continental slope. After Kulm and Fowler (1970). 53 ,oo^> CO •~-i c CD c c o O ^_^ c a • -* vO :jj CT^ tl rt - . 5 c c| . — 1 n 3 ^L W CI) u T3 0 c CO tl a) 11 i — i si cd -i-> 3 £ o <^ CO o CD a 0 U-, d o •*-> 0) CO !h 0 a. tl ti d CD rd C CD --1 0) 1 — 1 n a) 4-> u 1 tl 4-J 0) f 1 ^J IH J rrt tl -*-> T^ CO CD tl o • -H •1 — > T3 Id 0 2 g CD ti 3 54 to large-scale faulting. A number of small-scale anticlinal folds appear on the reflection profiles on the lower slope which have not been plotted in Figure 16. These folds are smaller in amplitude and wave- length than those of the upper slope and do not appear to detract from the general trend noted above. Nearly all of the faults determined, or inferred, from the reflec- tion profiles and shown on Figure 16 appear to strike in a north-south direction. They apparently are high-angle or nearly vertical and down- thrown to the west. This has also been observed by Silver (1969a, p. 24). These faults appear to be normal faults; although data sufficient to determine their true dip directions and attitude is lacking. The fault pattern on the lower slope appears to be one of a typical block-faulted region. Maloney (1965) classified similar faults off central Oregon as step faults and the intervening benches and hills as step fault blocks, although he was unable to determine if any displacement had occurred. Step faults differ from normal faults in that they are compressional in nature and both the hanging wall and foot wall move up, the foot wall moving up farther than the hanging wall. The net displacement is similar to that of a normal fault. The faults on the southern Oregon margin appear to be normal faults which are tensional in origin. If this is correct, then it must be reconciled with the apparently compressional origin of the folds which exist on the same margin (Figure 16). Even if both faults and 55 folds result from compressional forces, then the interpretation of east-west tension noted earlier from fault-plane solutions (Dehlinger, 19 69) would still have to be explained. A solution to this problem might be found by assuming a mech- anism where both tension and compression are possible. Such a mechanism has been postulated by Malahoff (1970) to account for simultaneous underthrusting of oceanic crust under island arcs and the formation of gravity faults in trenches; he suggests that a trans- formed thrust mechanism may exist under trenches, whereby a gravity fault located at the trench surface would become a thrust fault at depths of 30 to 40 km beneath the island arc. Assuming that the southern Oregon continental margin at its interface with the adjacent ocean basin presents an analogous situation, then the existence of normal gravity faults in an area of active underthrusting can be explained. Continental Shelf The southern Oregon continental shelf can be divided into two regions which do not appear to be structurally related (Bales and Kulm, 19 69). The dividing line of the two regions is the surface trace of a west-southwest trending angular unconformity within the late Cenozoic sedimentary sequence between the Rogue River and Cape Sebastian (Figure 16). This division can also be noted on the gravity map (Figure 12, arrow A). A series of parallel and subparallel 56 northwest-southeast trending folds characterize the northern region, with folds becoming gentler on the outer shelf and increasing in number towards the coastline. The Gold Beach shear zone may extend off- shore in a similar northwest-southeast trend. This latter feature reflects the northwest-southeast trend of shears found in northwestern California which were formed during the late Cenozoic Cascadan orogeny. A north-south trending angular unconformity between the late Cenozoic sedimentary sequence and the later Quaternary deposits is present along the shelf break southwest of Port Orford (Figure 17, A, at 26 km). The wedge of sediment above the unconformity suggests that the shelf may be prograding in this area. The unconformity apparently terminates north of the head of the Upper Rogue Canyon and is not observed south of the Canyon. The southern part of the shelf is characterized by a southwest- trending basin off Cape Sebastian which extends westward across the entire shelf (Figure 17, B) and southward to the southernmost portion of the shelf (Figure 16). It appears to be only slightly deformed by local folding and warping near the coastline (Bales and Kulm, 19 69; Figure 16). West of this basin, an anticlinal fold underlies the shelf edge (Figures 16 and 17, B). A north-south unconformity, similar to the one in the northern region, lies between this anticline and the upper slope (Figure 1 6). Several prominent reflectors observed on the seismic records at fM 57 c\j CO CM' ro CM' LU ^ or cm' o x CO o m CNJ 2 X. CD CNJ" CNJ 00 CM' 0> CM" (w) Hld30 H31VM o o o o o o o o o o o o cm ro qo i & w v C CU cq o u a CQ CD a a) u en CO en o ^ ^ d * 5 0 h h o CL c J-l 2 o. g Uj * O I o £ M HldSO 831VM O g I (09S) 3M± 13AV&1 AVM-Z oo Q W a o c v c •u C o u ^ a a d »— i 3 rd o s >H 0) O h c -1_J C QJ nj U C 0) ■-1 tn 'XI 0 (U t* C u ri '-' a> TJ 01 <*-! j2 O u to a T3 , | Sh ni

a Sh a Oj CO > (IV) Hld30 U31VM * °J P's; j/bj T3/WU1 tVM-S DC 64 c c 0) OJ U rti 4-1 u n • pH n Tfl <> C M Cfl ^ CL . h y, OJ X X Jh O rd o a rvj a) ~T r\) fM OJ ^ 3 DU h (oas) 3M± 13AV81 AVM-3 65 Lower Slope The lower slope consists of small-scale, north-south trending anticlinal folds and normal faults down-dropped to the west. They form a series of sediment traps and step fault blocks down to the base of the slope. The Lower Bench (2250 to 2500 m) is the most prominent bench- like structure on the lower slope. The most prominent normal fault is that forming a marginal escarpment at the base of the slope (Figure 21 at 64 km). Silver (1969a) also postulated the existence of such a fault and others similar to it on the northern California margin. He finds structural trends similar to those off southern Oregon (Figure 23). He has recognized the Klamath Plateau and other ponded sediment basins and benches and has described the northern California margin as a series of north-south trending anticlinal folds and high-angle normal faults down-dropped to the west (Figure 23). He believes his interpretation of the structure is consistent with compression and underthrusting directed broadly from the west. In general the sediment cover drapes conformably over the under- lying structures on the lower slope and appears to thicken near its base. The thickest section is found at the base of the lower slope and on the adjacent abyssal floor (Figures 18, A and D and 21). This thick section (Figure 23, lineO) consists of flat-lying, relatively undeformed strata which conformably overlie the irregular basement topography; 67 however, deformation of the sediments appears to increase with depth. A continental rise is not well developed off southern Oregon and probably reflects the lower sediment supply from the Rogue, Klamath and other rivers which feed the area. In contrast, the continental rise off the east coast of the United States results from the much larger influx of sediment supplied by the numerous large coastal rivers. Rogue Submarine Canyon The Upper Rogue Canyon, situated from 150 to approximately 1500 m, is an east-west trending submarine canyon that is predomi- nantly structurally controlled but exhibits erosional features (Figure 24). The Upper Canyon has about 400 m of relief and a distinctly sharp V-shape. An east-west fault can be inferred from the rapidly- steepening dips observed on either side of the Canyon which suggest down-dragged strata, in addition to what appears to be a fault sliver on the south wall (Figure 24). Other seismic reflection profiles across the Upper Canyon show similar fault features. The fault is possibly split into two branches: one strikes along the axis of the Canyon and the other along the steeper south wall which has been down-dropped to the north forming a small bench along part of the channelway (Figure 24). The fault terminates eastward on the shelf against a series of small-scale en echelon faults at the head of the Canyon (Figure 16). Little, if any, sediment accumulation exists in the Upper Rogue Canyon, ROGUE CANYON N 0 (km) 5-0.2 V.E. /Ox Figure 24. North- south Sparker profile across the Upper Rogue Canyon (SP-71, 124°50'W). 69 suggesting that it may still be erosional in nature. The Lower Rogue Canyon, trends in a northwest-southwest direction and extends from 1500 to 3000 m. It has a much broader cross-sectional profile (e.g., Figure 6, profiles K-K' and L-L') in contrast to the Upper Canyon. It is at least in part an extension of the northeast-southwest trending swale (Figure 2) which is located to the north of the Rogue Canyon. The two valleys join at about 1500 m and continue down to the base of the slope as one system. Good seismic reflection profiles from the Lower Canyon are lacking, but it appears from cores taken in the axis of the Lower Canyon that this portion may be filling. The Lower Canyon does not appear to be offset by the Upper Canyon fault and hence may be younger than the Upper Canyon. The writer concurs with Bales (1969) in the belief that the fault in the Upper Canyon may be of late Tertiary age, as is the S-shaped anticline west of the Rogue River mouth (Figure 16). Erosion of the Rogue Canyon was probably due to the influx of coarse clastic sediment during the late Pleistocene, when sea level was only slightly above 150 m. Other evidence (See Sedimentology) indicates that the late Pleistocene coarse sediment may have poured into the incipient faulted channelway during this time, initiating the development of the Upper Rogue Canyon. No buried channels have been found on the continental shelf to indicate a possible connection between the Rogue River mouth and the 70 head of the Rogue Canyon (Kulm, et al . , 1968a). However, the eastern flank of the S- shaped anticline to the west of the Rogue River could have acted as a pathway for sediment transport, during a lowered sea level. In a similar manner, other shelf structures, such as those now observed southeast of Port Orford, could have acted as pathways for sediment transport north along the shelf by-passing the Upper Rogue Canyon. If such a flow were diverted southwestward across the slope by banks or rock reefs, a feature similar to the swale on the could easily have formed. Continental Margin Structure in Relation to the Regional Setting Various theories can be developed to explain the evolution of the southern Oregon margin, each of which is dependent upon a different interpretation of the regional tectonics in the northeast Pacific. One interpretation of the magnetic data suggests that ocean-floor spreading is occurring eastward from the Juan de Fuca and Gorda Ridges and that this new crust is underthrusting beneath the southern Oregon and northern California continental margin (Silver, 1969b). However, based on fault plane solutions Dehlinger, (19 69) and Dehlinger, et al. (1970) believe that the region is presently experiencing tension in an east-west direction. The structure of the southern Oregon margin does not resolve this dilemma. The north-south fold system indicates compression 71 directed from the west, while the north-south faults are ambiguous in that they appear to be normal tensional faults down-dropped to the west. They may also be compressional step faults (i„ e. vertical faults where the movement is mainly in the horizontal plane) or simply slump features produced by sediment loading. If underthrusing is occurring along the southern Oregon margin, the leading edge of the descending oceanic lithspheric plate could produce components of both tension and compression. Tension would result in subsidence and down-dragging at the base of the slope, giving rise to normal gravity faults. Compres sion would produce uplift directed to the east forming compressional north-south trending folds. If such a mechanism operated along a structural hinge line along the present shelf break or upper slope, the resultant structural pattern would be similar to that mapped in Figure 16. In an analogous situation that was described previously, Malahoff (1970) was able to reconcile both tension and compression existing in an area of active underthrusting. It would appear then that the com- plex of structures on the southern margin are compatible with under- thrusting. Several additional points need to be considered in attempting to explain the margin structural pattern. First, the margin off southern Oregon and northern California may have been more affected by com- pression than the margin off central and northern Oregon. Second, the warped coastal marine terraces near Cape Blanco indicate 72 differential uplift in the vicinity of Cape Blanco. Third, faunal studies (See Stratigraphy) indicate that both uplift and subsidence has occurred on the southern Oregon margin. These facts indicate a complex struc- tural history for the southern Oregon margin; one that has varied with time. Several current interpretations of the regional tectonics do take into account changing structural trends with both direction and time, and may serve to explain many of the complex structures on the southern Oregon margin. Dehlinger (1969) has postulated that much of the move- ment of crustal plates, especially between the Gorda and North American plates, may be due to differential slippage caused by the varying effects of tidal friction on the low velocity layer between the plates. He also notes that the interaction of two plates at the continental margin need not be resolved by underthrusting alone, but can also be accounted for by local vertical uplifts. These uplifts may express themselves as fold-like diapiric structures which are not compressional in nature, and probably not unlike the small-scale folds observed on the margin in Figures 20 and 21. Silver (1969b) has postulated a general eastward movement of the Gorda Plate, but with a larger component of right lateral slip of the plate along the continental margin which moves the plate mostly in a north-northwest direction. Wilson (1965a) first theorized that sea-floor spreading and the resulting direction of plate motion has changed to its present northwest-southeast sense due to 73 rapid outbuilding of the margin sometime in the mid-Teriary, perhaps since Miocene time. Atwater and Menard (1970) have shown that this change in direction of sea-floor spreading may have begun as early as 55 million years ago (Eocene time). These theories suggest that some degree of spreading and under- thrusting occurred on the margin prior to mid-Tertiary time. Sub- sequent to mid-Tertiary time the direction of movement of continental and oceanic plates shifted to the northwest-southeast. Late Pliocene or early Pleistocene may have been a time of renewed underthrusting (Silver, 1969a); if so evidence for this episode should be found in the structure and unconformities on the northern California and southern Oregon margin. The current effects of underthrusting are considered by Silver (1969b), Dehlinger (1969), and others to involve only the southern portion of the Gorda Plate, affecting only that part of the mar- gin lying between Cape Sebastian (Oregon) and Cape Mendocino. The northern part of the Gorda Plate above Cape Sebastian may not be actively underthrusting at present, as is suggested by the fact that sedi ments have begun to accumulate at the base of the northern Oregon- southern Washington slope in the form of submarine fans, which may eventually coalesce to form an incipient continental rise. Localized uplift and subsidence may also play an important role in the present structural evolution of the southern Oregon margin. 74 STRATIGRAPHY Unconsolidated Sediments The geologic history of the southern Oregon margin can be inter- preted through a correlation of the various sedimentary units which occur on the margin and their continental counterparts. The strati- graphic work of Duncan (1968), Duncan, et al. (1970), and Griggs, et al. (1970) in the deep-sea environments west of the Oregon margin provide a basis for comparison with the results of this study and with other interpretations of the faunal stratigraphy of the Oregon margin, Planktonic Foraminifera-Radiolaria Abundance Duncan (1968) and Duncan, et al. (1970) reported the development of the first paleoclimatic stratigraphy for the deep-sea environments off Oregon, which is based mainly on the Radiolaria (R)/planktonic Foraminifera (P) ratios during the late Pleistocene and Holocene. Griggs (1969) and Griggs, et^ al. (1970) expanded this stratigraphy to give a more detailed paleoclimatic curve for the last 12, 500 years in Cascadia Basin. The stratigraphy developed by both Duncan and Griggs closely matches the glacial advances and retreats noted by Armstong, e_t al_. (1965) on the continent. Since 35,000 years Before the Present (B. P. ), there have been three warm periods recognized within the Pleistocene and marked by an increase in Radiolaria within the pre- dominant planktonic Foraminifera interval. These have been dated at 75 34, 000 years B. P. , 28, 000 to 25, 000 years B. P. , and 18, 000 to 16, 000 years B. P. The most abrupt change in the fauna occurred about 12, 500 years B. P. in Cascadia Basin and is considered the end of the Pleisto- cene and the beginning of the Holocene (or post-glacial) in this area (Duncan, 1968). The pelagic and hemipelagic sediments high in plank- tonic Foraminifera, principally Globigerina pachyderma and G. bulloides, characterize cooler or glacial periods, while Radiolaria-rich sediments characterize the warmer or interglacial periods. The abruptness of the transition depends on the sedimentation rate, which determines the length of the transitional interval between the Holocene and late Pleistocene. Griggs ( 1 9 69 ) and Griggs, et al_. (1970) noted that several cooler periods, marked by Globigerina-rich intervals, occur within the Holocene since 12,500 years B. P. These have been dated at 5000 to 4000 years B. P. and 2000 years B. P. ; the former can be dated precisely since it usually occurs just above a Mazema ash layer which dates at 6600 years B. P. , while the latter is somewhat tenuous. In an effort to determine whether or not these stratigraphic hori- zons occur in the continental margin deposits, an analysis similar to that performed by Griggs and Duncan was employed with the sediments examined in this study. Figure 25 illustrates the R/P ratios deter- mined for five cores (6706-2, 6711-1, 6711-2, 6711-6, and 6711-8). In each core a combined count of at least 100 Radiolaria and/or planktonic Foraminifera were obtained. The high sedimentation rates, LOWER SLOPE ROGUE CANYON .