Sample records for zaramella andrea marion

  1. Drill Bit Noise Illuminates the San Andreas Fault

    NASA Astrophysics Data System (ADS)

    Vasconcelos, Ivan; Snieder, Roel; Sava, Paul; Taylor, Tom; Malin, Peter; Chavarria, Andres

    2008-09-01

    Extracting the vibration response of the subsurface from noise is a rapidly growing field of research [Curtis et al., 2006; Larose et al., 2006]. We carried out broadside imaging of the San Andreas fault zone (SAFZ) using drill bit noise created in the main hole of the San Andreas Fault Observatory at Depth (SAFOD), near Parkfield, Calif. Imaging with drill bit noise is not new, but it traditionally requires the measurement of the vibrations of the drill stem [Rector and Marion, 1991]; such measurements provide the waves radiated by the drill bit. At SAFOD, these measurements were not available due to the absence of an accelerometer mounted on the drill stem. For this reason, the new technique of deconvolution interferometry was used [Vasconcelos and Snieder, 2008]. This technique extracts the waves propagating between seismometers from recordings of incoherent noise.

  2. General perspective view of the Marion Creek Bridge, view looking ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    General perspective view of the Marion Creek Bridge, view looking southwest. - Marion Creek Bridge, Spanning Marion Creek at Milepoint 66.42 on North Santiam Highway (OR-22), Marion Forks, Linn County, OR

  3. General perspective view of the Marion Creek Bridge, view looking ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    General perspective view of the Marion Creek Bridge, view looking southeast. - Marion Creek Bridge, Spanning Marion Creek at Milepoint 66.42 on North Santiam Highway (OR-22), Marion Forks, Linn County, OR

  4. Topographic view of the Marion Creek Bridge, view looking westbound ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    Topographic view of the Marion Creek Bridge, view looking westbound on the Santiam Highway. - Marion Creek Bridge, Spanning Marion Creek at Milepoint 66.42 on North Santiam Highway (OR-22), Marion Forks, Linn County, OR

  5. Marion McGregor Lee Loy

    ERIC Educational Resources Information Center

    Rossi, Joe

    2007-01-01

    This article presents an interview with Marion Frances Kaleleonalani McGregor Lee Loy who served as a teacher in the Hawai'i Department of Education from 1935 to 1974. Marion McGregor Lee Loy was born in 1911 in Honolulu. She attended Central Grammar and Lincoln Grammar schools before entering Kamehameha School for Girls in the ninth grade. Lee…

  6. Francis Marion National Forest forest plan revision - ecosystems & restoration needs

    Treesearch

    Mark Danaher

    2016-01-01

    The Forest Service is currently revising the previous 1995 Forest Plan for the Francis Marion National Forest in Coastal South Carolina developed in the wake of Hurricane Hugo which devastated the forest in 1989. Since 1995, the human communities surrounding the Francis Marion National Forest have grown and changed significantly. The revised Francis Marion Forest Plan...

  7. Distribution of Volatile Composition in 'Marion' (Rubus Species Hyb) Blackberry Pedigree

    USDA-ARS?s Scientific Manuscript database

    Fruit of 'Marion', as well as those of 'Marion's parents, grandparents, great grandparents and great-great grandparents that were available were evaluated for the volatiles that impact flavor. These various parents include blackberries from several species, raspberry and raspberry-blackberry hybrid...

  8. Sediment transport and deposition in Lakes Marion and Moultrie, South Carolina, 1942-85

    USGS Publications Warehouse

    Patterson, G.G.; Cooney, T.W.; Harvey, R.M.