-*.-_. zzr^r ^ ^ CONTINENTAL SHELF 1 □ Figure 25. Summary of lithologic and stratigraphic information from piston cores from the continental margin and adjacent Blanco Valley. Data for box cores 6708-44 and 6802-BCB after Chambers (1968). 77 particularly evident in the lower slope cores, caused the dilution of the biologic material by detrital grains and precluded the calculation of R/P ratios in most other cores. The variation in the R/P ratio in cores 6706-2, 6711-1 and 6711-2 have been dated using radiocarbon dates and/or the presence of Mazama ash. As can be noted from Figure 25, the R/P ratios in core 6706-2 (upper slope and benches) indicate an age for the core which ranges from 28, 000 to 25, 000 years B. P. near the bottom to 18, 000 to 1 6, 000 years B. P. near its surface. Since the top of the core is considered to be present, it appears that the surface sediment in this core and that from the surrounding area may be relict sediment of Pleistocene age. The low R/P ratio in cores 6711-1 and 6711-2 (lower slope) probably represents the 5000 to 4000 years B. P. interval of Griggs, et_ al . (1970). This interval correlates with a similar interval in cores 6604-12 and 6604-11 from Blanco Valley (Figure 25), and reflects similar sedimentation rates for the two environments. The ages inferred by the R/P ratios plotted for cores 6711-6 (lower Rogue Canyon) and 6711-8 (lower slope) have not been confirmed by dating. The topmost Foraminifera-rich zone in each core may represent the last cool period in the Holocene (2000 years B. P. ); however, this would produce an unreasonably high rate of sedimentation. It is more reasonable to assume they represent the 5000 to 4000 years B. P. interval until more precise dating can confirm their age. 78 Mazama Ash The Mazama ash which erupted from Mt. Mazama (Crater Lake, Oregon) 6600 years ago and blanketed large areas of the Pacific North- west is a good stratigraphic horizon in deep-sea and continental mar- gin sediments because of its widespread occurrence. Nelson, et al. (19 68 ) describe the occurrence of the ash in the various deep-sea environments of Cascadia Basin and postulate that ash deposition was effected by turbidity currents and not through aerial fallout. They sug- gest that the ash accumulated on the continental margin prior to its transport to the deep sea. The ash identified in this study has been verified as Mazama ash through comparison of its refractive index (1. 505) with known samples as identified by Nelson (1968), Duncan (1968) and others. The depth of maximum abundance of the ash has been used as the stratigraphic horizon dating approximating 6600 years B. P. in six cores (6711-1, 6711-2, and 6708-38, lower slope; 6706-6, upper slope; and 6708-23 and 6708-25, shelf) from the con- tinental margin (Figure 25). While the depth of occurrence of the ash varies within the margin or deep-sea environments, it does reflect differences in sedimentation rates. It is significant to note that the ash was found in only one upper slope core (6706-6) which suggests that older sediments persist to the surface in upper slope environments, or at least that the sedimentation rate is markedly lower here. 79 Radiocarbon Dating In order to establish the absolute ages of certain lithologic or biologic events recorded in the cores, and to facilitate correlation 14 12 among cores, the radiocarbon (C /C ) dating method was employed. Samples from five selected cores (6706-2 and 6706-3, upper slope; 6711-2, lower slope; 6708-37, Rogue Canyon; 6708-42, shelf) were dated utilizing the total carbon content (Figure 25; Appendix 2). Two other dates were obtained by Chambers (1968) from box cores taken on the southern Oregon shelf (6708-44 (42°35. l'N, 124°41.0'W, 150 m) and 6802-BCB (42°35. l'N, 124°41. l'W, 150 m)); these utilized car- bonate carbon from mollusk shells found in the cores. Faunal Stratigraphy of Consolidated Sediments The benthic and planktonic Foraminifera in two selected dredge haul samples (6708-36-1 and 6802-D3-1) and from consolidated rock fragments in the basal 20 cm in one piston core (6706-3) were identified and described in detail by Fowler (1970). His conclusions regarding the faunal assemblages are shown in Table 1. Both uplift and subsidence are inferred from the nature of the faunal assemblages in the consolidated rocks. In the two samples from the upper slope (6708-36-1 and 6706-3-11-2) the faunal data suggest that uplift has occurred on the southern Oregon margin. These data 80 Q o nj E E to -rt rt Q M 3 rt to 60 1 o 2 E CO > ^^ o 2 S 3 TJ E Ol U s +-» , o rt •iH QJ B E o 00 £ 1 CM rt o a. LO ft >- H o (32 O o o ft 3 (-0 rt o 0) a rt V cu 32 E co lO 43 3 CM t/j 1 E a) E fa rt E r^. rt QJ ^ i 3 O rt "rt 0 -rt rt 5 41 u U 3 rt ■t-J rt a, 0) O -rt rt U 3 o 2, 5 O rt u 3 rt (LI rt rt O u m QJ | ID O TJ ■!-H 01 0 a h 0) 1 ts TJ O TJ (LI 42 ■!-> rt o s QJ 1 rtj 01 fa rt o 1 01 42 E s bO £ ^ TJ s a at rt .rt 4-> rt CO 3 = 3 ai 01 J2 > ^ . .« 3 a. o I rt £ TJ rt b0 _ i! « i u a. CO ■c rt ai >> 3 O Oi 4£ U ni w TJ w rt 3 rt oi rt ■S E E 5 2 ° ro to ^ p OJ 3 o +-> CO -3 o o 00 o 2 QJ n OI ■rt 42 T3 T3 CO 00 3 ■rt 42 ■rt 4tf Crt ^— * rt O LO 4-1 CM >- I — 1 VO 43 42 II TJ 1 LO 0) 01 CM 3 a -H •r-l rt o + 4-> DO 43 CJ 0I 0 +-' S| OI o DO rt • •% 3 rt u LO rt 43 3 CM 42 CO II ft rt o CM 01 TJ 1 S o LO O • rt CM 01 TJ TJ rt rt o ft TJ LO CM 01 a II 42 <« <4-l o 01 bo 3 o lo| ■M rt a rt o 'o ft TJ • t-i E 42 ft 0< TJ LO oT . — i bO 3 i — i o 01 bO rt rt 0) 3 01 o o o O Ol li O rt > 01 TJ 01 rt ■6 3 rt Ol 1 TJ «+h E E o OI o • r^ u. ■rt >- 43 42 _b0 u bO LO 0) 42 1_> ^-< CO 42 ■rt 0> 01 0) rt • f-i ^ r£_ bo rt CO TJ VS o o 0) CO Ol a rt 0) DO 01 ■rt TJ 42 ^_^ rt 3 CO 43 3 rt a 3 Oi rt rt TJ 0' P rt 0) > o ft rt QJ E Crt o ■rt 3 TJ O a 3_ S CO 01 <4H 01 i\ 4-> rt 4>J 3 0) TJ 43 E 3 z, 42 -rt a. Oi TJ bb oi E 81 corroborate the findings of Byrne, e_t al . (19 66). Although the amount of uplift is less than the maximum reported by Byrne, et al. , for the central Oregon margin (60 m to over 1000 m), it is within their lower range of values. Faunal data from the lower slope sample (6802-D3-1) suggest that a significant amount of down-faulting or subsidence has occurred on the lower margin. Movement and deformation on the margin was subsequent to either "early Pliocene" or "mid-Pliocene. " However, the three samples are all approximately the same age (Table 1). This suggests that both local uplift and subsidence could have occurred at about the same time, with the uplift restricted to the upper slope and shelf and the down-faulting or subsidence restricted to the deeper portions of the slope. Rates of Sediment Accumulation The stratigraphic horizons established at various depths within the unconsolidated sediment section permit the calculation of approxi- mate rates of deposition. By noting the depth interval between a par- ticular horizon and the surface (which is assumed to be at zero or present time) and by averaging this interval over the time represented, an estimate can be obtained of the sedimentation rates within the dif- ferent depositional environments (Table 2). The sedimentation rates determined for the upper slope, benches, 82 nl (X 1 (2 o rtt u c < 01 o o o as o c 42 N 4_J >H «T ° 0 * H o O X 42 2 2 JS9 rt o a. a Q w 0 2 o U DC i I I I _i a- J-. J-. |~-4 l-M M-< O ^O CM CM 00 CO (M o> o CTl 00 00 CM m Ol m CM IO en ro »* — 1 ro CM h- — < CM > in o ■* ro o ^ O O ^ o 4-1 ° o o IO 9 CD o IO " <"" < « ~ < cd Q a g tH Ou « Su« 2 c»~ C-- CM CM o o o o m o o o O O m o O O O o + 1 VO CM + 1 IO + 1 § I o in i o o o 1 o o o 1 1 0^ o cR o ro o o ^ o ^ o ^ o 8 G\ ^ ro ^f o CM 00 CM 00~ ro m to *•"■* CM ~^^ »— 1 — ~~* *■-'' 1) ' — ■ aj *"■'' m fj +-> +-> 4-» i-i +-> +-> "+H 42 •t-i 42 nl Q n) Q nf a in in Tf m o 0 o >~ >- c c ctf rt U U 01 n HI a 01 a u a. g> 3> u a 6 o o o o o f— 4 ^> 00 in PC PC in u< Eh i-t (-1 tH t-t r~- r^ K r^ IO IO ^O IO IO IC VO R IO o in o LO o IN. O o o s h> 8 LO O O m o O ~-< 00 m in lO CM CM in Tf ro CM ^H CM m ro CM CO ro •H ro ro 3 S ■>* CM o o o o IO S 2 u a o u a R in CM I 00 o 00 o 00 o in ^^ LO ^^ g r>- + 1 tv. O o o + o o o vn •" 1 IO i_> t£> cm" IO Oi CM o t-t IX u 4i a (ii a) c 00 i> O o o ^ o •& DC o 1) 4-» 42 n O 4J u w r> T) 4-) r. CO o U a o a u Pu 1) II •u nt DC S Pu U s hJ u a a* o tu o rt o | X X a n 42 p, a > nt >- 11 U nl a O • »H a T2- 0) (2 rt 43 O 83 and the Upper Rogue Canyon are moderate, averaging about 10 cm/ 1000 years (Figures 29 and 31). In contrast, the lower slope is experiencing the highest rates of sediment accumulation, on the order of 40 to 50 cm/1000 years (Figures 30 and 31). Sedimentation on the shelf varies widely from a very low to a moderately high rate and ranges from 5 to approximately 50 cm/1000 years (Table 2; Figure 28). These com- parisons assume that the surface of all cores represents present time, and that sedimentation continued uniformly at the calculated rate during the stated time interval. If the surface or near -surface sediment of the upper slope cores (670602, 6706-3, 6708-37) and the shelf core (6708-42) is assumed to be late Pleistocene in age as stratigraphic data suggest, then the calculated rates would have to be increased to about 15 cm/1000 years in each core (25 cm/1000 years in the case of core 6708-42) and would represent only late Pleistocene sedimenta- tion. Consequently, the Holocene sedimentation rate for these shelf and upper slope cores would be nil. In any case, the upper slope environments appear to be receiving less sediment than the lower slope. This is also demonstrated in the Rogue Canyon, where the rate cal- culated in the lower Canyon (core 6711-6) is eight times greater than that determined from the upper Canyon (core 6708-37). This suggests that the lower Canyon may be more closely related to the rapidly-filling swale (core 6708-38) than to the upper Canyon, and hence the lower Canyon is filling much faster than the upper Canyon. 84 Estimates of the Holocene sedimentation rate on continental slopes vary widely. Gorsline and Emery (1959) calculated that the green muds of the southern California Borderland province deposit at a rate equivalent to 72 cm/1000 years, while the estimates given by Moore (1966) for the total basin-fill in the same area range from 4 to 160 cm/ 1000 years. Carlson (1968) notes that the rate of deposition on the floor of Astoria Canyon on the northern Oregon margin ranges from about 50 cm/1000 years near the mouth to more than 75 cm/ 1000 years near the head, but he estimates a 10 cm/1000 years average for the hemipelagic deposition on the adjacent protected con- tinental slope. Duncan (1968) calculated postglacial (Holocene) rates of deposition ranging from 29 cm/1000 years to 100 cm/1000 years in the deep-sea area adjacent to the base of the continental slope. Thus the rates shown in Table 2, especially those from the lower slope, are not unreasonable when compared to the foregoing estimates. It is therefore suggested that the lower slope is presently accumulating sediment three to four times faster than the upper slope. 85 SEDIMENTOLOGY The sedimentological character of continental margin deposits can provide important clues to the provenance and dispersal paths of the sediments, as well as to the physiographic and tectonic framework of sedimentation in the margin environment. In turn, a knowledge of margin sedimentation patterns, particularly with regard to the role of the continental slope as a temporary resting place for sediments during their transport from shelves to final deposition in deep-sea basins, is an essential link in the interpretation of the complete cycle of oceanic sedimentation. Unconsolidated Sediments Classification and Distribution of Sediment Types Three main sediment types were observed in the cores taken on the southern Oregon margin. These are olive gray lutite, gray lutite, and sand-silt layers. Although not part of a particular classification, these terms have been used previously (Duncan, 1968; Griggs, 1969) to describe similar deposits, and are based primarily on character- istics observed by megascopic examination supplemented with the use of a binocular microscope. The term lutite (Ericson, e_t al_ . 1961; Heezen, e_t al . , 1966) implies a sediment composed chiefly of silt- 86 and clay-size particles, or mud. The proportion of the two con- stituents may vary over a wide range from a nearly pure clay to a sediment composed almost entirely of silt, but most often the lutite in this study is a clayey silt (< 50% clay) or silty clay (< 50% silt) according to Shepard's classification (1954). The distinction between olive gray lutite and gray lutite is based solely on color. Sand-silt layers refer to terrigenous sand or coarse silt-size particles com- posed chiefly of detrital mineral grains. At this point it is important to note that in most cases a physio- graphic province corresponds to an environment of deposition. There- fore, in subsequent discussions relating to the margin sediments, the term "physiographic province" will also imply the depositional environ- ment represented by that province, and conversely, the term "depositional environment" will also imply the particular physiographic province which includes that depositional environment. Any exception to this will be explained fully in subsequent discussion. Using the data from the textural analysis, and plotting each sample according to its sediment type, depositional environment and age, the sediments were grouped according to the textural classifica- tion of Shepard (1954) (Figure 26). Similarly, the coarse -fraction constituents of each sample are plotted on a triangular diagram whose three end-members are grouped according to origin and hydrodynamic character: the detrital group consists of those mineral grains and 87 K O UJZ a. u >-z t-2 l-° *l s s i, Bfc! 3 li I 5 f| Ui °° ill o IXJ © > * o3d o z CO w o §11 XI 0) 4-1 -4-> co o -tJ r-H c CL 0) , — . £ ^ ■—i in X o OJ —" CO ^ CD o en s c! • — i X o rt CD CJ a co a o D • rH D ■4-1 CJ ^4-1 -W 0 CO o 4-1 o ex CD X 3 CO XJ (tJ 3 CJ CD CO ^4-1 a XI CO C CO rti rt 0 r— 1 4-> CD U CJ DO re! I— 1 c <-M nJ ■ — I Jh u X d 4-1 o W X o 0) cj • H rt pq vO fM V U d DO h 88 rock fragments (except mica) requiring strong transporting currents; mica-plant fibers -volcanic glass are platy and/or low-density grains sensitive to current action; the biologic-authigenic group consists of material which accumulates in situ such as benthic and planktonic Foraminifera, Radiolaria, fecal pellets, glauconite and pyrite (Figure 27). The location and distribution of the major and minor sediment types within each core illustrated for the four main physio- graphic provinces: shelf (Figure 28), upper slope and benches (Figure 29), lower slope (Figure 30) and Rogue Canyon (Figure 31). Olive Gray Lutite Olive gray lutite is the most widespread and abundant sediment type (Figures 28-31). It varies slightly in color, from olive gray (5Y 3/2, Geological Society of America Rock Color Chart, 1963) to light olive gray (5 Y 4/2). The texture also varies, but it is typically a homogeneous, poorly sorted silty clay or clayey silt, with occasional coarse constituents (Figure 26). The olive gray lutite is character- istically poorly consolidated and relatively high in organic matter and moisture content. Although found in every environment, it especially dominates the lower slope (Figure 30), where the coarse fraction com- monly is composed of more than 50% biologic material (Figure 27). In contrast, the majority of the coarse fraction in samples from the upper slope and benches consists of detrital minerals (Figure 27). 89 V a CO fc»N to -i-> < -J o o -l-> 1 CO bfi (Z d ai ■ *-J CD T3 U. ^ O 1- O z 0 < rd _J 0. T3 1 V < j-> o 0 2 ~h CO a; ^ (0 a) .- o « to o a ■- cu "~ o m 'co CM u W) • ft Figure 28. Lithology of piston cores from the southern Oregon continental shelf. Sediment ages are indicated next to the core. Circled numbers indicate rates of sedimentation. Physiogra- phic province boundaries reflect present dis- sected surface of margin. Data for box core 6802-BCB after Chambers (1968). UPPER SLOPE AND BENCHES 5 O 5 10 HUH !— =] KILOMETERS LEGEND = ASH _ OLIVE GRAY " LUTITE = SAND-SILT = GRAY LUTITE 3 =FORAMS = GLAUCONITE ROCK FRAGMENTS Figure 29. Lithology of piston cores from the upper con- tinental slope and benches. Sediment ages are indicated next to the core. Circled numbers indicate rates of sedimentation. Physiographic province boundaries reflect present dissected surface of margin. Figure 30. Lithology of piston cores from the lower slope. See legend and caption of Figure 29 for explan- ation. Figure 31. Lithology of piston cores from the Rogue Sub- marine Canyon. See legend and caption of Figure 29 for explanation. 94 Mica, plant fibers and glass are rarely more than 20% of the coarse fraction in any sample. All of the olive gray lutite observed by Duncan (1968) in the adjacent deep sea was postglacial (Holocene); however, at least some of that sampled from the upper slope and Upper Rogue Canyon (cores 6706-2 and 6708-37, Figures 25, 29 and 30) may be late Pleistocene in age. In both cases gray lutite occurs at the base of the core. As Duncan (1968), Carlson (1968) and Nelson (1968) have observed, the change from gray lutite to olive gray lutite may mark a significant time boundary, the transition from late Pleistocene to postglacial (Holocene) time. If the R/P faunal ratio indicates that this transition took place about 12,500 years B. P. (Duncan, et al . , 1970) than the gray to olive transition should have occurred at about the same time and would be observed at approximately the same depth as the shift from abundant planktonic Foraminifera to abundant Radiolaria. A transitional type sediment, light olive gray, lutite often occurs between the gray and olive varieties in the deep sea, but appears to be restricted to postglacial time (Duncan, 1968). The color transition has been noted in four slope cores (6706-1 and 6706-2 from the upper slope; 6706-4 and 6708-37 from the Upper Rogue Canyon), with one other core (6706-3 from the Klamath Plateau) composed almost entirely of gray lutite (Figures 25, 29 and 31). Two of these cores (6706-1 and 6706-4) lack sufficient stratigraphic 95 information to confirm the above hypothesis. However, a core adjacent to each of these from a similar environment (6706-2 and 6708-37, Figures 29 and 31, respectively) clearly indicates that the color shift occurs earlier than 12,500 years B. P. (Figure 26). This indicates that either the deposition of olive gray lutite began much earlier in the upper margin environments than in the deep sea and hence may be as old as late Pleistocene, or that the olive gray lutite in cores 6708-37 and 6706-2 should have been classed as gray lutite, a distinction not apparent from either megascopic examination or subsequent analyses (Figures 26 and 27). Gray Lutite Gray lutite commonly occurs as a clayey silt or silty clay (Figure 27) and varies in color from olive gray (5Y 4/1) to dark greenish gray (5GY 4/1) or medium dark gray (N4). It is generally cohesive and stiffly compacted, and hence tends to be better con- solidated than the olive gray lutite. The coarse fraction content of the gray lutite comprise only two or three percent of the total sedi- ment, and consist mainly of detrital grains with only 25 to 40% biologic material (Figure 27). As noted above, gray lutite was penetrated only in cores from the upper slope and Upper Rogue Can- yon. If it is indicative of the late Pleistocene, it probably underlies the thicker accumulation of Holocene sediment on the lower slope. 96 In continental margin environments, gray lutite cannot be distinguished from olive gray lutite on the basis of texture (Figure 26) or coarse fraction (Figure 27). It may be distinctive initially only in its color until its age can be established. Gray lutite from the continental mar- gin is similar texturally to that found by Duncan (19 68) in adjacent deep-sea environments, and closely resembles the gray clay found by Griggs (1969) in Cascadia Channel. However, the coarse fraction in these latter types often contains 75% or more biologic material, principally planktonic Foraminifera and Radiolaria, as compared to the lesser amounts found in gray lutite from the margin. This reflects the increased pelagic rain of shells and tests and the smaller amount of detrital material found in deeper and more distant environments. Sand-Silt Layers Terrigenous sand and coarse silt occur in all environments and in all cores except 6711-1 (Figures 28-31). It is the dominant sediment type in cores from the continental shelf (including 6706-7 from the head of the Rogue Canyon) and from the Cape Blanco Bench. It occurs in these latter environments in thick sequences, and in all other environments as distinct layers with sharp and well-defined contacts interbedded with olive gray lutite or gray lutite. X-radiographs revealed that all of the sands sampled were essentially featureless and structureless with no grading observed within a unit or layer except for some minor 97 grading noted in core 6706-5 from the Upper Rogue Canyon. Texturally, this type varies from sandy silt to sand, but still can easily be dis- tinguished from the lutites (Figure 26). Generally the coarse fraction consists of more than 75% detritals. The age of the sand-silt layers in the various environments is more difficult to determine, and can be either Holocene or late Pleistocene. Coarse fraction analyses reveal an abundance of iron- staining, solution pitting, and alteration in the sand-silt materials from the surface of both the shelf and Cape Blanco Bench, but these characteristics are absent from other parts of the slope. The thick accumulations of sand-silt on Cape Blanco Bench are more suggestive of shelf sands than of the thinner sand-silt layers on the lower slope. Chambers (1968) recognized staining and alteration in sands from the southern Oregon shelf; Emery (1965, 1968) notes that staining and alteration of shelf sands may be indicative of their relict nature. It is postulated here that the sand-silt on the surface of Cape Blanco Bench is also of a relict nature, and may be late Pleistocene in age. However, all of the sand-silt layers in the cores from the lower slope are Holocene. Minor Sediment Types Several minor sediment types were observed which were dis- tinctly different from the three main types just described, but were not 98 widespread enough either vertically or areally to warrant their clas- sification as major sediment types. These types usually consist of sediments whose coarse fractions contain appreciable amounts of distinctive constituents. The most important of these were: Foraminifera-rich lutite (6706-2 from the upper slope; 6711-1 and 6711-2 from the lower slope); volcanic ash-rich layers, usually lutite (6708-23 and 6708-25 from the shelf; 6706-6 from the wall of the Upper Rogue Canyon; 6711-1, 6711-2, and 6708-38 from the lower slope); shell layers (6708-42 from the shelf); rock fragments (6706-3 from the Klamath Plateau); and glauconitic layers, usually sandy (6706-1 from Middle Bench, 6706-3 from the Klamath Plateau, and 6706-4 from the Upper Rogue Canyon). The glauconitic layers occur at the surface of two cores from the upper slope close to the relict surface sediments. Glauconite is normally authigenic, or formed in place, and is often relict (Emery, 1968). Because of its association with relict, or possibly relict sediments, it is postulated that the glauconite found in the upper slope cores is also relict (i. e. a sediment formed out of equilibrium with existing conditions). The tentative relict age of the olive gray lutite at the surface of core 6706-2 may also apply to the adjacent olive gray lutite and surface glauconite in core 6706-1. Similarly, the late 99 Pleistocene age of the gray lutite in core 6706-3 (Figure 25) also suggests that the surface glauconite in this core is not modern but relict. The coarse sand in core 6706-7 from the head of the Upper Rogue Canyon, the late Pleistocene age of the olive gray lutite in core 6708-37, and the surface glauconite in core 6706-4 (Figure 31), all suggest that the Upper Rogue Canyon may be another relict area on the upper slope. Without additional evidence for the absolute ages of the surface sediment in cores 6706-1, 6706-3, and 6706-4, it is difficult to state with certainty that these sediments are late Pleistocene in age, although the existing evidence suggests that they may be. Relict sedi- ments need not all be of late Pleistocene age, they could have formed early in Holocene time under different conditions than those encountered today. The Present Sediment Pattern The stratigraphic and sedimentological evidence from the upper slope environments, especially the benches, strongly suggests that the surface sediment in these areas is both modern and relict in nature. Certain areas, such as Cape Blanco Bench and the Klamath Plateau, may be covered entirely with relict sediment while other upper slope areas contain patches of both modern and relict material. The data of Chambers (1968) suggest that the continental shelf off southern 100 Oregon is also surfaced with both modern and relict sediment. In contrast, the lower slope appears to be completely mantled by modern mud. By combining the data available on the surface sediments of the southern Oregon margin, a generalized genetic classification emerges. Figure 32 illustrates six types thought to be present on the margin. Modern sand occurs on the inner shelf and modern mud is found on both the central shelf and lower slope. Mixed sand and mud and the relict sand found on the shelf have their analogs in the mixed modern and relict deposits and the relict sand found on the upper slope. The shelf, upper slope, and lower slope are distinct physiographically and as sedimentary environments. The upper slope appears to be more closely related to the shelf than to the lower slope, and could be thought of as a transition zone between the two, not unlike the sedi- mentary regime of the Borderland province off southern California (Gorsline and Emery, 1959; Moore, 1966). Textural Relationships Certain statistical measures related to the texture of the sedi- ment have been examined in an attempt to test whether or not the three environments (shelf, upper slope, lower slope) can be dis- tinguished more precisely. Mean diameter (M ) has been plotted 9 against both the phi deviation (cr ) and phi skewness (a ) with respect 9 9 Figure 3 2. Distribution of surface sediment types on the southern Oregon margin. 102 to environment (Figures 33, A and 34) and sediment type (Figures 33, B and 35). The sorting of the majority of sediments, particularly the olive gray lutites from the lower slope, is generally poor (Figure 33). No sharp demarcation can be made between sand-silt layers and lutites on this basis, either environmentally or by sediment type. A division of the sediments into two groups is apparent in both Figures 34 and 35. One group represents sand-silt layers, which are coarse-grained and extremely fine -skewed, and both olive gray and gray lutite appear as another group which is fine-grained and coarse -skewed. The coarse (negative) skewness of the lutite may be attributable to the presence of sand-sized tests or shells of Foraminifera or other pelagic organisms. No environmental distinction can be made among any of the three sediment types which would indicate their particular environment of deposition. Conversely, it also appears that no clear textural subdivisions can be made among the various environments which would indicate the presence of a distinct sediment type. Mineralogy An analysis of the mineralogy of the margin sediments can be most useful, not only in distinguishing among the various sediment types, but also in determining the provenance of the sediment and its dispersal patterns. 103 9r o SHELF • UPPER SLOPE □ BENCH a UPPER ROGUE CANYON g 6 _ ♦ LOWER SLOPE ^ L±J Q X 4 g 3 X 2 A "♦ O 5- °H? ♦ ♦.♦♦♦♦# e ♦♦♦♦!♦ r* > 4 " A _ n „oA°° „ ♦. ♦ I* H 31" ° o un o cu, aao . ° oo', □ »t ' ■ A *ft,°fat*2tA o o ogcfitfS o 2 - TO CTrtn O J lDo° I I 2 3 4 5 6 7 8 9 10 II 12 PHI MEAN DIAMETER (M0) 81- • OLIVE GRAY LUTITE nGRAY LUTITE o SAND- SILT b •• • • 0 ••••••• > o00 • • *<*»• B Q- o°o0o°oo8 I ■ I0' I I I I I I I 1 1 II 2 3 4 5 6 7 8 9 10 II 12 PHI MEAN DIAMETER (M0) Figure 33. Phi Mean Diameter versus Phi Deviation according to (A) depositional environment and (B) sediment type. .9 .8 .7 .6 .5 .4 Si 3 3 °-v ■*> 00 3UJ IT g O 8/ \ o I . CO 1 o I ^ fe S I- e^ / •i ,, / •n < SQ . o fO t /feld- /SPATHI Vl SAND \ \ o OD !=? 5 g Q 2 r-. N 1 <■£> «> 1 UJ < CO * i 0 R SLOPE R ROGU ANYON Q- o -1 to c 3 . . U u CO a 0 i— I PQ 0 ^ co >, *H Jd u CD CJ rt -UJ o 4J (ti h c CD co cj o CD a 4-> 0 GO CD Sh CO Sc. U d o rgs a a 5-* R i—i CD CO a a o O O U a o ^^ 0 -UJ ■* -UJ rt Lfl CO o ■ — < o> o «+H V CL a CO CO 4-1 o nJ CD o ■— ( X! u i—i ^ 1—1 nJ CJ 0 (D X! T3 -^ CL, c! ID (TJ rt no a GO ^ 2: -UJ O CD -~^ X! U s ■» tuO -t-> ?H 4-J • m * d c> c> O O co ro -* ro to ro "tf ro O fl (M 00 Ol ro LO "* ^ ro ■* M m m h -h CM c CU o U a (U B a c o M [-. 74 '£ 2 rt ai CO a. •* -H -H CM (M in i cm in h n N m o cm oo CO CM CM a o Tl_^ 4-> CO rt -tJ >» 0 CO a rt ro U rt 0 ^ U CO* *"* J= u CU >- 4-» 1 > ^> CU cu 3 n 0) 3 O •r-l +■> c5 rt ^H 0 bo o Eh o o a rt rt s rt rt a E O U t: 0 £ "o -t-» CO CQ D D _J PQ 2 D z u .— < rt U O t-> •^H -J rt c *t B cu CU 60 c rt g 8 'c( o t- U Q rt O u 1 o CO H ,_i rt rt u* 2 cr> ■-H CO * 'H 3 ~H CO CT, CO rt o cu £ ~ CU C CU .s c bO cu 13 rt Q • rn -a a rt CU T3 co rj CU .£5 p. CO CU 3 cr rt a o i a o a >s -a o 3 CU 0 cu rt p. •s .B 3 rt G. O E -c ^ So >s -a o oo CO CTl CO CU "&, E rt cu P c rt cu J3 T3 Dh C o 0) o 3 « rt E be o ,C 3 rt X) cu O T3 ■a d B J3 E o I ^ r2 cd U cu n-i bC j2 113 The relative abundance of pyroxene and amphibole is a consistent indicator of the source area contributing sediment to the margin environments. The minerals included in the pyroxene to amphibole ratio (P/H ratio) constitute the majority of the non-opaque, non- micaceous and unweathered constituents and therefore adequately represent the entire heavy mineral suite. The average P/A ratio in the margin environments is similar to that obtained for the Klamath- South Coast Basins (Table 4). The P/A ratios indicate that the sediments in all margin environ' ments are influenced more by the mineral assemblages in the Klamath South Coast Basins than those from any other basin. Despite the fact that the river runoff of the Columbia Basin is over seven times that for the Klamath-South Coast Basin (Figure 37), the distinctive assemblage of the latter basin still appears to dominate the southern Oregon margin environments. It may be argued from the data of Table 4 that the lower slope, which has a relatively high P/A ratio of 0. 7, is also influenced by sediments from the Umpqua Mid-Coast Basins and/or the Columbia River. However, the clay mineralogy suggests that only the sediments from the Umpqua Mid-Coast Basins have any noticeable effect on the lower slope off southern Oregon. In general, surface sediments from the benches and upper slope exhibit higher percentages of hematite and limonite in their heavy mineral fractions than do sediments of other environments. 114 Considerable iron staining, pitting, and alteration is also evident in the light mineral fractions from these environments, especially from the benches. As previously noted, this evidence may be an indication of the relict nature of the benches sediments, and to a lesser extent of most upper slope sediments. Shelf sediments show a higher average percentage of heavy minerals, particularly opaque grains, than do the other environments. Chambers (1968) has plotted the zones of high heavy mineral concentration on the shelf; he has also noted that opaque grains may be indicative of stillstands of sea level due to their high specific gravity and resistance to offshore transport. The shelf samples examined in this study are within, or adjacent to, high heavy mineral zones mapped by Chambers and the results from these samples appear to be consistent with his data. Clay Minerals The clay mineralogy reinforces the hypothesis that the Klamath- South Coast Basins are the most important source for the sediments of the southern Oregon margin. The clay fraction from 18 sediment samples representing shelf, upper slope and lower slope environments, and from three rock dredges was examined to determine the per- centages of the various clay minerals present (Figure 38). The clay minerals from all the environments on the southern Oregon margin consist predominantly of chlorite and illite, with only minor amounts 115 LU LlI co Z o o > 2 LU < h O < o UJ UJ UJ -I o a. o 3 o o 0- O CO t/> a: CO z o <_> U. (r I cr cr _l LU o LU LU LU a z a. * z I a LU a. o o CO o CD 3 -J ro cz* O i ro C\J i O (0 , — CD O to CO f>- co o ' £ Ld CVJ I o ■z. _l o z> DL 00 LU < z 5 CD CD LU 1- CO < LU O < < CO a O _l CD O z a. _J < CC UJ CO _l i LU (9 Q UJ or LU CD LU < 5 > cr 5 a < cr o < _i a _l o it ® cu a) G co •M CO £ cd O CJ S o ° u -C > ' H CO Q £ ^ .1-1 d o T3 CD co T3 O 0) ^ CJ <^ a, -o co —I O * co fl G O -^ i^ co :>> ^ e u * O o3 O • TJ co >H £ 00 .^ -—I 5 a — c o U O "SI oo ro CD U ofl 116 of montmorillonite. No kaolinite was found. The presence of amphiboles of clay size was also noted in six of the sediment samples and all the rock samples examined. In all sediments and sedimentary- rocks, chlorite is the most abundant mineral (average 60%), with lesser amounts of illite (average 35%). Only minor amounts of mont- morillonite occur in the sediments (average 5%). The Rogue River, which drains the metamorphic terrane of the Klamath Mountains, is noted for its high chlorite content. Duncan (1968) and Duncan, et al. (1970) examined the clays from this river and obtained values of 5 1 /o chlorite, 26% illite, and 23% montmoril- lonite; Drake (1969) obtained similar values in her analysis (43% chlorite, 38% illite, and 19% montmorillonite ) (Figure 38). When compared to the clay mineralogy of the Umpqua River of any of the Columbia River sub-basins (Figure 38), the Rogue River appears to be the dominant influence affecting the clay mineralogy of all sediments on the southern Oregon margin. Other rivers draining the Klamath Mountains may also be chlorite -rich and may contribute to the relative uniformity of the clay mineral content of margin sediments. The fact that all the samples examined are so nearly alike in clay mineral com- position attests to the pervasive influence of chlorite -rich sediments from the Klamath Mountains, which masks any influence from the Columbia sub-basins, and to a lesser extent, from the Umpqua Mid-Coast Basins. 