    1996-01-01

    Lakes Marion and Moultrie, two large reservoirs in the South Carolina Coastal Plain, receive large inflows of sediment from the Santee River. The average rate of sediment deposition for both lakes during the period 1942-85 was about 0.06 inch per year, or about 800 acre-feet per year. The rate during 1983-85 was about 0.037 inch per year, or about 490 acre-feet per year, reflecting the decreasing trend in sediment inflow. This is a reversal of a trend toward increasing suspended- sediment concentrations in streams that were caused by farming practices in the southern Piedmont from about 1800 to about 1920. Only a small part of the eroded sediment has been carried out of the Piedmont, but the remaining sediment is becoming less available for transport. Sediment deposition is concentrated in several areas of upper Lake Marion where the velocity of the incoming water decreases significantly. Beds of aquatic macrophytes appear to encourage deposition which, in turn, creates favorable habitat for the plants. The rate of sediment accumulation in Lakes Marion and Moultrie averaged 650,000 tons per year during 1983-85, reflecting a trap efficiency of 79 percent of the total sediment inflow of 825,000 tons per year. Thickness of post-impoundment sediment varies from about 11 feet near the mouth of the Santee River in Lake Marion to 0 feet in Lake Moultrie near Bonneau. Sediments in Lake Marion tend to have finer texture and higher contents of organic matter, nutrients, and trace metals than those in Lake Moultrie.

  9. Iris Marion Young's Imaginations of Gift Giving: Some Implications for the Teacher and the Student

    ERIC Educational Resources Information Center

    Galea, Simone

    2006-01-01

    The paper discusses Iris Marion Young's idea of asymmetric reciprocity that rethinks typical understandings of gift giving. Iris Marion Young's proposals for asymmetric ethical relationships have important implications for democratic contexts that seek to take differences seriously. Imagining oneself in the place of the other or expecting from the…

  10. NASA Sees Heavy Rainfall in Tropical Storm Andrea

    NASA Image and Video Library

    2017-12-08

    This NOAA GOES-East satellite animation shows the development of System 91L into Tropical Storm Andrea over the course of 3 days from June 4 to June 6, just after Andrea was officially designated a tropical storm. Credit: NASA's GOES Project --- NASA Sees Heavy Rainfall in Tropical Storm Andrea NASA’s TRMM satellite passed over Tropical Storm Andrea right after it was named, while NASA’s Terra satellite captured a visible image of the storm’s reach hours beforehand. TRMM measures rainfall from space and saw that rainfall rates in the southern part of the storm was falling at almost 5 inches per hour. NASA’s Terra satellite passed over Tropical Storm Andrea on June 5 at 16:25 UTC (12:25 p.m. EDT) and the Moderate Resolution Imaging Spectroradiometer or MODIS instrument, captured a visible image of the storm. At that time, Andrea’s clouds had already extended over more than half of Florida. At 8 p.m. EDT on Wednesday, June 5, System 91L became the first tropical storm of the Atlantic Ocean hurricane season. Tropical Storm Andrea was centered near 25.5 North and 86.5 West, about 300 miles (485 km) southwest of Tampa, Fla. At the time Andrea intensified into a tropical storm, its maximum sustained winds were near 40 mph (65 kph). Full updates can be found at NASA's Hurricane page: www.nasa.gov/hurricane Rob Gutro NASA’s Goddard Space Flight Center

  11. Andrea Watson | NREL

    Science.gov Websites

    energy solutions for emission mitigation, international climate change strategies, and renewable energy technical decision making. Andrea's expertise lies in strategic planning, change strategies, and decision

  12. 77 FR 771 - Proposed Establishment of Class E Airspace; Marion, AL

    Federal Register 2010, 2011, 2012, 2013, 2014

    2012-01-06

    ...) Global Positioning System (GPS) Standard Instrument Approach Procedures at Vaiden Field. This action... airspace at Marion, AL, providing the controlled airspace required to support the new RNAV GPS standard...