117 Russell (1967), Duncan (1968), Duncan, et al . (1970), and Harlett (1969) have all noted that diffractograms of late Pleistocene deep-sea clays exhibit sharper peaks than those of Holocene deep- sea clays. Duncan et al. (1970) report that more illite than chlorite is present in late Pleistocene deep-sea clays than in the Holocene clays. They postulate that the change in character between the late Pleistocene and Holocene clays is probably due to variations in the relative rates at which the Columbia River and Snake River sub- basins contributed their clay mineral assemblage to the sediment load of the Columbia River. They state further that the Upper Columbia River sub-basin was a major source for the illite-rich deep-sea sediments during the Pleistocene, but as the region emerged from glaciation montmorillonite -rich sediments of the Lower Columbia and Snake River sub-basins began to dominate the sediment load. This change is abrupt, and not due to more gradual marine diagenetic processes; it is associated with the late Pleistocene -Holocene faunal boundary (ca. 12, 500 years B. P. ) and has been verified quantitatively by similar abrupt increases in the chlorite -illite and montmorillonite - illite ratios (Duncan, 1968; Duncan, et al. , 1970). Chlorite -illite ratios were calculated for the samples examined in this study; the samples represent the tops and bottoms of nine cores from the various environments (Appendix 7). Although the Pleistocene-Holocene boundary was not evident in the cores examined the chlorite -illite ratio 118 did increase upward in all but two cores, a trend which is in general agreement with Duncan's data. In addition to determining the change in clay mineral character with depth, Duncan (1968) noted that three Holocene clay mineral groups are evident in the deep-sea which radiate outward from the Columbia River. The montmorillonite content decreases while chlorite, and to a lesser extent, illite, increase with distance from the river. Group three, in the deep sea off central and southern Oregon, contains the most chlorite (mean value 40%) and the least montmorillonite (mean value 32%). The clay minerals from the continental margin appear to belong to this group. However, the higher chlorite content of margin sediments compared to those in group three suggest either that a significant amount of chlorite is retained on the margin and may never reach the deep-sea environment, or that the chlorite is diluted as the clays move seaward. The occurrence of amphibole in the clay mineral fraction from three cores and three dredge samples (Appendix 7), representing mostly upper slope environments, may be significant. Heath (1969) suggests that the amphibole may have been produced as glacial rock flour during the late Pleistocene and may have deposited on the mar- gin in that form. However, except for a few valley glaciers that may have been present, the region was not glaciated during the Pleistocene. In addition, the sediment in these samples is not all Pleistocene in 119 age. A fuller explanation for the presence of amphibole must await a more detailed examination. Organic Carbon Sedimentation rates in late Pleistocene was at least six times higher than that in the Holocene in the deep-sea environments off Oregon (Duncan, 1968). Duncan found that the organic carbon content of postglacial lutite is up to five times higher than that of the late Pleistocene lutites. He attributed this apparently reversed trend to the higher influx of sediments in the late Pleistocene masking the organic carbon content of these older sediments. Peterson (1969) reached the same conclusion by calculating sediment and organic car- 2 bon accumulations in terms of gm/cm /1000 years. From his cal- culations, Peterson found that although the sedimentation rate and organic carbon influx was higher during the late Pleistocene, the absolute amount of organic carbon preserved in the older sediments was less due to dilution, differences in preservation, and depth of burial. In this study, samples from near the top and bottom of eight cores representing different depositional environments were analyzed for their percent total carbon, organic carbon, and calcium carbonate (Table 5). Two additional cores were more fully examined by Peterson (1969). 120 Table 5. Total carbon, organic carbon and calcium carbonate in selected margin sediments. Percent Percent Sample Depth in Depositional . b Age Sediment total organic Percent number core (cm) environment typec carbon carbon CaC03a 6708-25-4 230 Shelf H OGL 1. 16 0.90 2. 17 6708-25-16 583 H OGL 0. 95 0.63 2.67 6708-42-1 230 Shelf H S-S 0. 53 0.44 0.76 6708-42-2 454 LP S-S 1.02 0. 54 3.97 d 6706-5-(l) 6706-5-(2) 17 183 Upper Rogue Canyon LP LP OGL OGL 1. 41 0.69 1.36 0. 47 0.48 1. 77 6706-5-(3) 255 LP OGL 0.70 0. 53 1.35 6708-37-1 15 Upper Rogue LP OGL 1. 70 1.57 1.03 6708-37-7 575 Canyon LP GL 1. 56 1. 14 3.50 6706-3-1 5 Klamath LP GL 1.00 0. 95 0.41 6706-3-10 377 Plateau LT GL 0.84 0.67 1. 44 6706-2-12 302 Middle LP OGL 1.34 0.85 4.08 6706-2-17 375 Bench LP GL 1. 18 0.74 3.66 6706-6-8 155 Upper H OGL 1. 14 0. 94 1.53 6706-6-15 355+ Slope H OGL 1. 18 0. 89 2. 47 6711-6-1 10 Lower Rogue H OGL 2.30 2.00 2.45 6711-6-8 340+ Canyon H OGL 2.22 1.84 3.00 6708-38-1 5 Lower H OGL 1.82 1.60 1.82 6708-38-21 472 Slope H OGL 1. 52 1.32 1.62 d 6711-2-(l) 19 Lower H OGL 2.06 1.42 5. 32 6711-2-(5) 154 Slope H OGL 1.06 0.97 0.62 6711-2-(10) 410 H OGL 1.33 1.20 1. 16 Percent by weight LP -late Pleistocene, H - Holocene, LT - Late Tertiary. OGL - olive gray lutite, GL - gray lutite, S-S - sand-silt layers. i Data and sample number designations after Peterson (1969) 121 In those margin cores where the surface, or near-surface, samples were examined, the average percent organic carbon was found to be 1. 3, with surface samples from the lower slope being among the highest values recorded (1. 6% and 2. 0%). These values compare favorably with the 1. 6% median value from the adjacent con- tinental slope at 41059'N, reported by Gross, et al. (1970). Sediments accumulating on the continental slope are thought to have high organic - carbon concentrations where the oxygen-minimum zone impinges on the bottom. Off Oregon the oxygen minimum zone extends to depths of 500 to 2000 m and appears to coincide with the high organic carbon content of slope sediments in this depth range. All but one core (6708-42, a sandy core from the shelf) showed an increasing percentage of organic carbon upwards, and all but three cores (6711-2, 6708-38, and 6706-2) showed a decreasing percentage of calcium carbonate upwards in the core. In terms of absolute values of organic carbon, samples from cores considered to be entirely Holocene in age showed generally higher values of organic carbon and lower values of calcium carbonate than cores considered to be entirely late Pleistocene in age. It is not possible to determine whether this increase is abrupt. Although the Holocene sedimentation rates com- puted in this study are high (Table 2), it is possible, as Duncan has noted, that Pleistocene rates could have been several times higher both on the margin and in the deep sea. If this hypothesis is correct, 122 then the decrease in sedimentation rate from Pleistocene to Holocene time may not have occurred as abruptly on the margin as in the deep sea. Consolidated Sediments Classification and Distribution of Rock Types A comparison of the consolidated rocks recovered from the dredge hauls with the unconsolidated sediments from the cores reveals certain trends which may provide a clearer understanding of the geological events that have occurred on the margin. Figure 39 shows the location and major lithology of the rock types collected in the ten dredge hauls. Samples from nine of the rock types were sufficiently coarse-grained to permit a detailed petrographic examination (Appendix 8); these have been grouped according to the classification of Williams, Turner and Gilbert (1954) and are shown in Figure 36, C. The remaining samples were too fine-grained for petrographic identi- fication (Figure 39; Appendix 8). The lithified samples fall into two groups: coarse-grained impure sandstones, mostly arkosic and lithic wacke (Figure 36, C), all but one of which are from the shelf and upper slope, and fine- grained mudstones and siltstones from the lower slope (Figure 39), some of which are calcareous. When one compares the light mineral ARKOSE/ARKOSIC WACKE V LITHIC WACKE FELDSPATHIC V_^\ WACKE ^m ^ .-SHELLS DENOTES LENGTH AND DIRECTION OF DREDGE HAUL Figure 39. Rock types from the southern Oregon margin Numbers next to circles denote percent of each rock type. 124 composition of the unconsolidated sediments (Figure 36, A) with that of the consolidated sediments (Figure 36, C), it is apparent that the unconsolidated sediments have a higher percentage of unstable grains than do the lithified sediments. This would seem reasonable since the sediments lose much of their unstable elements during the various stages of diagenesis and repeated cycling as they become lithified. In general, both the unconsolidated and consolidated sediments are arkosic in character. A comparison of the light mineralogy of the consolidated sediments on the margin with that of southwestern Oregon continental rocks examined by Dott (1965) (Figure 36, B) shows that the composition of the shelf and upper slope sedimentary rocks are similar to those of the Upper Cretaceous Series and the Dothan Formation. This suggests that the late Tertiary rocks of the shelf and upper slope could have been derived from either the Jurassic or Late Cretaceous rocks of the Klamath Mountains. A complex of Tertiary and pre-Tertiary strata crop out in the coastal and near-coastal regions of southwestern Oregon. These have been described in detail previously (see Regional Geology). The early Tertiary strata exposed along the southern Oregon Coast Range were folded and faulted during the late Tertiary Cascadan orogeny (Dott, 1965), and the resulting north-south-trending anticlines and synclines probably extend westward from the coast offshore and underlie the shelf and upper slope, and perhaps the lower slope. All but the most 125 resistant Tertiary strata probably occur only in synclinal basins or down-faulted blocks, and most likely only the younger Mio-Pliocene strata are widespread, or exposed, on the upper margin. Maloney (1965) reports only Miocene and Pliocene rocks from the shelf, banks, and slope off central Oregon. The age of three rocks recovered from the margin off southern Oregon date back to the early or mid- Pliocene (Table 1). Most, if not all, of the predominantly arkosic rocks from the shelf and upper slope (6708-22, 6708-36, 6708-41, 6711-D1; Figure 39) may also be Mio-Pliocene in age, although it cannot be stated with certainty that they are from contiguous strata. The calcareous, nodular mudstones (6711-D2 and 6711-D3) may be widespread on the upper slope, particularly on the Klamath Plateau, since similar rocks have been recovered by Silver (1969a) on the Klamath Plateau off northern California. The rocks are probably concretionary in origin, but the time of their formation is unknown. Rock fragments in the base of core 6706-3 (Figure 39) have been dated as mid-Pliocene (Table 1). This suggests that a distinct unconformity, probably a result of tectonism, exists between the Tertiary rocks at the base of the core and the overlying late Pleistocene sediments. In general, the lithology and age of the rocks recovered by Silver (1969a) from the northern California margin are similar to those found in this study. Clay mineral contents of the three rock samples from the slope 126 (6802-D3-1, 6802-D2, and 6706-3-11-1; Figure 38) show the same high chlorite percentages as unconsolidated sediments from the slope, which indicates that the contribution of the chlorite-rich sediments from the Klamath Mountains has been strong, at least since the Mio- Pliocene. The large difference in the chlorite-illite ratio in core 6706-3 between the mudstone fragments at the base (C/I = 3. 9) and the immediately overlying gray lutite (C/I =1.5) reinforces the evidence of a distinct erosional boundary at that depth in the core. In general, the grain-to-matrix ratio of the consolidated sedi- ments from the margin approximates the sand-mud ratio of the over- lying sediments. On the shelf, the grain-matrix ratio of the rocks approaches one and the sand-mud ratio of shelf sediments is also commonly one or slightly greater. In the remaining margin environ- ments, both the grain-matrix ratio of the rocks and the sand-mud ratio of the overlying sediments are usually less than one, reflecting an absence of coarse constituents in these environments. In examining the composition of the consolidated and unconsoli- dated sediments of the southern Oregon margin, certain generaliza- tions can be made. First, the light mineral composition of the sedi- ment samples is generally similar to the rocks, but the latter have more stable constituents. Second, the upper slope sediments and rocks have a composition more closely approaching that of the Upper Cretaceous Series and Dothan Formation of the Klamath Mountains 127 than that of the Eocene or other early Tertiary strata of the southern Oregon Coast Range. Third, the lower slope sediments and rocks have a higher percentage of unstable constituents, particularly feldspars, than do those of the upper slope. They tend to reflect the influence of the early Tertiary strata of the southern Oregon Coast Range as well as that of the Klamath Mountains. These relationships suggest an orderly sequence of events. The Mio-Pliocene strata exposed or underlying the shelf and upper slope were apparently derived from the uplift and erosion of pre- Tertiary Klamath Mountain strata, probably during the Nevadan orogeny or late Cretaceous disturbance. The Quaternary sediments exposed or underlying the lower slope were probably derived in part from a later uplift and erosion of mid- and late Tertiary sediments, probably during the Cascadan orogeny, and in part from the pre- existing Klamath Mountain strata. 128 PROCESSES OF SEDIMENTATION A Proposed Model for Modern Sediment Transport on the Southern Oregon Continental Margin Many factors influence sedimentation on the southern Oregon con- tinental margin. These include the character of the source rocks, climatic effects, the quantity of sediment supplied by the continental drainage, the influence of coastal morphology and submarine topography, and the various oceanographic conditions which prevail over the margin. A generalized model of modern sediment transport processes (Figure 40) is proposed which attempts to relate these factors to the sediment distribution pattern presently found on the margin, and to that present during the late Pleistocene and early Holocene. The Initial Regime of Sedimentation The major sources of the unconsolidated sediments on the southern Oregon margin lies in the adjacent coastal complex of pre- Tertiary and Tertiary rocks of the Klamath Mountains and southern Oregon Coast Range. The relative dominance of sediments from the metamorphic terrane of the Klamath Mountains throughout the entire margin has already been shown. To a lesser extent, sediments from the Tertiary rocks of the southern Oregon Coast Range are contributed to the margin, but this contribution is more noticeable on the lower FLUVIAL TRANSPORT LONGSHORE DRIFT LUTUM TRANSPORT SUSPENSATE BOTTOM TURBID w LAYER SLUMPING Figure 40. Schematic model of modern sediment transport processes on the southern Oregon margin. 130 slope than in other margin environments. The major drainage complex supplying sediment to the margin is that of the Klamath-South Coast Basins. Additional sediment may be added to the southern Oregon margin by the major streams of the Umpqua Mid-Coast Basins, particularly by the Umpqua River, which drains portions of the northern edge of the Klamath Mountains as well as the southern Oregon Coast Range. River runoff data for these streams has been compiled by Hagenstein, et al. (1966) and by the U. S. Geological Survey (1964, 1966). Over 87 % (21 . 5 million acre feet) of the total runoff of the Klamath-South Coast Basins comes from the Klamath and Rogue Rivers; one-half (7. 5 million acre feet) of the total runoff of the Umpqua Mid-Coast Basins is supplied by the Umpqua River (Figure 37). The runoff of all the streams in both basins is only 20-25% of that supplied by the Columbia River (Lockett, 1965). How- ever, mineralogical analyses in this study have shown that the influence of the Columbia River, and that of the North Coast Basins to the south of it, have a relatively minor affect on the sediments of the southern Oregon margin. The amount of rainfall over southwestern Oregon varies widely with seasons, but is greatest during winter months, which is reflected in the stream runoff. For example, although the average discharge at the mouth of the Rogue River is approximately 3000 to 4000 cubic feet per second (cfs) for the entire year, it reaches a maximum of 15, 000 131 to 16, 000 cfs in January, and a minimum of 1500 cfs in September, following the low summer rainfall season. Since the amount and size of sediment discharge from a river is a function of the rate of flow of the river (Postma, 1967), the Rogue River discharges more sediment in winter, by an order of magnitude, than in summer. The last rise of sea level has created numerous bays and estuaries at the mouth of the major streams along the northern Califor- nia and southern Oregon coasts. These features have a marked influence on the supply of sediment to the southern Oregon margin. Fluctuating tides influence the interaction of the discharging fresh river water and the encroaching oceanic salt water within these bays and estuaries. According to Postma (1967) various accumulation mech- anisms, particularly settling- and scour-lag effects, tend to trap sediment inside these basins or at leas-t restrict their movement to the near-shore regions. Kulm and Byrne (1966) have shown this to be the case in Yaquina Bay, Oregon. Postma noted that this sediment trapping is most effective for relatively fine-grained sediment, but may vary between wide limits depending on local conditions to any size between five to ten \i and 100 u. As Chambers (1968) has noted, the Holocene sea level rise has caused the trapping and restriction of sand-sized sediment on the con- tinental shelf off the Rogue River. A belt of sand is present from the shoreline to a depth of about 50-75 m, but only sand shoaler than 132 15-20 m (modern sand) is considered to be in equilibrium with the present environment (Figure 32). The sand between 20 and 70 m is probably relict sand of Pleistocene age. Dietz (1963) and Vernon (1966) conclude that modern beach sand is generally not transported offshore to depths greater than 20 m, except where the heads of submarine can- yon extend into this depth. No work later than that of Chambers, including the present study, has detected any appreciable quantities of modern Holocene sand at the surface in margin environments deeper than the inner shelf, although Roush (1970) and Neudeck (1970) have cited evidence from sedimentary structures and from oceanographic conditions which indicate that sand movements to depths of 80-90 m or deeper may be possible. The Holocene sea-level rise on the southern Oregon margin pro- duced the following: 1. large accumulations of relict Pleistocene sand on the submerged benches and the deposition of large quantities of Holocene mud on the lower slope; 2. a six-times greater Pleistocene sedimentation rate in the deep sea compared to the Holocene (Duncan, 1968); 3. non-deposition of Holocene and modern sands on much of the outer shelf due to trapping in the bays and estuaries. This evidence all points to a large influx of coarse-grained sediment on the margin during the late Pleistocene when sea level was lower, followed by a greatly reduced and finer-grained influx of sediment deposited in Holocene time as a consequence of rising sea level. 