  13. A Cultural Resources Survey of the River Trace Permit Area Marion, Crittenden County, Arkansas

    DTIC Science & Technology

    1990-09-01

    located near the City of Marion, Crittenden County, Arkansas. The tract is within the direct impact area of a proposed lake that will be part of the River...channel. The area is presently situated in a swale between low ridges, and was part of Marion Lake until the early twentieth century. Properties...PRPOSED LAKE ................... B M N225 ~\\\\ \\\\r~~i N * ’~ 7~;777 ........o 7____ *, -T , .1 contour interval fee h~t 0 Mile PROJET ARE Map 0

  14. Another Vision of Progressivism: Marion Richardson's Triumph and Tragedy.

    ERIC Educational Resources Information Center

    Smith, Peter

    1996-01-01

    Profiles the career and contributions of English art teacher Marion Richardson (1892-1946). A dynamic and assertive woman, Richardson's ideas and practices changed British primary and secondary art teaching for many years. She often used "word pictures" (narrative descriptions of scenes or emotions) to inspire her students. (MJP)

  15. Management Practices of Soybean Producers in Marion County, Tennessee.

    ERIC Educational Resources Information Center

    Hall, William A.; And Others

    The purposes of the study were to: (1) determine some major characteristics of Marion County soybean producers and their farms; (2) more accurately determine which recommended production practices soybean producers were using in 1968 and 1969; (3) study the relation between use of recommended production practices and yield levels; and (4) identify…

  16. Marion Richardson: "Art and the Child," a Forgotten Classic

    ERIC Educational Resources Information Center

    Armstrong, Michael

    2015-01-01

    Marion Richardson was a revolutionary art teacher and schools inspector. First published in 1948, her book "Art and the Child" is one of the most remarkable educational documents of the period between the first and second world wars. This article reviews Richardson's philosophy and practice of art and suggests its continuing…

  17. San Andreas drilling sites selected

    NASA Astrophysics Data System (ADS)

    Ellsworth, Bill; Zoback, Mark

    A new initiative for drilling and coring directly into the San Andreas fault at depths up to 10 km is being proposed by an international team of scientists led by Mark Zoback, Stanford University; Steve Hickman and Bill Ellsworth, U.S. Geological Survey; and Lee Younker, Lawrence Livermore Laboratory. In addition to exhuming samples of fault rock and fluids from seismogenic depths, the hole will be used to make a wide range of geophysical measurements within the fault zone and to monitor the fault zone over time. Four areas along the San Andreas have been selected as candidates for deep drilling: the Mojave segment of the San Andreas between Leona Valley and Big Pine, the Carrizo Plain, the San Francisco Peninsula between Los Altos and Daly City, and the Northern Gabilan Range between the Cienga winery and Melendy Ranch. These sites were chosen from an initial list compiled at the International Fault Zone Drilling Workshop held in Asilomar, Calif., in December 1992 and at meetings held this winter and spring in Menlo Park, Calif.

  18. Subtropical Storm Andrea

    NASA Technical Reports Server (NTRS)

    2007-01-01

    The circling clouds of an intense low-pressure system sat off the southeast coast of the United States on May 8, 2007, when the Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA's Terra satellite captured this image. By the following morning, the storm developed enough to be classified as a subtropical storm, a storm that forms outside of the tropics, but has many of the characteristics--hurricane-force winds, driving rains, low pressure, and sometimes an eye--of a tropical storm. Although it arrived several weeks shy of the official start of the hurricane season (June 1), Subtropical Storm Andrea became the first named storm of the 2007 Atlantic hurricane season. The storm has the circular shape of a tropical cyclone in this image, but lacks the tight organization seen in more powerful storms. By May 9, the storm's winds reached 75 kilometers per hour (45 miles per hour), and the storm was not predicted to get any stronger, said the National Hurricane Center. Though Subtropical Storm Andrea was expected to remain offshore, its strong winds and high waves pummeled coastal states, prompting a tropical storm watch. The winds fueled wild fires (marked with red boxes) in Georgia and Florida. The wind-driven flames generated thick plumes of smoke that concentrated in a gray-brown mass over Tampa Bay, Florida. Unfortunately for Georgia and Florida, which are experiencing moderate to severe drought, Subtropical Storm Andrea was not predicted to bring significant rain to the region right away, according to reports on the Washington Post Website.

  19. NASA Sees Heavy Rainfall in Tropical Storm Andrea

    NASA Image and Video Library

    2013-06-06

    NASA’s Terra satellite passed over Tropical Storm Andrea on June 5 at 16:25 UTC (12:25 p.m. EDT) and the MODIS instrument captured this visible image of the storm. Andrea’s clouds had already extended over more than half of Florida. Credit: NASA Goddard MODIS Rapid Response Team --- NASA Sees Heavy Rainfall in Tropical Storm Andrea NASA’s TRMM satellite passed over Tropical Storm Andrea right after it was named, while NASA’s Terra satellite captured a visible image of the storm’s reach hours beforehand. TRMM measures rainfall from space and saw that rainfall rates in the southern part of the storm was falling at almost 5 inches per hour. NASA’s Terra satellite passed over Tropical Storm Andrea on June 5 at 16:25 UTC (12:25 p.m. EDT) and the Moderate Resolution Imaging Spectroradiometer or MODIS instrument, captured a visible image of the storm. At that time, Andrea’s clouds had already extended over more than half of Florida. At 8 p.m. EDT on Wednesday, June 5, System 91L became the first tropical storm of the Atlantic Ocean hurricane season. Tropical Storm Andrea was centered near 25.5 North and 86.5 West, about 300 miles (485 km) southwest of Tampa, Fla. At the time Andrea intensified into a tropical storm, its maximum sustained winds were near 40 mph (65 kph). Full updates can be found at NASA's Hurricane page: www.nasa.gov/hurricane Rob Gutro NASA’s Goddard Space Flight Center

  20. Some Reflections on Problem Posing: A Conversation with Marion Walter

    ERIC Educational Resources Information Center

    Baxter, Juliet A.

    2005-01-01

    Marion Walter, an internationally acclaimed mathematics educator discusses about problem posing, focusing on both the merits of problem posing and techniques to encourage problem posing. She believes that playful attitude toward problem variables is an essential part of an inquiring mind and the more opportunities that learners have, to change a…

  1. The Andrea Levialdi Fellowship

    NASA Astrophysics Data System (ADS)

    Fieschi, Roberto

    My first encounter with Cuba dates back to winter 1967-1968 at the Cultural Congress of La Havana, a very large international event to promote greater understanding of the reality of the Cuban Revolution. In fact the person invited was my friend and colleague Andrea Levialdi (Andrea already knew Cuba and loved it) who, unable to participate, allowed me to go in her place. So I landed at the airport of the "first free country in Latin America" with the delegation of the Italian Communist Party. In Havana I met other Italian physicists whom I already knew, among them Bruno Vitale and Daniele Amati. They, like me, were embarrassed by the generous hospitality of `Havana Libre,' especially in a country which was going through such difficulties. Despite our best efforts we did not succeed in receiving a more modest welcome.

  2. 77 FR 47361 - Francis Marion Sumter National Forests Resource Advisory Committee

    Federal Register 2010, 2011, 2012, 2013, 2014

    2012-08-08

    ... office, Large Conference Room, 4931 Broad River Road, Columbia, SC 29212. Written comments may be... inspect comments received at Francis Marion Sumter National Forests, 4931 Broad River Road, Columbia, SC..., 4931 Broad River Road, Columbia, SC 29212; (803) 561-4058; Email [email protected] . Individuals who...

  3. San Andreas tremor cascades define deep fault zone complexity

    USGS Publications Warehouse

    Shelly, David R.

    2015-01-01

    Weak seismic vibrations - tectonic tremor - can be used to delineate some plate boundary faults. Tremor on the deep San Andreas Fault, located at the boundary between the Pacific and North American plates, is thought to be a passive indicator of slow fault slip. San Andreas Fault tremor migrates at up to 30 m s-1, but the processes regulating tremor migration are unclear. Here I use a 12-year catalogue of more than 850,000 low-frequency earthquakes to systematically analyse the high-speed migration of tremor along the San Andreas Fault. I find that tremor migrates most effectively through regions of greatest tremor production and does not propagate through regions with gaps in tremor production. I interpret the rapid tremor migration as a self-regulating cascade of seismic ruptures along the fault, which implies that tremor may be an active, rather than passive participant in the slip propagation. I also identify an isolated group of tremor sources that are offset eastwards beneath the San Andreas Fault, possibly indicative of the interface between the Monterey Microplate, a hypothesized remnant of the subducted Farallon Plate, and the North American Plate. These observations illustrate a possible link between the central San Andreas Fault and tremor-producing subduction zones.