133 The Concept of Lutum Transport The initial premise upon which the present sediment transport model (Figure 40) is based is that since the modern time (3000 years B. P. to the present, when sea level is assumed to have been at about the same level it is today), and indeed since the beginning of Holocene time (12,500 years B„ P. ), the sediment influx to the margin has been primarily one of fine-grained silts and clays, or lutum. Lutum here refers to fine-grained sediment still in transport or suspension, as distinguished from lutite, which is fine-grained sediment that has been deposited. As opposed to the late Pleistocene, when great quantities of coarse sediment blanketed the margin and deep sea, very little modern sand reaches depths greater than 20-30 m. As this study has shown, Holocene (including modern) lutites are ubiquitous on the mar- gin, but the thickest accumulations are found on the lower slope. These lutites are generally uniform in character with no visible grading, as are the sand-silt layers interbedded with them. These sand-silt layers were probably deposited by mass downslope gravity movements, such as slumping; it is also possible that they are winnowed deposits, or even the finer-grained tails of unrecognizable turbidites. However, the evidence precludes a turbidity current origin for much, if not all, of the Holocene deposits, and suggests instead that a process of large- scale lutum transport must be called for to account for the Holocene 134 and modern deposits. Such a process has been described by Moore (1970) to account for much of the basin-filling occurring on the southern California Borderland. A similar process is postulated to occur on the southern Oregon margin. The Oceanographic Regime In order to fully understand lutum transport on the southern Oregon margin, it is also necessary to understand the oceanographic regime in which it operates. The oceanographic factors play a major role in determining the routes of sediment transport and dispersal, and consequently the ultimate sites of deposition. Many of the oceanographic factors that are important to sediment transport vary seasonally and shift in direction in response to the changing wind pat- terns. Sea and swell and many of the currents on the margin change direction in this manner; therefore, it is important to distinguish be- tween the oceanographic regime in the winter (Figure 41, left) and that which exists in the summer (Figure 41, right). The Westwind Drift is a major ocean surface current which moves eastward across the North Pacific Ocean. As it approaches North America at a point about 500 km west of the continent near 45° N latitude it divides into two parts: one part of the Drift flows north and one part south as the California Current (Dodimead, Favorite, and Hirano, 19 63). The Davidson Current flows northward as a subsurface Figure 41. Schematic diagram of the major ocean currents on the southern Oregon margii Left, represents winter conditions; right, summer conditions. 136 current on the coastal side of the California Current (Figure 41) and breaks through to the surface during the winter months (Sverdrup, et al_. , 1942); it has been reported as far north as 50° during January and February (Burt and Wyatt, 1964). Stevenson, et al. (1969) report a subsurface current flowing about 70 km offshore from Oregon which has a southerly component of movement from the surface to 500 m during the entire year except for some slight northward movement during the winter months (Figure 41, left). They measured velocities averaging 5-10 cm/sec, but reported no velocities greater than 18 cm/sec below 50 m. Maughan (1963) measured velocities as great as 25. 8 cm/sec at a depth of 10 m over the central Oregon continental slope. Collins (19 67) measured northerly subsurface currents with velocities of 1 3 to 27 cm/sec on the central Oregon shelf at depths of 10, 20, and 60 m during the winter months (Figure 41, left), but reported a southerly movement at these same depths during the summer (Figure 41, right). However, throughout any year the net movement was to the north over the shelf. Longuet-Higgins (1969a), in studying mass transport by time- varying ocean currents, determined that in regions of large bottom gradient the Stokes velocity factor (the difference between Eulerian mean velocity and Lagrangian mean velocity at a given point) may be an important consideration. He stated that the Stokes velocity may at times be opposite in direction to that of the surface wave propagation 137 direction, and hence the resulting mass transport may be opposite in direction to that expected from the wave propagation direction. He farther reasoned (1969b) that such a situation may exist along the con- tinental shelf-slope transition off Oregon, where a significant Stokes jet may give rise to a southerly mass transport along the upper slope with the flow being opposite to the wave-generated northerly mass transport over the shelf (Figure 41 ). Upwelling has been noted to occur along the Oregon coast (Smith, 1964), and is particularly well developed off southern Oregon (Pattullo and Denner, 1965; Smith, e_t ah , 1966). The upwelling is explained by the seasonal shift in wind direction off the Oregon coast. During the late spring and summer, the predominant winds are from the north- northwest and blow onshore to drive the surface water offshore. This surface water is then replaced by upwelling subsurface water (Figure 41). During the winter the winds and surface water movement reverses direction and the upwelling ceases. Neudeck (1970) has postulated that upwelling may be a significant factor in inhibiting the offshore and down- slope flow of bottom turbid layers. The strongest winds occur off Oregon during the winter months, when severe storms produce high energy wind and wave conditions (Cooper, 1958; Kulm and Byrne, 1966). Neudeck (1970) has related the occurrence of greater amounts of more competent waves in winter to an increase in the maximum depth of rippling of bottom shelf sediment. 138 He has shown that rippling occurs roughly parallel to the coastline to maximum depths of about 200 m in winter and only 50-100 m in summer. The seasonal shift in winds produces similar shifts in the pattern of waves and in the resulting littoral drift directions along the coast. During the winter months wave directions due to sea are from the south-southwest (Figure 41); however, the prevailing swell comes from the south-southwest a majority of the time only during January and February (Kulm and Byrne, 1966). During the summer, both sea and swell come from the north-northwest a majority of the time (Figure 41, right). Thus, it can be inferred that littoral drift along the Oregon coast is to the south during the summer and north during the winter (Figure 41). Quantitatively, the northerly drift in winter appears to be greater due to the larger sediment influx, during this season and the higher energy waves available to move it. The existence of a net northward littoral drift over the Oregon shelf has been verified by textural studies (Gross and Nelson, 1966; Gross, et al . , 1967) and heavy mineral patterns (Kulm, et al . , 1968b; Chambers, 1968). The work of Collins, cited above, and Mooers, e_t ah (1968) also suggest that a net northward movement exists over the Oregon shelf during the year. Tides along the Oregon coast are mixed semi-diurnal with a range of 6 to 10 feet (Pattullo and Burt, 1962). Mooers, etal. (1968) believe that the tidal influence accounts for a large percentage of the current 139 speeds measured on the Oregon shelf, and although these tidal current speeds average less than 10 cm/sec, they may cause turbulence or instability at the pycnocline due to shear. Tidal currents may be related to internal waves and internal tides which Mooers (1968) believes may be a contributing factor to subsurface turbulence at the shelf edge. Normal alternating tides and tidal currents may have a pronounced effect on the movement of suspended sediment within the water column and on the bottom sediment down as far as abyssal depths (Swift, 1969; Johnson and Belderson, 1969). However it is significant to note that Neudeck (1970) has considered these normal tides rela- tively un-important compared to the influence of surface waves, especially long-period swell, in forming ripples and in generating a turbid transport system on the Oregon shelf. Storm tides may affect the bottom; however, these are exceptions to the normally predicted tides. Bottom photographic studies have been made by Neudeck (1970) on the Oregon margin to determine the extent of bottom current activity. He found indirect evidence of current directions (no current velocity measurements were made) in the trends of rippling, scouring, and in the movements of benthic organisms. On the upper slope off southern Oregon, Neudeck found indications of both a southwest-moving bottom current and scouring at a depth of 1000 m in two nearby stations within the swale to the north of the Upper Rogue Canyon (Kulm, 19 69). 140 Korgen (19 69) has actually measured near-bottom current velocities (but not directions) on the central Oregon slope. He recorded velocities at depths of 700-900 m which range from 8. 4-17. 4 cm/sec (at 75 cm above the bottom) to 9. 6-45. 0 cm/sec (at 300-350 cm above the bottom). At another station, at a depth of 1600 m, he recorded velocities of 0. 4-12. 9 cm/sec (at 150-200 cm above the bottom). Harlett (1970) has verified a southerly bottom current direction on the upper slope, but was unsuccessful in recording its velocity. Neudeck has observed many areas of north-south trending bottom ripples on the southern Oregon shelf, which he has related to long- period swell. He also observed extensive areas of near-bottom turbid layers over the southern Oregon shelf and slope. His results show that: 1. in general the turbidity decreases with distance offshore, but increases sharply within 10 to 20 m off the bottom; 2. relatively turbid waters were observed on the upper slope in the swale north of the Upper Rogue Canyon, and in the Upper Canyon itself, while the ■water over submarine banks was relatively clear; 3. marked seasonal changes were evident in the turbidity of surface waters due to planktonic organisms, but the turbidity was usually higher during summer upwel- ling; 4. although near-bottom turbid layers were dense over the shelf during the entire year, bottom turbidity was more evident over the upper slope in winter when its generation by waves was greater and when the downslope movement of the bottom turbid layers was 141 unopposed by upwelling. Based on all previous work it is evident that the predominant surface, subsurface, and bottom current direction, as well as the resultant net transport, is to the south on the upper slope, while the predominant current directions at all depths over the shelf are to the north (Figure 41). Normal tides superimpose an oscillatory east- west component of movement on this system, possibly at all depths, while upwelling superimposes a west-to-east, subsurface to surface transport during the summer. Turbid Layer Formation and Transport Across the Shelf The mechanism of lutum transport described by Moore (1970) calls for the downslope transport of fine-grained lutum as bottom or near-bottom low-density turbid layers. These turbid layers are often thin and relatively slow-moving, without sufficient erosive power to remove fine sand or the tests of Foraminifera into deeper water. Depending on topographic or other conditions, they may be channelized into thicker, faster flows, or may remain as un-channelized sheet flows. In the initial formation of turbid layers mixed sand and mud of the southern Oregon rivers reaches the bays and estuaries. Much of the coarse material is separated from the lutum in the bay immediately seaward of the river mouth; part of the sand is restricted to channel 142 banks or tidal flats inside the bay (Kulm and Byrne, 1966), or sub- jected to alternating littoral drift and restricted to the beaches or the inner shelf shallower than 20-30 m (modern shelf sand, Figure 32). The lutum fraction is separated into a very fine-grained clay and colloidal suspensate (Figure 40) which often contains entrapped organic debris, and a silt, clayey silt, or silty clay fraction which eventually becomes the bottom turbid layer. The finest fraction (the suspensate) is carried slowly westward in suspension in the water column by alternating tidal currents, by-passing the shelf and upper slope and settling in the manner of hemipelagic sediments to deposit on the lower slope or adjacent deep sea (Figure 40). Part of the lutum fraction may remain in the bay or estuary; how- ever, most of it deposits over a wide area relatively near the river mouth (Modern Shelf Mud, Figure 32). This newly deposited lutum is subjected to re-suspension by long-period swell, and if the re- suspended cloud is of sufficient density, it will form a bottom or near- bottom turbid layer. Since long-period swell is most dominant during the winter when the sediment supply is greatest, turbid layers are most apt to form during this season. Long-period swell is a more effective agent for re-suspending the lutum than normal sea waves. It is mostly this former type of wave, with a period of 8-12 sec and frequently greater than 14 sec off Oregon (National Marine Consultants, 1961), that may attain sufficient bottom orbital velocity over all shelf depths 143 to erode fine sand and the finer-grained sediment. Inman (1957) has demonstrated that it is the long-period waves, of both sea and swell, which most often attain the critical orbital velocity of 1 0 cm/sec neces- sary to erode or ripple fine sand. Moore (1966) has also suggested that the long-period swell is the prime agent responsible for re-suspending lutum. Neudeck (1970) suggests that certain trends are evident when the frequency of long-period swell increases during the winter, among these is a corresponding increase in turbidity and turbid layer formation. Sundborg (1956) calculated that a minimum velocity of 10 cm/sec is necessary to erode very fine sand (62 fj.). Postma (1967) notes that the "critical erosion velocity" for fine-grained silt and clay varies according to its degree of consolidation (i. e. , water content). For example, a freshly deposited medium silt (30 (jl diameter) containing 90% water would require a velocity of 15-20 cm/sec to initaite movement, whereas the same size sediment containing 50% water would require a velocity over 100 cm/sec. In each case, a somewhat lower velocity is required to maintain the sediment in suspension. Thus, it is only the long-period waves, mostly swell, which can provide the necessary velocity for re-suspending recently deposited lutum on the southern Oregon shelf. Once the lutum is re -suspended into the water column as a near- bottom turbid cloud, it is subjected to the influence of coastal currents. A small part may move south over the shelf in the summer, but the 144 bulk of the material moves north in response to the net mass transport over the shelf (Figure 40). During the winter, when the sediment supply and coastal current velocities are greatest, the near-bottom turbid cloud more nearly approaches a very low-density current which moves northward under the action of coastal currents and westward across the relatively steep shelf under the action of gravity. The bottom turbid current continues to move in this fashion until one of several events occurs. It may continue to move north until it impinges against a natural topographic barrier, such as a headland or a rock reef. Numerous resistant headlands, chains of sea stacks, and rock reefs, are present along the southern Oregon coast. Cape Blanco, Humbug Mountain, Cape Sebastian, and Cape Ferrelo are examples of the first, and the rock reefs off the southern end of Cape Blanco and off the Rogue River are examples of the last (Figure 2). These barriers may tend to constrict the northward flow and temporarily increase the density of the current, or they may disperse or divert the flow. In either of these cases, the action of alternating tides together with the shelf gradient will gradually move the turbid water outward to the shelf edge. In summer when the northward coastal currents are weak, or if the re-suspended turbid cloud is initially in a less-dense dispersed state, the westward component of movement may be stronger than the northern component. In this event, the turbid cloud will move as a sheet flow directly, but more slowly, westward across the shelf to the 145 shelf edge. Thus as a consequence of the various oceanographic con- ditions in both winter and summer, together with the influence of bot- tom topography, the bottom turbid layer may reach the shelf edge most often as a broad sheet-like cloud (Moore, 1966). Transport Down the Slope Upon reaching the edge of the shelf the turbid cloud is subjected to: 1. the increased gradient of the upper slope; 2. the influence of a predominantly southerly mass transport and a resulting component of southwesterly bottom currents; and 3. possible turbulence arising from internal waves, a factor which may keep the turbid layer stirred up enough to prevent deposition on the outer edge of the shelf. In addi- tion, topographic features may be present which would tend to channel or divert the flow. The submarine topography of the southern Oregon continental slope reveals many features which may serve as natural barriers or channelways to sediment movement (Figures 2, 4 and 5). The most obvious of these are the numerous east-west trending submarine valleys, especially the Rogue Canyon, Brookings Seavalley and Blanco Seavalley. These valleys serve to channel sediment from the upper to the lower slope. Another important valley is the unnamed swale to the north of the Upper Rogue Canyon; this swale trends northeast- southwest and probably feeds sediment into the Lower Rogue Canyon. 146 The topography of the southern end of Cape Blanco Bench (Figure 5) strongly suggests that a system of north-south channels is present which may also serve to channel sediment. Less obvious, but perhaps as important, is the character of the continental slope itself. The downslope gradient of the continental slope gives rise to normal down- ward gravity flow and slumping of sediment from the shelf edge to the upper slope, and from the upper to the lower slope. The benches on the upper slope, which interrupt the steadily descending gradient of the continental slope, can be considered as rela- tive "highs" because of their less steep gradient (approximately l°-2° versus an average of 3° 37'). In effect, this may have several con- sequences depending on the direction and velocity of sediment-laden currents or water masses which may move over such "highs. " Downslope -moving bottom or near-bottom sediment loads tend to deposit coarser-grained particles on these benches upon reaching the abruptly decreased gradient. Any other slope currents would tend to flow faster over these "highs" than over the lower slope, since the "high, " in effect, has constrained the current to a narrower space in the water column. From a topographic viewpoint, the benches, and much of the upper slope, may be regimes of coarse sediment from which the fine sediment is being continually removed or winnowed. As the turbid bottom layer progresses down the slope it is often 5-10 m thick (Neudeck, 1970), and may become channelized by entering 147 sea valleys such as the Upper Rogue Canyon, the swale, or other val- ley systems (Figure 40). Neudeck has found high turbidity in both the swale and the Upper Rogue Canyon. In the swale, he also observed indications of southwest-trending bottom currents and scouring at the same station (970 m) where a significant amount of turbid water was observed. He postulates that the scouring and current activity were probably generated by the turbid, more dense, near-bottom waters flowing down the swale in response to gravity. This explains the relatively low Holocene sedimentation rate (9 cm/ 1000 years, Table 2) determined for the Upper Rogue Canyon, as well as the scouring observed in the swale to the north of the Canyon. One must assume that the velocity of the turbid layer increases down-canyon due to an increased thickness and density of the layer, and/or due to an increased gradient. Moore (1970) suggests that the velocity of the turbid layer increases only slightly above 10 cm/sec in its transit, and this is not sufficient to transport shallow-water foraminiferal tests into deeper water. But Neudeck and others (Owens and Emery, 1967; Stanley and Kelling, 1968) suggest that the velocity of such turbid layers may increase enough to cause distinct scouring. It is evident that whatever the resultant velocity of the turbid layer may be, it is enough to pre- vent significant deposition of lutum in the upper reaches of the Rogue Canyon, and perhaps in the shoaler portions of other valleys. The 148 northward-flowing lutum may have sufficient velocity to by-pass the Canyon head and eventually divert into the swale to the north. In any case, large quantities of lutum do become funneled, or channelized, and eventually deposit on the lower slope. This process may be the most important mode of supply for the fine-grained sediment on the lower slope. When combined with the by-passed suspensate fraction, the two probably account for the total Holocene and modern lutite cover on the lower slope. That portion of the bottom turbid layer which does not become channelized probably flows down the upper slope in a sheet-like man- ner. It has been established that the Holocene sediment cover on the upper slope is thin, and often patchy enough to reveal "windows" of older, late Pleistocene, relict sediment (Figure 32). This is true of the Upper Plateau Slope as well as the benches. Moore (1966) contends that once lutum has deposited, it forms a fairly stable mass and is not prone to failure by slumping. If this assumption is correct, then the slumping observed on the Upper Plateau Slope (Figures 21 and 22) is probably a Pleistocene phenomenon, due to the failure of great masses of sediment deposited there during lowered sea level. It is postulated that this process did occur, which left large areas of still older Pleistocene material exposed on the upper slope. The unchannelized flows of Holocene and modern lutum do not deposit at as high a rate as the channelized flows reaching the lower slope (10 cm/1000 years 149 versus 50 cm/1000 years); hence they have failed to cover the under- lying Pleistocene material completely, leaving a patchy sediment dis- tribution pattern. In addition, the increased gradient from the shelf edge to the Upper Plateau Slope and the possibility of a Stokes "jet" here may be sufficient to prevent deposition of all but the coarsest lutum fraction on the Upper Plateau Slope. The greater portion of the un-channelized lutum reaches the less- steep bench areas on the upper slope, and is subjected to further diver- sion, mostly in a southerly direction. The channels on the Cape Blanco Bench may serve as routes of north-south transport for lutum from the Umpqua drainage. Whatever quantity does move south within these channels is fed into the swale to the north of the Rogue Canyon. Silver (1969a) has observed similar channels on the southern Klamath Plateau which empty into Trinidad Seavalley. He believes that they are routes of sediment transport into Gorda Basin. Although no such channels are evident on the Klamath Plateau off southern Oregon, the southerly mass transport over this region may sweep much of the lutum into Brookings Seavalley, which carries it to the lower slope. Core 6711-1, from Brookings Seavalley, shows the highest rates of deposition (60 cm/1000 years) of any core on the lower slope. Although it appears that the abruptly decreased gradient from the Upper Plateau Slope to the benches on the upper slope would give rise to the deposition of much of the lutum, several facts contradict 150 this. First, the southerly mass transport mentioned previously may sweep the lutum into an available downslope valley; second, the sedi- ment distribution pattern itself (Figure 32) shows that little Holocene or modern lutite is present on the benches; third, the near-bottom current velocities measured by Korgen (8. 4 to 45. 0 cm/sec at a depth of 700-900 m) indicate that the velocities are high enough on the upper slope to keep the bottom turbid layer in transit, or even to winnow any newly deposited lutum from the bench surface. Topographic highs, such as those at the western edge of the Klamath Plateau are another factor which affects turbid layer movement. These highs effectively dam any westward-flowing lutum, and give rise to a thicker ponded accumulation behind the high. Although this undoubtedly occurs, part of this accumulation may also be winnowed and diverted into down- slope channels. Deposition of Lutum on the Lower Slope The distribution of Holocene and modern lutite on the lower slope and the computed Holocene sedimentation rates (50 cm/ 1000 years average, Table 2) attest to the fact that the lower slope receives the major portion of lutum generated from southern Oregon drainages. Lutum reaching the upper slope is either channelized directly and funneled onto the lower slope, or it may deposit on the upper slope for a brief period and later become re-suspended and channelized. In 151 either case, much of the lutum eventually funnels into the major dis- tributary channels and accumulates on the lower slope. Much of the lutum from the Upper Rogue Canyon and from the swale to the north feeds into the Lower Rogue Canyon. Sedimentation rates suggest that the Lower Rogue Canyon (core 6711-6, approxi- mately 75 cm/ 1 000 years ), the swale (core 6708-38, 38 cm/ 1 000 years), and Brookings Seavalley (core 6711-1, 60 cm/1000 years) are rapidly filling with lutum. Thus it is not unreasonable to assume that all the seavalleys on the lower slope, and the inter-valley areas as well, are also filling as rapidly. The entire lower slope is growing and out- building in this manner. Duncan (1968) has examined the Holocene sediment in Blanco Valley adjacent to the southern Oregon slope, and has calculated Holocene sedimentation rates ranging from 44 to 100 cm/1000 years in four cores from this area. The Holocene section in these cores suggests that a distinct bulge of Holocene sediment is building up at the base of the southern Oregon slope. Data presented here Indicates that the lower slope is upbuilding in a similar manner. Downslope gravitational processes, both within valleys and in intervalley areas, are the dominant mechanism for transport of lutum on the lower slope (Dott, 1963; Shepard, 1965; Moore, 1970). Although bottom currents are probably present in the lower slope valleys, none have been adequately measured. The newly-deposited lutum may be a stable mass, as Moore (1970) has suggested. However, the margin is 152 relatively active tectonically, and the lower slope possesses a fairly steep gradient (up to 8°); both of these facts suggest that considerable slumping may occur on the lower slope, such as that observed in Figure 21. This slumping is not as large-scale as that postulated to have occurred on the upper slope during the Pleistocene, but it still may be quantitatively important. Locally, the numerous small anti- clines on the lower slope may serve as dams which give rise to a num- ber of thickly ponded areas behind them. This is similar to the ponding noted on the upper slope, but on a smaller scale. The mineralogy of the lower slope sediments indicates a greater influence from Tertiary rocks than do the sediments from other environ- ments. Much of this may come from the Umpqua Mid- Coast Basins, especially from the Umpqua River. If this assumption is correct, and a similar turbid transport mechanism operates north of Cape Blanco, as Neudeck (1970) suggests, then some Umpqua sediments may be diverted to the south around Coquille Bank and onto the southern Oregon margin. Blanco Seavalley would be expected to collect much of this sediment and funnel it into the deep sea. However, no core has been taken in Blanco Seavalley to test this hypothesis, and Duncan's cores (1968) in the adjacent deep sea do not reflect this. However, it is sig- nificant to note that one core near the mouth of Blanco Seavalley (6604-12, Figure 25) exhibits the thickest Holocene section of any of Duncan's cores. Consequently, it is not unreasonable to suggest that 153 Blanco Seavalley contributes the major portion of this Holocene material in the form of lutum. The Late Pleistocene and Holocene Sedimentation Patterns The great influx of sediments during the Pleistocene originally- covered much of the southern Oregon shelf, upper slope, and lower slope (Figure 42). The benches are primarily Pleistocene phenomena, having formed by the ponding of Pleistocene sediment behind anticlinal folds on the upper slope. The Upper Rogue Canyon, although struc- turally controlled, was primarily cut during the Pleistocene, as were the majority of channels and valleys. Evidence from glacial marine sediments in the adjacent deep sea (Griggs and Kulm, 1969) suggests that the paleo-oceanographic regime was essentially the same in the late Pleistocene as it is today. If the major winds and currents were essentially the same, then it is not unreasonable to assume that the paleo-oceanographic regime over the continental margin was also similar. As a consequence of the rise of sea level during the Holocene, the deposition of coarse-grained material was greatly reduced. Even so, a uniform Holocene sedimentation rate of 50 cm/1000 years over the entire margin should have covered all the environments with a sediment cover approximately six meters thick since the start of Holocene time. However, the model of lutum transport just described 0 -l-i +■> 0 o 2 •rH 4-> 3 •H £> +-> •lH ro M •H +j Tl co •r-l 1—1 •n nS 0) 4-1 d M 0) rt a T3 • H CD T3 tsl 0) •i-i i—i UJ rt <1> a t— I -i-> fl) d T3 o C\J <* CD u d 155 has shown that such a uniform sediment distribution does not exist. A thin Holocene cover, averaging only about 3-4 m thick, is present on the upper slope, benches, and in the Upper Rogue Canyon, and a much thicker cover, probably averaging 10 meters exists over the lower slope, especially in the seavalleys (Figure 43). Pleistocene sediments were laid down over a longer time period, and at a higher rate, than Holocene sediments. Hence the Pleistocene section is probably more widespread and thicker than the present Holocene cover, perhaps as thick as several hundred meters or more (Figure 43). The uneven deposition of lutum during the Holocene has left large areas of Pleistocene sediment still exposed on the surface of the shelf and upper slope. This sediment is now relict, being out of equilibrium with the present environment. Even though Holocene and modern lutum now dominate the lower slope, the total thickness of these deposits is not great, and would probably be penetrated frequently below 10 meters. '^i 8— LU X en < i- UJ o o o w 10 q- M <0 <" wft Ll_ S t _l •as LU Ou T o to I _l - — < \- z — LU UJ 2 > I— lO CO Z LU Z ( > lu a. < <_> z o tf &UJ CO SU 2 q < ^ tt UJ 2 H z X £ < (0 ,fl -M fl 0 fl O •iH fl -»-> 0 u -t-> ■t-l CO •H J-l 4-> T3 CD 1 — 1 •H rtj T3 CO -t-> u fl a <1) £ o •r-l <+H Tl 0) CM CO CO n 0 U fl X! OJU O h 0) x) -u •j-j O «4H O PI PI 0 M •H O rrt 0) ti CO 1 fl CO O CO CUD 0 CO u ^ o 0 T3 fl CO M N N CD M ^ U > — . M a 4-> U C — •" X 'a js i a « CO Q > cu co u cd CD z x" O O £ ■i-i to CD co D a o -u O CO h-l l>s X CD XI rd co CD CO cd El rd CO CD u o o o X CD in cd X u ■r-l X pq U PQ i 00 o 00 sO X cd « 00 o r~ X CD O CD to CU u o CJ C! 0 o In u-< CD rd co CD r— I a CD GO rd CJ o X rt u o •H X CD XI -i .-i H E-1 * III II I (Tir^'^,mt^O|Ji,*,Lr)ooo^--im Hmm^lioioo|0''ii') i *~\ i ^s i i r-H r^ i^ t^ i **\ i /^ r^ irt VO*-i^tJ NOOWNNlOHWrn rofOrorotNjrocOfMroromwim cococororocococococo it^^o^coa^fMOOoo'^ot^ i oo o u"> vo cm i in o cm IMrt^^^SN^^lO '■O I CM of) ■"tf" LO lO I ■■* rr> I 000>HlOOOCfflO>NHOOH I CO K CT> CTl c/5 lO^OVD I lONiniO^NNinNNOim l-rhvo-^roro I-* ■* |ini/)Ol(MB)NNNOl')NH I m m Ifl ^ ^ I <0 O CM I LO CO -r* I (M I O -l ,-H .-H .-I CM CM (M Cu ~H iH 1 o 1 R vo lO (Mrfl'ftniONMlJl I I I I I I I I I I I I I I I I I I I I I I ooooooooooo (Mn^mioNtosi ■ i i i i i i i i i CMCMCMCMCMCMCMCMCMCM I I I I I I I I I I oooooooooo 187 OJ S 3 (3 > G O +■» o ni « o to Qu EaajraiuiEjoj DTipuag BJajIUIUIBJOJ DIUOrppiBJd SUIOIEIQ SUEIJBJOipBy SJ3qiJ JUEld SSBjS OTUBOJC>/\ EDIJM SUIElS JB}Ii:j3Q pajunoo suibjS psjoj. Asp % CM ,-i .-i ,-h ^ 1ITS % PUBS % H o u co < adAx juauiTpag 3§V (uid) 3ioo ui ipdaQ jaquinu ajduiBs OOUI'tHOOiriHrtNH I N I l(M 1 E — i -. — t ■. — I r — h- tx CM I II I CO CO «— I »H ,— I t-I i— I *-l ^ —I rf tfl ^ N N IO ■* ro # (M *-H i ro H «H tH f— ' H I «-H I I I I I I I I I l cm ro -h *-h i i I I I ro CO i H CM «-« «-i K IO CO nhimimnnI'iohi/) Hininmioi'uiuiro CMCMCMcococo icm i i mHHfo^coNmh I I I I I I I I I I I I I CO I CO *-l I U0 *-( lrt i-l I I I I I I I I I I 1—1—1 1 KtvinCTiLOIOCOIOO'-O inNM^OClNNOlS h o> lo I'lOOOl'roiOi'mN© (J\lOiBNniiOtONOiO Ocoo r— j*-tvrtmij"ir\ir-^,— 1»— ir\i ems^-*^ia\onrm\-m^t/\r^r> n\ n\ — * iaj w_i u/ ui ui i\j <_; ui i— i i^«* i\j ^' uj i^ mj i^ v» rrj ui >iP O^lOKliflNOHHN Oi-i^O^OOO'-'mOO &i CT> t-i i'l'I'mtO^fOfO COCOCOCMCMCOCOCOCOCO N N •* oo ii-icommipTfroip UiNNOitpmtOHOj^ n in in CM Ii/)hO\0>I3>«)hO UIOloNOlHiONrtuI O'-'CO Tf immTii-^i'rtl'*mLn ^iflcom^inininni1 in ^f in ctv i^ioioiol'rti/ii/i n s N * i/i (o io t-it-i i1 n m T— I I ^H -H rt K H N hJ hJ hJ o^oooooooo ooaouDODod OOO CL,o-cuD-,pL,oup-a,D-iCu &,a.fca.CuCueH(Lhh _ _ _ _}_|_J_l_i_l,-)_3.-l|-4 _}_|,-l_]_|_ll_q_J_l_I I H s •*cocMm»-iLoo'rtQQ mmwmmmmmt-»r--. o m r-~ CMCMCOCOCOCOCOCOCOCO r o ■^H O .— I 1—1 1 H ro 1 CM 1 CO 1 CM PJ 1 CM 1 CM r-j r\! CM | CO ro i ro ro ro 1 ro i ro 1 ro 1 CO 4 -t 1 O r IO o IO IO o s IO o IO R IO o ID O IO <0 o IO IO o rs c o •IH *-> o « 01 u HI a. ajiuoDnBTni i i in Sjajpd p3Daj BjajmiuiBjoj oiipuag EjajIUIUIBIOJ DTUO!>fUB|cJ SUIOJBIQ SUBIIBfOipB^J saaqij JUb^j SSTB];S DTUBDJOyV 1 '^,H'-i'-''- CO » (O OOCOOONNloSoOOOOO 00 CM J2 0\ lo CO (M CM ro •* CM CM H co ■* I I CL, £ CM —t tv H -* N -* CM I I r-» O ^ i^ N N n s m LO 189 X> S I a o a o to 0- 9JTUODnB^9 siajjad xeo3£ BJ3JTUIUIBJOJ Ditpuag BJ3JIUTUIBJOJ SUIOJBTQ SUBIIBpipB-tf siaqij lUEy SSBjS OlUBOJOy\ EDIJAJ SUIBlS TBlIJ^an -h tv. 00 CM ^* f^ t~^ IN ■* 00 a CT} CM ffl N O CM cm ro h in CM •* CO CM mWrtj-NtMHrM CO CM * CM Ifl -H N *H CO <-• <-< CM i-H I Tt< CM I CM vo CM cm •*inroinroro,*iO ro i i i H i '-• u"> H ■ i i a a CO w ro " _i ^ m CM I CM ro in ro * I CM ro ■* O * I CM * CM -h ro N LO CMCM in CM CM NhN(MI/)1'N0> «-l I CM-* l£> 00 (O lO « O'-<'*00tN00OlO N H rl ■* tTi oo mw^i/iiomNN m s n ro* LO CM f^ i-H IO N N m«)N0OLOC11ll/l CMCMCM lo in * in N d> ifl 1/1 IO ID 0"*0\<£>N.'-'COCM N ID N NlOSiOOiOifliO K N 00 psiunoo suiciS jbiox in ■* * in ro ro Q CM CM CM * Id 00 * Tf i£> -* oo io in •* r-t CM CM CO CO CO lomifliO'-i^toi1 io m Oi Ortirt^iooiinH in in oo cocOrororoCMCMro rococo Abtd o/ oo g mcM m -l O T-H «H lO ^f <0 adXj^ luauixpag a§V (uid) 3IOD ui ipdsQ aaquinu afduiBS o o o o ►J J II o o IN fn \o '■O CM I o o r-. r-- _} _l J J .. o u a o oo- x> 6 o 4-> o B u u 01 0- SJajpd p303j BjajraiuiBioj Diqiuag BjajiutuiBaoj DIUOPfUBU SUIOJEIQ SUBIiejOipBy SJ3qiJ jXIBJd ssrejS otuBopy\ BDty\f SUIBjg JB3IJJ3Q parninoo suibiS psiox Abjo % Ills % 190 pUES % (MH^iniflTfrflM fO 00 ^ W 'f (fl N(M(MNN(\1N(M bO H tM N ^ W to I h h N h (M >* fl I n * ^ i/i E — • cvi •■ — • E — ■ »— • E-* •— * E — • I I I iH I I I I i-l .-I CM CM CM CO CMCO'-icOCMt-ICM'-h h m N n m ■* Nio^ini'ioin"* ^oo^omcM'*''^,al•''tfrO''*>*mco m n o t oi h mooinHoooiiOH H^rtNiONNiOH CM LO t-i O CM m rOOOCMOCMcorOio rocororocoro^roco co CM co co rO co cocococorOMiroco inNNHKimOlOlO 00 O CM O ■— ' 00 CMWCMCM',;tl/">'-i'-i i-H »-H *-« *-H t-4 *— t i-Hi-H CM rOCOrO ts*'-< x x iiiiiiii wiflwininininmin to t/i c/j O O O ooooooi/jt/oi/oc/)!/) 3§V hJhJ.-1i-Ji-1i-1i-J.i-J.-I . — I . — J i — I , — J . — J , — ^ 1,1,1,1111.1, OJ s i o u a Ou Pu < (uid) 3jod ui qidaQ jaqiunu ajdutBs inocooooLnom o o o m o m iocMr-^o*oocMr-s CMOmCTiCMCTi^Hmvo ^OLOiOf^. Lnt^OfMt-^OO rtNNN(OfO'!|lil7 H N rO CO .-I i-h ,-) CM CM -* io I I o IO O *-i I I I I IO IO o o 3 IO IO IO a 1 I in i VO 1 i i I 1 1 i IO 1 < CO CO u ro Q ro w ro ro 0 ro X ro 1 '-- N fs. K 00 00 00 O0 CO 00 CM oj CM CM (M 1 OJ CM 1 IO ip io IO IO >o IO iO IO 10 10 IO lO IO IO IO 191 aiiuoonFj;} ~* 1-4 » n m n N OJ E c a o u 0 rt o to PU BJ3JIUIUIB.IOJ onpuag BJ9JTUTLUBJOJ SUIOIBIQ SUBUBJOipBy saaqtj iub^j SSBjS OTUBOJOy\ BDTJV SUIBlS jBjtjiaQ paiunoo suteiS p3}o.L CM CM Ifl ^ H N H H N (Tll/)^Hl/)HN^(Mrt I I I I I I — < H * m w Tf n w io lO'tmro^NiMrororo -^ ^ H i i i i i i i i i i i i i i i i i i i i i i i i i i i i i i i i i i a a H H rt (M I "-• o ct> m ■* mm cm m m m cm m m pUBS % 1 O 51 N (J> ^ lO «) (O T3 g U m Pu a. < adAj, juauiTpag aSy (uid) 3iod ui tpdaQ jaquinu ajdiuB$ ,~ ,~ Irt I— 1 1—J rt Irt rt (-J "? <■£> in in in in in in in CM CM CM CM CM CM CM I I I I I I I 00 00 CO 00 00 00 00 o o o o o o o K N tv t^ t*^ t^ t-- lO lO IO IC lO lO lO O'-icMm'^in^ot^ OOCTlrtrtrtrtrtrtrtrt i i i i i i i i i i inmininmininminin CM CM CM CM I I I I I I I I I cm H m ^ t^ in in -*mr^ m ■* n (Jii'inooioOi'^iooooo io t^Lnmioioinvo ooinin^^m'i'iol'N »-i rt OlONOlNrtlBOONM N CO rt *-l *-t y-l ■*-! r~i y-i tH 00t"N.CTl OOOOOOOOOO w in w fx p_i o_i M-< iJ-< HH rt rt JU |Jm hU t-M rt rt I— H rtJ ommmmmininmo cr> en en QCM-^ipoOOCM^COrt n ^ ^fi'^^minininio CM CM CM CM CM 00000000000000000000 oooooooooo NKSNNNNSNN • -o C o •rH u rt a o b a, S-Wjpd psDaj EjajIUTUIBIOJ oiqiusg BJ3JIUIUIEJOJ SlUOJEIQ SUBUEJOtpBy saaqij jUEid SSBfS DIUEDpyV SUIEjS JEIJJaQ pajunoo suiejS Je;ox ins % pUES % adAx luauiipas (mo) ajoo ut qjdaQ jaqtunu a^duiEg H oo Q co ro I 00 o a "* "* m rg rj i i 192 I CM -H [-H [-. I I I I I I I I I I I I I I ^ ^1" ro (M -h I I I I -h ^ ■* N (N N lO rH VI (M ^ Tf 0000 "* CM CM f-> CO l£> —i in co oo io m ^ w to oo roOi cm N n co io oo n n> ro ^ K 'I' ■* to ifl tn oi ro ■^•i-^ o O* «-< p* 0\ r< ca o i i i i i i i i l l i i i i I CM •* \o VO lO 00 ro rt rt i I I O O cm oo ro cm oo oo n n in m o oo >-« r». m n to » m io n n o\ loior^Ktvr^K t^ "*lo ^f m n m io io oo o o » co o\ "* in «i en m ^f s in h ^ CM CO ro CM CM CO vo m in CM IM ro ■* CO (M « h 00 O (J -J 1-1 l-J "? U U O U j H _i c/> ooooouu o- o. a. n. b. a, a. a. _| -J J J J J J _l rt" rtJ j _l a o o a o o o in in in o m o m "-i O O rt Q «-< N- h cm m ^ m in rt n ro ^ in m s i i i i i i i t^ t>. t-x t>» k k t^ co ro ro ro ro ro ro 0- kJ 53 DC ro co ro 00 ^ lO in j j j j j j j u u u u u o o o o o o o o o MH I— l— < HM mU |Jh ^-< HH ro in ro ro ro ro ro O *0 rt *o rt ' CM CM ro ro ■* »-l lO Q ^ H lO rt ID I I I I I I 00 00 00 00 00 00 00 o o o o o o o r>. t^ r>. rx t^ t-x rx o c io id m c io 00 CM ■* lO CO O N lO oo O f- 00 R 00 O N 6 g > a o • rH u SJTUODriB^Q ■BjajiuiuiBJoj onpuag BJ3JTUIUIBJOJ omoppiB^ SUIOJBIQ a u 0- SUBTIBJOipB^ saaqij iub^ ssb^S diubo^oty BOIJAJ SUIBJS ]Eq.TJJ3Q psjunoD sutbjS imojL ABp % ins % PUBS % •H 1 o u co Xi a. < ad^x luauttpag 3§V (uio) 3joo in ipdaQ jaquinu ajduiBg (Nj ,H ^H [_ f_ ,_, ^ i i i i i i i i i i i i i i r-i £— < [— » *-» T-H I I I I 1 I I I I I I I rtrt^mfONN^^N i I I I I I I I I I I I I I I I I I I I I I I I OiOi^^OMajmo^Niodi OOI»ONU1NO>Nl()NaiNOO OOOOOOOOOOOOOOOOOOOOOOCTlOO cO'-ioooooiOLnLntMC^Locg'* mmfOfONNNNmNNNN (M ro h N ^f CM CO tH CM >-h CO -^h cOCT^CMOOlO'OQOr^VOt'^cOi© N (O H to ^ in H OOOOOOOOOOOOOOOOOOLOOOOOOO I I I I I I I I I I I I 1 CuCuD_0-D_Q-CL,QJp-,CuCijD-D-. _}_}.Jt-ll-3l-J.-J.-ll-ll-ll-l>JtJ o m o cm t^ a. a, i i i CT) CTl CTi CO CO CO I I I 00 00 00 o o o t^ N - >-icocO00cO00cO00'-ic\] ■- rg n n m m >-t — i 193 £P w> bo bo HHroiH HH in-* N n m i i i i 9 0- cu cu eu Ph J _] _} J _| n m n o ro ■* 00 10 ■* »-i CM 00 00 00 00 oo o> m W rt 00 (M (M CO "* (O W 00 OO N N N N (M N N ro (fl N »h ^h »-i 0\ ro o W rt Tjl («| ■"* O t*- ro m to -h H (\| H H rt lO Cfi W) ffi to ■* 0\ ■* CM K •* <* CTi a> l/l I/) W M I/) 00 CO I I I I I II 00 00 00 00 00 00 00 eu ts. r^ h» K rx t-^ r^ r^ r-- IOIOIOC10IOIOIfllOlOmrvror^ miooo'^fr^otx i i •* m rt C N 00 lO N I I ■* to Tf in ■-H Tj< *-* ^H ro^fmrfminrorortrO'^ i h n ^< ■<* ■* c\j *h in OO^NooNOO^f) — < -h-h (M CM CM HmmcMooioc^'^iO'H^HHiMfo cm ro cm I *-H »H i r- oo i h n h Irt N H N «-H -H I I I I I I I I till I HNHNmOO^ONO •*r^ m m >t * ^ n n O N in O ""> O *-< t"^ 0\ Oi mO'-'tv»>OrOOOOOOf^OlM rt 00 OO Ol Q N ^ lO (O ^ IO ^ 'I' ^ 00 IO '^'lomQI^Tt^OOOOOlOOOO 00 N Ol to M 03 N 0'-<<^o^mt^c\]moin'<4. m CT\ ro o mmNNmNmminNmNfO^ i ^ in -tf *** in OO N 0\00KC,lWH0ii,rtHOO h oOOOOO'-'OVO'-fO^tH^H^H I DO IO ifl 03 N N 00 0\ t-H I H t-H °? -l»-l*-lf-liH*-H<-lT-li-li-H-,-l*-l»-l<-l tH *-l T-l *H i-l *H iH lO lo ^aK)iOlOIO>0>OlO>0>0'0'0>0 lO c >o © lO * w 195 si x> £ a >. a o o U pu BJ3JIUIUIBIOJ BjajiutuiBjoj DTUOUpiBXd SUIOIBIQ SUBtlBJOtpBy saaqij WBid SSBjS DIUBDJOA BDIJAJ SUIBJS JBlUjaQ pa^unoo suibiS jbiox Abjd % PUBS % adAj^ juauiipas aSy (uid) 3joo ui i^daQ laqumu ajduiBj; HlOnhrtMNrt .-I H "-« LT) H OOt^O'*'-l mKiiomsootOKifomNH r-^O'-'^t'f'OCTi'-i'— '<--l<\]CMt\lfMrO O'-ifMrO^LOiOl^ooi O N r~.ooc^>-(<-<*-ii-i«-i*-ii-i7-it-i HNm^uicSMO'iHH I I I I I I I I I I I I I I I I I I I I I I I NpjrjNfMojfMrjNNNN mmuiirimminuiuimif) i I i i i i i i i i i i i i i i i i i i i i i 6 3 c _o u 3}iuoonBfQ srjajpd pEDSj BI3JIUIUIBIOJ oiqiuag EJ3JIUIUIBJOJ oraoppiBu SUIOJBIQ PU SUBIJBpipBy SJaqij jUEid SSBfS OtUBDJOA BDIJAJ suibjS jBjuiaQ pajunoo sutbjS ]ejox Asp % Ills % pUBS % adAx juauiipas 3Sy (uid) 3jod ui qidaQ aaqtunu a^diuBg 196 i in o I I IO N 11 4 I I I I I I I I O^ CM t-i CM inNdlNHlMtONH CM CM 'tf »■! (Ol/llOTfl/lIfllOlflN «H CM CM «-i i - — l 1 I I •rt H •* O I 1 III ^ I O, CX 6. rv Dl, Hrt»*H(flHHIfl CM CM I I I I I I I I I I I I CM rt t» •-! O »H t/1 •^h ro -h Oi O fM O ro m co N to ci ro IJlOHNfJiflOI'W oo NSfOWNOOSKiO O CMCMCM-HCMCMCMrOCO co + LO o irt O O lO VO in m h in oo n h m w ^ ^ ^ ui ifl to I1 m io n Oi o O o m m in o o o o in o 0ONCONNN«lO\^ (7> 00 O M TH -H CM LO tO^t'NN^I'tMN I O ■* CO I ~ CTl lC^CMtvmCT>O^0OT-i I ■<$< -g< -sp -31 -^ tj< icmcmt-ht-h'^H'^hcM'-i i 000^,-h(M'^ I ^OhiMhionn I I —I I J J J J J J J _1_)_1_J_1_1_1_1_1 _) OOOOOOO OOOOOOOOO o a OOOOOOO OOOOOOOOO O 0 i-U —L J-. J-i —U J— HH J-h J-" HH l-M —U J— JU JU |_1h l-M -a 1^ r^ r^ t^ f^ t^ t^ (0 0 10(0 10 10 10 10101010IOI0101010 CO 00 CM O O ■-< to th O It^oOfMrOO-*'-!^ I "-" lO 00 in in io m m in in iionoooooo^jNoo i m ^ ■sf 0> ^ CM ■*t< in in ►J hJ hJ o o O O in in (O (O in o t— < *— t —i —i T^ ~H CM — i —| (M co ■* in to [^ 00 *-H «-H CM CO in i i 1 m in i m i in (O i to 1 to 1 lO lO lO 1 1 1 10 1 to 1 VO 1 00 1 oo 1 00 1 00 1 (O 197 01 E 3 a > c o 6 0) o o> ft "8 3 o u * 9r airaoonEjQ sjajpd psoaj BJ3JTUILUEJOJ Dtqqnag Baajiuiui-EJOj DTUOriifUBJjJ SUIO4BIQ SUEIJEJOipBy saaqy iubjj SSBjS DIUBDXOA EOIJ/\[ suibjS peqijjaQ pajunoo suieiS xeiox PUBS % adAj, luaunpas a2y (uio) ajoo ui qidarj jaquinu ajduiBS S *< 3" in h H i co cm co ro i-i C<1 -* -* CM — I I CM I I iflNmmtcd^mc I1 lO M IO UlNNHmmHrtK) i i i i ■ i i i i i i i i i ro ■<* t"^ f^ ft a, o. a oo 5 -^ o ^ T CI «1 in 00 O CO I I I I I I I I 00 00 O * B> 00 O N H rt N N Ol N O » m N io h ro CM *# m co m m o o O «h J J hJ hJ l-J o u u u 3 o o o o 0 dc dc ac rn in in O to in m o in O in t-H CO cj ro ro W * N 00 till 00 00 00 00 I I I I cocgc\i<-i<-i-HC«j«-i*-< lONhHiflHrtfUN I I I I I I I I I I I I I I I I 0»- o rx rt n) "ft • iH 'a < en a 3 3 OJ 01 £ 0 3 S OJ O CJ OJ c OJ 3 TJ 01 £ a ■ft E o B 0 CO y o 4-> rt TJ P T) 0) 'ft 0) u E 4-» rt £ 0) rt bo O "a. M n) •rH S 0 01 cu Ui TJ u. ^) Mh 3 ra E 3 p4 ■cH co u u w 3 »fH OJ .^H QJ a 3 ft A O £ 01 I C 0) V a bo CO > OJ "3 (3 3 OJ 0) OJ 3 CJ o U 'EL E bo OJ -3 ft -^ 1 TJ LO 0) bfl a 1 — 1 cj c bo 5 6 ft 01 -3 CO '5- ft TJ a to (X) CO CO •m a QJ 11 ,,H +^ X 01 TJ TJ TJ i 3 3 3 V! ^-< _2 o Z y U 0 ft c a 3 1 XI O TJ OJ >« APPENDIX 4. TEXTURAL ANALYSES OF SEDIMENT SAMPLES 198 Depth osua Interval M^ V V Sample in cores Sediment Percent Percent Percent Textural Number (cm) Type Sand Silt Clay Classification 6706-1-1 0-5 OGL 6. 51 2. 17 -0.05 5.2 67. 5 27.3 CLSL 6706-1-2 8-11 S-S 5.75 2.22 0.08 55.2 28.7 16. 1 SLSN 6706-1-3 55-60 OGL 7.15 2.38 -0.08 8.6 50.9 40.5 CLSL 6706-1-4 100-105 OGL 7.40 2. 10 -0.08 1.6 55.0 43. 4 CLSL 6706-1-5 145-150 OGL 7.83 2.04 -0.04 3.0 47.6 49. 4 SLCL 6706-1-6 200-205 OGL 10.58 5.41 -1.00 2.4 25. 9 71.7 SLCL 6706-1-7 250-255 OGL 8.85 1.52 0.00 1.8 28.3 69. 9 SLCL 6706-1-8 300-305 OGL 9.02 3.88 0.58 1.8 58.7 39. 5 CLSL 6706-1-9 346-352 GL 8.71 4.99 0.23 30. 2 22.2 47.6 SSC 6706-1-10 405-410 GL 10.08 5. 92 -0.13 15.0 20. 9 64.0 SLCL 6706-1-11 425-430 GL 10.00 6.00 1.00 2.0 98.0 0 SL 6706-1-12 430+ GL 10.83 5. 17 -1.00 1.4 31.4 67.2 SLCL 6706-2-1 0-5 OGL 5.62 2.72 -0.61 29.3 42. 8 27. 9 SSC 6706-2-2 45-50 OGL 6.74 2.43 0.38 3.0 66.9 30.1 CLSL 6706-2-3 110-115 OGL 7.23 2.41 -0. 17 5. 9 49.4 44.7 CLSL 6706-2-4 150-155 OGL 8.43 2.20 0.02 4.3 39.4 56.3 SLCL 6706-2-5 185-190 OGL 9.70 3.66 0. 10 3.5 35.0 61. 5 SLCL 6706-2-6 196-198 OGL No data 6706-2-7 200-205 OGL 8.00 2.66 0. 18 5.7 49.3 45.0 CLSL 6706-2-8 223-224 S-S 2.80 1.04 -0. 10 100.0 0 0 SN 6706-2-9 245-250 OGL 6.36 2.58 -0.17 22.2 46.0 31.8 SSC 6706-2-10 253-254 OGL 6.84 2.94 -0.06 19.3 42.4 38.3 CLSL 6706-2-11 282-283 S-S 2.75 1.03 -0.09 100.0 0 0 SN 6706-2-12 295-302 OGL 6.90 2.51 -0.21 3.8 55.3 40.9 CLSL 6706-2-13 300-315 OGL 7.36 2.22 -0.08 6. 1 50. 9 43.0 CLSL 6706-2-14 320-321 OGL 7.22 2.32 -0.09 11.2 46.3 43. 5 CLSL 6706-2-15 330-335 OGL 6.57 2.51 -0.20 15.8 49. 4 34.8 CLSL 6706-2-16 355-360 OGL 8.22 3.28 0.20 4.3 49.3 46.4 CLSL 6706-2-17 370-375 GL 8.06 1.43 -0.06 3. 1 42.6 54.4 SLCL 199 Appendix 4. (Continued) Depth osua Interval Sample in cores Sediment M + % % Percent Percent Percent Textural Number (cm) Type Sand Silt Clay Classification 6706-2-18 6706-2-19 6706-3-1 6706-3-2 6706-3-3 6706-3-4 6706-3-5 6706-3-6 6706-3-7 6706-3-8 6706-3-9 6706-3-10 6706-3-11 6706-3-12 6706-4-1* 6706-4-2 6706-4-3 6706-4-4 6706-4-5 6706-4-6 6706-5-1 6706-5-2* 6706-5-3 6706-5-4 6706-5-5 6706-5-6 6706-5-7 6706-5-8 6706-5-9 6706-5-10 385-390 390+ 0-5 50-55 100-105 118 150-155 200-205 250-255 300-305 350-355 375-377 377+ 380+ 0-10 60-65 92-97 108-112 145-150 150+ 0-10 25-30 85-90 90-91 93-95 96-97 107-110 115-117 120-121 125-127 GL GL GL GL GL GL GL GL GL GL GL GL GL OGL OGL OGL S-S GL GL OGL OGL OGL S-S S-S S-S OGL OGL S-S OGL 7.44 7.36 6.95 7.11 5.73 7.29 7.80 7.54 8.39 7.71 7.02 8.49 6.84 7.23 7.61 7.53 2.12 7.54 7.67 6.94 6.81 9.51 2.93 2.71 2.62 7.11 6.83 3.81 6.56 2.29 2.62 2.73 2.27 1.43 2.36 2.54 2.47 3.74 2. 96 2.56 4.85 2.69 2.41 2.01 2.62 1.01 2. 10 2.41 2.62 3. 11 5.63 2. 10 1.00 1.21 2. 15 2.53 1.43 2.54 -0.04 -0. 12 -0.07 0.01 0.47 No data 0.07 0.06 0.04 0.47 0.25 0.00 0.62 0.00 -0.08 0.03 -0.02 -0.03 -0.04 -0.08 0.06 0.04 0.53 -0.01 0.04 5.3 4.5 14.6 7.0 1. 9 4.4 5.0 6. 1 6. 1 7.2 11.0 21.0 14.0 4.3 5. 1 2.9 99.7 3.2 3.6 12. 8 14.2 16.7 96.3 94.7 0.03 100.0 -0.08 9.3 0.21 0.05 0. 13 10.0 95. 1 15.7 51.4 50.0 46.4 56.0 85.9 57.0 49. 1 50.8 56.3 51. 9 51.4 44.9 51.0 49.9 41.3 52.6 0.3 41.5 38.3 52.2 48.3 45.3 2. 1 1.7 0 40.2 37.2 3. 1 55.3 43.3 45.6 39.0 37.0 12.2 38.6 45.9 43. 1 37.6 40. 8 37. 5 34.0 35.0 46.8 54.6 44. 5 0 56.3 59. 1 35.0 37. 5 38.0 1.6 3.6 0 50.5 53. 8 1.8 28.9 CLSL CLSL CLSL CLSL SL CLSL CLSL CLSL CLSL CLSL CLSL SSC CLSL CLSL SLCL CLSL SN SLCL SLCL CLSL CLSL CLSL SN SN SN SLCL SLCL SN CLSL 200 Appendix 4. (Continued) Depth osua Interval Sample Number in cores (cm) Sediment Type % cr a+ 9 Percent Sand Percent Silt Percent Clay Textural c Classification 6706-5-11 138-140 OGL 6.71 2.63 0.09 11.-3 44.2 44.5 SLCL 6706-5-12 145-148 S-S 2.92 1.61 0.07 97.3 2.2 0.5 SN 6706-5-13 185-187 OGL 6.47 2.58 0.27 18. 3 55.6 26. 1 CLSL 6706-5-14 207-210 OGL 6.70 2.53 0.43 11.6 61.4 27.0 CLSL 6706-5-15 227-230 OGL 6.82 2.67 0.08 8.3 53.6 38.1 CLSL 6706-5-16 243-245 OGL 7.15 2. 16 -0.03 7.4 41.3 51.3 SLCL 6706-5-17 270-273 OGL 6.93 2.61 0.05 11.6 40.8 47.6 SLCL 6706-5-18 273+ OGL 5.83 2.44 0.31 9.3 51.0 39.7 CLSL 6706-6-1 0-5 OGL 5.50 2. 19 0.54 3.9 41. 1 55.0 SLCL 6706-6-2 15-20 OGL 6.89 2.14 0.04 6.3 61.3 32.4 CLSL 6706-6-3 35-38 OGL 7.20 2.12 0.21 6. 1 60.6 33.3 CLSL 6706-6-4 60-65 OGL 7.09 1.71 0.16 2.2 68.5 29.3 CLSL 6706-6-5 80-83 OGL 6.95 1.35 -0.17 2.6 71.4 25.9 CLSL 6706-6-6 100-103 OGL 6.45 2.34 -0.08 15.0 56.8 28.2 CLSL 6706-6-7 125-127 S-S 3.14 0.66 -0.41 100.0 0 0 SN 6706-6-8 150-155 OGL 8.01 3.39 0.40 4.8 58.0 37.4 CLSL 6706-6-9 190-193 OGL 7.20 1.86 -0.10 11.5 50.4 38. 1 CLSL 6706-6-10 199-201 S-S 3.17 0.67 0.07 100.0 0 0 SN 6706-6-11 250-255 OGL 7.05 2.29 -0.02 7.4 55. 8 36.8 CLSL 6706-6-12 285-288 OGL 6.56 2.09 0. 12 7. 7 67.7 24.6 CLSL 6706-6-13 320-323 OGL 7.50 3.24 0. 18 13. 58 46. 1 40. 4 CLSL 6706-6-14 350-355 OGL 6.93 2. 48 0.04 10. 4 54. 3 35.2 CLSL 6706-6-15 335+ OGL 7. 19 2.65 0.22 9. 2 56.4 34. 4 CLSL 6706-7-1 0-5 S-S 5.25 2. 33 0.81 67. 4 18. 4 14.2 SLSN 6706-7-2 40-45 S-S 10.28 7. 35 0.77 45. 1 16.8 38. 1 CLSN 6706-7-3 70-75 S-S 5.28 2. 39 0.82 63.3 21. 7 15.0 SLSN 6706-7-4 120-125 S-S 5.15 2.41 0.74 65.0 20. 2 14.8 SLSN 6706-7-5 171 S-S 5.23 2.40 0.81 64. 9 20.3 14.8 SLSN 6706-7-6 195-200 S-S 4.78 1. 90 0.76 71.8 16.0 12.2 SLSN 6706-7-7 225-230 S-S 5.01 2.21 0.70 68.2 12. 8 19.0 CLSN 201 Appendix 4. (Continued) a OSU Depth Interval Sample Number in cores (cm) Sediment Type 9 \ •> Percent Sand Percent Silt Percent Clay Textural c Classification 6706-7-8 253 S-S 5.53 2.53 0.78 64.6 18.9 16.5 SLSN 6706-7-9 275-280 S-S 5.18 2.20 0.80 65.7 10.2 24.1 CLSN 6706-7-10 285-290 S-S 4.83 1.92 0.77 63.9 25.4 10.7 SLSN 6706-7-11 315-320 S-S 5.11 2.21 0.80 64.5 23.0 12.5 SLSN 6706-7-12 345-350 S-S 5.61 1.88 0.84 63.4 12.8 23.8 CLSN 6706-7-13 385-390 S-S 5.09 2.21 0.79 66.9 20.1 13.0 SLSN 6706-7-14 410-415 S-S 4.75 1.86 0.83 66.1 25.3 8.5 SLSN 6706-7-15 445-450 S-S 4.96 2.09 0.77 67.8 22.0 10.2 SLSN 6706-7-16 465 S-S 5.32 2.41 0.74 56.2 28.3 15.5 SLSN 6706-8-1* 0-10 S-S 3.64 0.57 0.40 83.9 8.3 7.8 SN 6706-8-2* 35-40 S-S 6.68 3.75 0.81 63.2 16.8 20.1 CLSN 6706-8-3 95-100 S-S 3.40 0.54 0.04 89.7 7.8 2.4 SN 6706-8-4 250-255 OGL 6.23 2.41 -0.32 18.2 51.7 30.1 CLSL 6706-8-5 355-360 OGL 6.13 3.26 0.43 35.5 34.0 30.5 SSC 6706-8-6 375 OGL 6.55 2.60 -0.29 17.3 45.0 37.7 CLSL 6708-20-1 0-5 S-S 3.82 1.77 0.31 83.6 9.0 7.8 SN 6708-23-A 4-6 S-S 4.52 2.16 0.59 51.6 36.3 12.1 SLSN 6708-23-B 50-52 S-S 5.49 2.21 0.38 42.9 42.4 14.7 SLSN 6708-23-C 75-77 S-S 4.31 1.98 0.51 48.9 39.2 11.9 SLSN 6708-23-D 98-100 S-S 4.29 2.38 0.59 52.8 34.8 12.4 SLSN 6708-23-E 124-126 S-S 4.93 2.31 0.37 43.9 41.8 14.3 SLSN 6708-23-F 168-170 S-S 5.28 2.15 0.31 36.0 49.0 15.0 SNSL 6708-23-G 200-202 S-S 4.36 2.17 0.39 53.9 34.7 11.4 SLSN 6708-23-H 205-207 S-S 4.71 2.98 0.57 52.8 36.7 10.5 SLSN 6708-23-1 213-215 S-S 5.62 1.99 0.10 20.2 62.8 17.0 SNSL 6708-23-J 241-243 S-S 5.41 1.89 0.13 19.3 66.0 14.1 SNSL 6708-25-1 0-5 OGL 7.18 1.18 -0.07 1.7 70.0 28.2 CLSL 6708-25-2 75-80 OGL 6. 70 1. 95 -0. 29 8. 5 58. 9 32. 7 CLSL 6708-25-3 150-155 OGL 7.17 2.22 -0.13 3.5 55.4 41.1 CLSL 6708-25-4 225-230 OGL 6.62 2.01 0.09 6.0 68.7 25.3 CLSL Appendix 4. (Continued) 202 Depth OSU Interval Sample i in cores ! sediment M* 0" 4> a* Percent Percent Percent Number (cm) Type Sand Silt Clay c Classification 6708-25-5 287 OGL 7.65 3.42 0.52 12.8 67.6 19.6 CLSL 6708-25-6 320-325 OGL 7.38 2.06 0.09 2.8 59.0 38. 2 CLSL 6708-25-7 350-355 OGL 6.63 1.91 -0.33 3.6 64. 9 31.4 CLSL 6708-25-8 395-400 OGL 6. 13 1.62 -0.24 9.6 81. 9 8.5 SL 6708-25-9 420-425 OGL 6.89 2.53 0.01 5.3 59.4 35.3 CLSL 6708-25-10 440-445 OGL 6.81 2.29 0.08 7.3 56. 1 36.6 CLSL 6708-25-11 460-465 OGL 6.86 2.09 -0.41 9.0 48.3 42.7 CLSL 6708-25-12 480-485 OGL 6.41 2.48 -0.36 17.0 47.2 35.8 CLSL 6708-25-13 500-505 OGL 8.01 3.28 -0.21 10. 9 29. 5 59. 5 SLCL 6708-25-14 520-525 OGL 7.95 3.98 0.64 16. 2 46.8 37.0 CLSL 6708-25-15 540-545 OGL 7.22 3.29 -0.12 18. 2 35.8 46.0 SLCL 6708-25-16 578-580 OGL 7.24 3.35 0.03 17.0 41. 1 41. 9 SLCL 6708-25-17 610 OGL 7.88 3. 95 -0. 17 17. 5 26. 5 56.0 SLCL 6708-33A 1-3 S-S 3.48 0.67 0.48 81.8 12.4 5.8 SN 6708-33B 21-23 S-S 4.03 1. 26 0.73 77. 9 2. 1 7.0 SN 6708-33C 41-43 S-S 2.97 0.31 0.22 91. 1 10. 9 6.8 SN 6708-33D 61-63 S-S 3.47 0.70 0.54 82. 4 27.3 6.7 SN 6708-37-1 10-15 OGL 7.41 2.08 0. 15 2. 1 60. 1 37.7 CLSL 6708-37-2 100-105 OGL 7.68 2.33 0.31 2.5 58.8 38.7 CLSL 6708-37-3 200-205 OGL 7. 14 2.21 0. 12 4.0 61.6 34. 5 CLSL 6708-37-4 300-305 OGL 7.42 3. 12 0.46 2.5 61. 7 35. 9 CLSL 6708-37-5 400-405 GL 7.39 2.48 0.20 1.6 61.0 37. 5 CLSL 6708-37-6 505-510 GL 7.41 2.30 0.17 3.2 58.7 38. 1 CLSL 6708-37-7 570-575 GL 7.59 2.08 0.08 1. 1 57.0 41.9 CLSL 6708-37-8 583 GL 7.54 2.45 0.22 1.0 59.9 39. 1 CLSL 6708-38-1 0-5 OGL 8.44 2.32 0.06 0.5 44.8 54.7 SLCL 6708-38-3 55-60 OGL 8.33 2.80 0.17 1. 1 50. 1 48.9 CLSL 6708-38-5 100-105 OGL 7.41 2.28 0.26 2. 3 61. 9 35.7 CLSL 6708-38-7 150-155 OGL 7.96 2.33 0.11 1. 1 52.4 46.5 CLSL 6708-38-9 200-205 OGL 7.56 2.70 0.08 4.6 52. 5 42.8 CLSL Appendix 4. (Continued) 203 Depth OSU Interval Sample in cores S Number (cm) ediment Type M % a+ Percent Sand Percent sat Percent Clay Textural c Classification 6708-38-11 250-255 OCL 8.52 2.97 0.39 1.2 53. 8 45.0 CLSL 6708-38-13 300-305 OGL 8.23 2. 93 0.42 1. 5 56. 5 42.0 CLSL 6708-38-15 350-355 OGL 7.40 2.45 0.04 1.3 57.6 41. 1 CLSL 6708-38-17 400-405 OGL 10.63 5.37 0.60 1.8 50.4 47.8 CLSL 6708-38-19 450-455 OGL 7.53 2.31 0.18 1.2 59.4 39.4 CLSL 6708-38-21 469-472 OGL 9.93 4.84 0.51 0.7 51.6 47.7 CLSL 6708-38-22 472+ OGL 10.68 5.32 -1.0 1.3 40.2 58.5 SLCL 6708-39-1* 0-10 S-S 5.87 3.03 0.65 55.7 23.2 21. 1 SLSN 6708-39-2* 25-30 S-S 6.57 3. 17 0.21 30.9 39.2 29.9 SSC 6708-39-3 65-70 S-S 4.77 2. 59 0.58 72.3 11. 9 15.7 CLSN 6708-39-4 113-115 S-S 4.93 2.42 0.41 70.9 8.0 21. 1 CLSN 6708-39-5 125-130 S-S 7.33 2. 16 -0.04 4.6 54.6 40.8 CLSL 6708-39-6 130-135 S-S 3.62 0.37 0.05 84. 5 6.2 9.3 SN 6708-39-7 180-185 S-S 8.48 5. 11 0.76 35. 8 39.8 24.4 SSC 6708-39-8 230-235 S-S 7.39 3.37 0.39 15.6 50.4 34.0 CLSN 6708-39-9 280-285 S-S 3.87 0.56 0.41 82.7 7.2 10.1 SN 6708-39-10 330-335 S-S 3.12 0.20 -0.18 73.8 6.0 20.2 CLSN 6708-39-11 380-385 S-S 4.85 1.52 0.74 68.4 17.3 14.3 SLSN 6708-39-12 412-417 S-S 9.85 6. 15 0.73 22.4 43.1 34.5 SSC 6708-39-13 427 S-S 5.83 3.07 0.87 67. 5 15.6 16.8 CLSN 6708-40-1 0-5 S-S 8.67 5. 32 0.62 46. 1 13. 9 40.0 CLSN 6708-40-2 30-35 S-S 9. 71 6. 25 0.24 29. 1 19.7 51.2 SNCL 6708-40-3 40-42 S-S 5.02 1. 94 0.78 72. 9 16. 5 10.6 SLSN 6708-40-4 87-92 S-S 9.62 6.38 0.73 45.6 13. 3 41. 1 CLSN 6708-40-5 123-127 S-S 8.59 5.23 0.63 46.0 15. 4 38.6 CLSN 6708-42-1 225-230 S-S 3.42 0.62 0.21 93.6 4.8 1.6 SN 6708-42-2 434-454 S-S 3. 10 0. 31 0. 11 98.8 1.2 0 SN 6708-43A 0-10 S-S 3.15 0.36 0.45 89.7 7.8 2.5 SN 6708-43B 10-20 S-S 3. 12 0.33 0.05 93.8 4.2 2.0 SN 6708-43C 20-30 S-S 3.26 0.38 0. 17 89.4 7. 7 2.9 SN 204 Appendix 4. (Continued) Depth osua ] Interval Sample Number in cores S (cm) ediment Type M 0" Q* Percent Sand Percent Silt Percent Clay Textural c Classification 6708-43D 30-40 S-S 3.23 0.41 0.21 90. 9 7.4 1.7 SN 6711-1-1 0-5 OGL 6.67 2.20 -0.08 0.2 68.6 31.2 CLSL 6711-1-2 45-50 OCL 7. 19 2.44 -0.09 0.3 58. 2 41. 5 CLSL 6711-1-3 95-100 OGL 6.68 1. 50 -0.30 0.3 76.7 23. 1 SL 6711-1-4 145-150 OGL 7.63 1. 78 -0.28 0.2 58.9 40. 9 CLSL 6711-1-5 195-200 OGL 8.20 2.21 -0.01 0. 1 45. 9 54. 1 SLCL 6711-1-6 248-252 OGL 6.66 1. 98 -0.01 0. 1 70. 9 29.0 CLSL 6711-1-7 295-300 OGL 6.46 1.70 -0.03 0. 5 78. 9 20.6 SL 6711-1-8 345-350 OGL 8.32 2.89 0.27 0.3 53. 7 46.0 CLSL 6711-1-9 363 OGL 6.43 2.06 0.04 5.6 71.0 23.4 CLSL 6711-1-10 375-378 OGL 6.31 1.84 0.37 0.8 81.0 18. 2 SL 6711-1-11 385-388 OGL 8.30 4.46 0.05 18. 9 30. 1 51.0 SLCL 6711-1-12 390-392 OGL 7.34 1. 75 -0.04 0. 9 60. 2 38. 9 CLSL 6711-1-13 397-399 OGL 6.92 1.69 -0. 13 0.7 68. 5 30.8 CLSL 6711-1-14 404-408 OGL 9.36 3.42 -0.03 0. 8 33.9 65.3 SLCL 6711-2-1 0-5 OGL 5.98 4.45 -0.42 17.6 33.6 48.8 SLCL 6711-2-2 25-30 OGL 9.77 6.23 0.32 16. 3 34.0 49.7 SLCL 6711-2-3 50-55 OGL 7.78 2.64 0.01 6.3 46.6 47. 1 SLCL 6711-2-4 72-80 OGL 7.28 2. 19 0.04 8. 1 47.6 44.3 CLSL 6711-2-5 103-108 OGL 8.01 2.42 0.21 7.3 50.0 42.7 CLSL 6711-2-6 125-130 OGL 8.22 1.99 0.30 6.8 40.2 53.0 SLCL 6711-2-7 135-145 OGL 7.24 2.02 -0.23 8.8 46.9 44. 4 CLSL 6711-2-8 180-185 OGL 8.26 1. 87 -0.05 4. 1 38.6 57. 2 SLCL 6711-2-9 217-223 OGL 8.31 2.30 0.03 0. 9 45. 2 53. 9 SLCL 6711-2-10 260-265 OGL 8.60 3.68 0.24 0.4 51.6 48. 1 CLSL 6711-2-11 280-285 OGL 8.81 3.09 0. 14 0.5 45.6 53. 9 SLCL 6711-2-12 305-313 OGL 8.74 3.48 0.06 0.7 44.2 55. 1 SLCL 6711-2-13 320-324 OGL 8.00 2.48 -0.01 1.0 48. 2 50.8 SLCL 6711-2-14 342-348 OGL 8.55 3.05 0.28 1.3 51. 2 47. 5 CLSL 6711-2-15 360-365 OGL 10.96 5.04 0.34 0. 9 44.4 54.6 SLCL 205 Appendix 4. (Continued) Depth OSU Interval Sample in cores S ediment M U q Percent Percent Percent Textural Number (cm) Type 4> 4> 4> Sand Silt Clay c Classification 6711-2-16 387-393 OGL 8. 17 3.25 -0. 10 1.6 42. 5 56.0 SLCL 6711-2-17 415-420 OGL 8.50 2.45 0. 14 2. 2 45.6 52.3 SLCL 6711-2-18 430 OGL 9.07 2. 74 0.05 0. 9 37.5 61.6 SLCL 6711-5-1 0-5 OGL 7.84 1.61 -0.03 0.8 50. 9 48. 3 CLSL 6711-5-2 35-40 OGL 8.22 1.81 -0.08 0.4 41.6 58.0 SLCL 6711-5-3 70-75 OGL 8.37 2.55 0. 24 0. 5 51. 9 47. 7 CLSL 6711-5-4 105-110 OGL 8.45 1.80 -0.04 0.6 38.6 60.8 SLCL 6711-5-5 140-145 OGL 8.89 3. 35 0. 15 0.2 46. 2 53.6 SLCL 6711-5-6 175-180 OGL 8.93 3. 18 0.06 0.3 41. 9 57. 7 SLCL 6711-5-7 195-200 OGL 7.93 1.42 -0.07 0.3 48.3 51. 4 SLCL 6711-5-8 210-215 OGL 8. 15 1. 93 0. 10 0.5 49.8 49. 7 CLSL 6711-5-9 250-255 OGL 9.43 3.22 0.22 0. 1 43.2 56.6 SLCL 6711-5-10 280-285 OGL 10.22 4.66 0.25 0.2 43.4 56.3 SLCL 6711-5-11 315 OGL No data 6711-5-12 315-320 OGL 8.20 1.99 -0.03 0. 1 44.6 55.3 SLCL 6711-5-13 350-355 OGL 8.80 2. 94 0. 10 0. 1 44. 1 55. 8 SLCL 6711-5-14 375-380 OGL 8.40 1. 99 -0.03 0. 1 41.0 58.9 SLCL 6711-5-15 410-415 OGL 10.78 4.58 0.24 0.2 39.6 60. 2 SLCL 6711-5-16 445-450 OGL 8.03 2. 15 0.00 4.4 44.8 50.8 SLCL 6711-5-17 475-480 OGL 8.35 2.03 0.09 0.6 46.3 53. 1 SLCL 6711-5-18 490 OGL No data 6711-5-19 500-516 OGL 8.36 2.43 0. 14 2.4 46.9 50.7 SLCL 6711-5-20 516+ OGL 10.28 4. 18 0.57 1.0 49.4 49. 7 SLCL 6711-6-1 0-10 OGL 10.85 5. 15 -0.02 3.7 29.0 67.3 SLCL 6711-6-2 50-55 OGL 11. 16 4.84 -1.00 0.2 22.3 77. 5 CL 6711-6-3 100-105 OGL 11.77 4.23 -0.66 0.8 17.3 81. 9 CL 6711-6-4 150-155 OGL 11.93 4.07 -0.22 1. 5 15.2 83.2 CL 6711-6-5 200-205 OGL 11.41 4. 59 -1.00 0.5 19. 3 80. 1 CL 6711-6-6 255-260 S-S 7.63 3. 79 -0.36 16.4 19.2 64. 4 SLCL 6711-6-7 335-340 OGL 10.34 4. 25 -0.03 1. 5 27.7 70.8 SLCL 206 Appendix 4. (Continued) Depth osua Interval Sample in cores S ediment M 4> 4> Sand Silt Clay c Classification 6711-6-8 340+ OGL 12.35 3.65 -1.00 1. 5 11.3 87.2 CL 6711-8-1 0-5 OGL 8.22 2.58 0.07 0.3 48. 7 51.0 SLCL 6711-8-2 50-55 OGL 7.82 3. 10 0. 10 0.3 54.3 45.4 CLSL 6711-8-3 100-105 OGL 8.83 2.72 0.42 0.2 52. 2 47.5 CLSL 6711-8-4 150-155 OGL 8.91 2.92 0.60 0. 3 55.6 44. 1 CLSL 6711-8-5 200-205 OGL 7.25 1.75 -0.27 0. 2 56.3 43. 5 CLSL 6711-8-6 245-250 OGL 7.45 1. 92 0.03 0.3 59.6 40. 1 CLSL 6711-8-7 300-305 OGL 7.02 2.36 -0. 17 0.2 59.2 40.6 CLSL 6711-8-8 350-355 OGL 8.92 3.66 0.44 0.6 53.7 45. 7 CLSL 6711-8-9 400-405 OGL 10.69 5. 31 0.72 1. 4 54.0 44.6 CLSL 6711-8-10 410-412 S-S 5.20 2.22 0.39 37.7 49. 7 12.6 CLSL 6711-8-11 415-416 OGL 7.53 3.57 0.20 17. 1 43. 4 39.5 CLSL 6711-8-12 417-419 OGL 9.05 4. 26 0.35 3.0 49. 5 47. 5 CLSL 6711-8-13 430-431 OGL 11.84 4. 16 -0.71 2.4 15. 3 82.3 SLCL 6711-8-14 450-455 OGL 8. 14 2.38 0.02 1. 1 47.2 51. 7 SLCL 6711-8-15 500-505 OGL 7.43 1. 72 -0.06 1. 8 57. 3 40. 9 CLSL 6711-8-16 555-560 OGL 10.95 5.05 -1.00 0.7 39. 1 60.2 SLCL 6711-8-16 560+ OGL 9.73 4.02 0.33 0.4 46.9 52. 8 SLCL Asterisk after sample number indicates Phleger core sample from multiple corer trip weight; all other samples are from piston cores. b M+)= Phi Mean Diameter; 0"^ = Phi Standard Deviation (Sorting); qi - Phi Skewness; computed from equations of Inman (1952). "SN = sand; SL ~ silt; CL = clay; SSC ~ sand-silt-clay; SNCL = sandy clay; SLCL = silty clay; SLSN = silty sand; CLSN = clayey sand; SNSL = sandy silt; CLSL = clayey silt; textural classification after Shepard (1954). 0) o o o o c o < > o £ V v IS T3 to QJ nj [i, E tH s & tn 2 "5 b * t; rt 3 o a 1 rt g & § bO J2 M O £ B, g 6 on 2 207 m m -* ro CM ro H CM r\j O in in ro CM m CM r\) O ro § 00 ■* 00 ro «5 fM in k CO ro o- ro O vo ■* ■* O m" m* I —* d -« t^ rg ^r on rg' ro rg" f_ o\ T — T T m *r — ■ vo rC m* rg' rg' oorvQ O on vo on r-v rr vo ^T rg rg -h" ro" ON* ^* -a" oo —« r*. [_ ON* ro" vo' [_ Ol CM 'O O rg on ro co <*" T* d rW d o o H O Id -i « N rf vo 0% rg* rs." ro ro" oo" f_ r»! X on co ■* **■ vo ** m o m m on o "t" m" vo' — «" m" c\ rg* rg* — . — vo ^r | oo — » co r*. oo vo vo vo ro -r -r o m <* rg d -* ON* —I -h" 8.3 * "S — a. ~ rt n c« ri + o (4 e a « fo 2 i: i! 7? c 00 o TJ 0) rt •J 7 8 £ u & -j 0 < Q < £ ■-■' H 0- < •3 « * + C q. 3 X X -3 •Bg = U3.»S-0.3gS8S! j « ? : s « i ° <* ^ >- s a 3 E i S j .S u : » g | . = .^ u O 0 o- 0 a 9 ■a 3 I 2 2 J! £ E 0 » :; e u a "3 0 r^ 1 '3 3 o J3 :> + ■r 6 3 0 O c -1 u B i a OS to s c I F 0 3I-9E-80Z9 3I-SE-80Z9 9W"90Z9 SS§ X to I-Z-90Z9 U jaquinu 3ldlUE5 o ^ s — tf ^ . -^ -^ m ro O — • Ot 00 "1 1^ O N t O * CO* ^ -■* ■*" -I ^1 Oi N O N O IN ^r N -> o H i i i H O VD ■* (M CO iO « i O lO CO Oi O CTi CO t^ a\ m oi CO CO cd rg *o . rg m r^ vo rg r^ in io oS i/" l^ ro' o rg* a" oi oo u m 6 n d rg j *D io oo ■<*• in o m o O «o ro" r\i -" vo i ~" 00 IO ro CO (_ "** O ^ ^ oo S. -i oo ■*" N N [- I* [- (-, O -h "* -? in d * ! o O *** 0"v CO — * io O o\ N t in co io co in rg' ■* d ro" cd I *D ■* oo ■* (_ O* \o' ro" rC ro" to irt co CT\ [_ io" N S X 0 u a >. *-> a g 3 1/3 < Physiographic province 0) ■M ■ H O 43 O 2? ■t-> a o •i-i i-i o s £3 O S 2? Chlorite/ illite (C/I) ratio Other minerals present 6711-2-1 5 OGL H LS 62 30 8 2. 1 6711-2-12 313 OGL H LS 58 41 1 1.4 6706-3-1 5 CL P KP 57 41 2 1.4 amphiboles 6706-3-10 377 GL P KP 59 39 2 1.5 amphiboles 6711-6-1 10 OGL H LRC(LS) 52 43 5 1.2 6711-6-8 345 OGL H LRC(LS) 49 49 2 0.9 6708-37-1 15 OGL LP URC 71 26 3 2.7 6708-37-7 575 GL LP URC 56 43 1 1.3 6708-25-4 230 OGL H SH 53 43 4 1.3 amphiboles 6708-25-16 580 OGL H SH 62 37 1. 1.7 amphiboles 6706-1-1 5 OGL H US 61 36 3 1.7 6706-1-11 430 GL LP US 63 35 2 1.7 6706-2-12 302 OGL LP MB 57 40 3 1.4 6706-2-17 375 GL LP MB 56 42 2 1.3 6708-38-1 5 OGL H LS 64 33 2 1.9 6708-38-21 472 OGL H LS 58 40 2 1. 5 6706-6-8 155 OGL H URCW(US) 69 29 2 2.4 amphiboles 6706-6-15 355 OGL H URCW(US) 68 32 1 2. 1 amphiboles 6802-D3-1 a c d LT LS 67 31 2 2. 1 amphiboles 6802-D2 a c H CBB 64 32 4 2.0 amphiboles 6706-3-11-2 377+b c e LT KP 78 20 2 3.9 amphiboles Sample from dredge haul. Rock fragment from piston core cutting head. '6802-D3-1 = Mudstone; 6802-D2 = Mud; 6706-3-11-2 = Mudstone. Early Pliocene. Mid-Pliocene. 212 bC O 01 3 •iH s a •i-i rt J- o JS 0) PS g 0 rt a! U rt 01 E ai 0. rt 13 8 CO SI u TS < ai o -C ro 5 a d o oi L0 0 Ei) PS rt o « .2 a no ■— ' PL to A s X 1) B o •M o M nl in 0- s 4-' 3 a cj o M 'rt LO CD CL bO H "3 o oj c« in 0 4= t^ aj rt P< o 0) w E £ rt 3 oo 3 rt "3 ° rt n bO >- IT -3 QJ '3 43 K o> s •IH rt bo +j 3 t, O in co CM CO 5 I fM CM 00 O fv 1 S E xT e 01 0) XI s 3 O I =1 01 ■8 rt ■a in o CM in ro in CM ro ro ro I 00 s o "cL 3 ST «» o ^ ^ ^3 +J rt 3 01 1 3 o E o 3 • l-( 4- t/l rt rt 3 3 0 0) 01 M l 3 u 75 c« oi •iH a o H o u o s V 8 o CO tx rO O 5 a 0 c .3 t, t-i 01 01 -3 3 -3 O X CO bO rt 4) P 3 0 a 4-> nl tL rt a) 4-1 5 m m ro m 0) 3 > o> •** Si i-j (X O 0) E •- h 3 b 8 ^ 3 rt rt ... 3 •!-< O 3 4-> O rt nl p o» bO +i rt rt 5 rt O O 01 o rt 8 0) o rt 8 o CM m CTi in ^ ~-l ~* rt C\] m o ro m in ro 13 0) rt O CT -a H 01 D T3 D 3 O) 3 a o ai 3 rt « oi Bo ^ +J t. E ~ 3 CT1 11 3 -3 u rt il feb b a -a v 2 -a rt r 03 3 O O 3 5 Si 3 T3 T3 U XI 3 3 O Wj 11 -rt J2 3 xi E O rt n. in O o Ui ^^ m XI rt 3 bO rt D 8 in vo in ^-H o ffv o m 3 . 1 rt r-H o a o o c D o o T3 1 o XI T3 4-> rt *0 QJ a -3 o >• C rt ID bO O o 01 o *-> . to X) 3 o 4-> bO •a 3 .— 1 o 0) (0 c c b o o n +J ■4-J 0 CO ca T3 T3 XI T3 rt 3 3 3 3 j 2 2 s 2 U OJ VO vo 214 Appendix 8. (Continued) NOTES a Other--includes: heavy minerals, authigenic minerals, opaque grains and all unidentified grains. b , Nomenclature assigned to all samples examined petrographically according to the classification of impure sandstones of Williams, Turner and Gilbert (1954). c Shell fragments comprised 25% of haul. d , , Number in parentheses indicates a re-computed percentage, wherein only quartz, feldspar and rock fragments are assumed to constitute 100% of the sample. e All samples of this rock type contain < 10% grains, and are too fine-grained for petrographic examination; identification based on megascopic and binocular-microscopic examination. Samples are from rock fragments in basal portion and cutting head of piston core 6706-3; 50% of fragments are siltstone and 50% are mudstone. AN ABSTRACT OF THE THESIS OF JOSEPH JOHN SPIGAI for the DOCTOR OF PHILOSOPHY (Name) (Degree) in OCEANOGRAPHY presented on (Major) (Date) Title: MARINE GEOLOGY OF THE CONTINENTAL MARGIN OFF SOUTHERN OREGON Abstract approved: L. D. Kulm The continental margin off southern Oregon, which includes the shelf and slope from Cape Blanco to the Oregon-California border, exhibits a distinctive marginal-plateau structural pattern which divides the margin into the continental shelf, the upper continental slope and its associated benches, and the lower continental slope. Lutum trans- port and deposition have dominated the sedimentary processes on the margin since the start of Holocene time. The structure of the southern Oregon margin is characterized by north-south trending compressional folds, and near-vertical faults which have been down-dropped to the west. Large-scale folds on the upper slope have ponded sediments behind them resulting in the forma- tion of the Klamath Plateau, Cape Blanco Bench, and other bench-like features. Development of the structural pattern is most likely a result of the compressive underthrusting of the oceanic lithospheric plate beneath the southernmost Oregon-northern California margin and the crustal extension which exists throughout the nearby continent and ocean basin. Useful stratigraphic horizons within the late Pleistocene and Holocene margin deposits include Mazama Ash (6600 years B. P. ) and several recognizable shifts in the abundance of Radiolaria and plank- tonic Foraminifera, particularly one dating from 5000-4000 years B. P. Holocene sedimentation rates vary from an average of 10 cm/ 1000 years on the upper slope to an average of 50 cm/1000 years on the lower slope, indicating that the lower slope is out-building and up- building more rapidly than the upper slope. The paleo-depth range of Pliocene fauna in sedimentary rocks from the margin suggests that subsequent to their deposition both uplift and subsidence occurred on the southern Oregon margin. Sediments from the southern Oregon margin consist primarily of olive gray lutite, gray lutite, and sand-silt layers. Olive gray lutite is Holocene in age and is ubiquitous on the margin, with the thickest accumulation (10 m average) found on the lower slope, while the distribution of Holocene lutite on the upper slope is thin and patchy (3-4 m or less). The gray lutite appears to be a late Pleistocene deposit, and the sand-silt layers reflect both ages. The surface sedi- ment distribution pattern on the shelf consists of modern inner shelf sand, modern central shelf mud, and mixed deposits of both types. Relict deposits are present at the shelf edge. The lower slope consists entirely of modern mud, but the surface sediment on the upper slope and benches consists of both modern and relict deposits, and mixtures of the two. The mineralogy of the unconsolidated and consolidated sediments from the margin indicates that the Klamath Mountains have been the dominant source for these deposits since early Tertiary time. This is reflected in the abundance of blue-green hornblende and other heavy minerals indicative of the Mesozoic rocks of the Klamath Mountains; the same source is suggested for the abundant chlorite found in the clay fraction of margin sediments and rocks. There are indications in the mineralogy of lower slope sediments which suggest that the Tertiary strata of the southern Oregon Coast Ranges may be a secondary source for the deposits in this environment. When compared to the upper slope sediments, those from the lower slope have a higher feldspar content, a higher pyroxene -to-amphibole ratio, and an apparently higher illite content. As a result of the Holocene rise in sea level, the deposition of coarse elastics on the southern Oregon margin has been restricted to the inner shelf. Consequently, only the fine-grained lutum discharged from rivers is deposited on the outer margin environments. Sub- marine topography, oceanographic conditions, and gravity are impor- tant factors which effect transport and deposition of lutum on the mar- gin. A model of modern lutum transport by bottom turbid layers and fine -particle suspensate is proposed for the southern Oregon margin. Long-period swell is believed to be responsible for much of the formation of bottom turbid layers on the shelf. Once formed, these turbid layers move north and west over the shelf under the influence of shelf currents, alternating tidal action, and gravity; upon reaching the slope they are funneled into submarine valleys and deposited on the lower slope and adjacent deep sea. Lutum deposited on the upper slope is eventually re-suspended and transported by southerly bottom cur- rents into down-slope valleys; very little lutum remains behind on the upper slope. Deposition of the fine-particle suspensate as well as slumping and other gravitational processes contribute to the lower slope sediments. The end result of modern lutum transport is the continual up-building and out-building of the lower slope. CASE BINDER ^^3 Syracuse, N. V. Stockton, Calif. SpVgai Thesis II S66845 Spigai Marine aeoloav of thp continental margin off southern Oregon. thesS66845 Marine geology of the continent* margin 3 2768 002 01580 2 DUDLEY KNOX LIBRARY