  4. Mary Ritter Beard and Marion Thompson Wright: Shaping Inclusive Social Education.

    ERIC Educational Resources Information Center

    Crocco, Margaret Smith

    1997-01-01

    Examines contributions by Mary Ritter Beard and Marion Thompson Wright to inclusive social education curricula. Beard established the field of women's history; Wright promoted the application of black history. Both saw social betterment as the goal of knowledge and sought inclusive understanding of the nature of legitimate knowledge in schools.…

  5. Flavour Profiling of 'Marion' and Thornless Blackberries by Instrumental and Sensory Analysis

    USDA-ARS?s Scientific Manuscript database

    The flavour of thornless blackberries grown in Pacific Northwest including 'Thornless Evergreen', 'Black Diamond', 'Black Pearl', 'Nightfall', ORUS 1843-3, 'Waldo', NZ 9351-4, and 'Chester Thornless' as well as 'Marion' was profiled by sensory evaluation and instrumental analysis. Sensory results sh...

  6. Deus Ex Machina: Tradition, Technology, and the Chicanafuturist Art of Marion C. Martinez

    ERIC Educational Resources Information Center

    Ramirez, Catherine S.

    2004-01-01

    The visual art of Marion C. Martinez is examined. Through technology, Martinez reproduces and transforms traditional Indo-Hispanic art forms, at the same time, underscores New Mexico's history as a dumping ground for technological waste.

  7. CloudSat Profiles Tropical Storm Andrea

    NASA Image and Video Library

    2007-05-10

    CloudSat's Cloud Profiling Radar captured a profile across Tropical Storm Andrea on Wednesday, May 9, 2007, near the South Carolina/Georgia/Florida Atlantic coast. The upper image shows an infrared view of Tropical Storm Andrea from the Moderate Resolution Imaging Spectroradiometer instrument on NASA's Aqua satellite, with CloudSat's ground track shown as a red line. The lower image is the vertical cross section of radar reflectivity along this path, where the colors indicate the intensity of the reflected radar energy. CloudSat orbits approximately one minute behind Aqua in a satellite formation known as the A-Train. http://photojournal.jpl.nasa.gov/catalog/PIA09379

  8. A Return to Love in William James and Jean-Luc Marion

    ERIC Educational Resources Information Center

    Rocha, Samuel

    2009-01-01

    In this essay Samuel Rocha primarily addresses, and challenges, the modern conception of reason and the lowly place of intuition, feeling, and love in what has become traditional philosophy and education. Drawing upon the rich thought of William James and Jean-Luc Marion, Rocha introduces the reader to a certain harmony between their ideas, most…

  9. Bound Volatile Precursors in Genotypes in the Pedigree of 'Marion' Blackberry (Rubus Sp.)

    USDA-ARS?s Scientific Manuscript database

    Glycosidically bound volatiles and precursors in genotypes representing the pedigree for 'Marion' blackberry were investigated over two growing seasons. The volatile precursors were isolated using a C18 solid-phase extraction column. After enzymatic hydrolysis, the released volatiles were analyzed u...

  10. Low strength of deep San Andreas fault gouge from SAFOD core

    USGS Publications Warehouse

    Lockner, David A.; Morrow, Carolyn A.; Moore, Diane E.; Hickman, Stephen H.

    2011-01-01

    The San Andreas fault accommodates 28–34 mm yr−1 of right lateral motion of the Pacific crustal plate northwestward past the North American plate. In California, the fault is composed of two distinct locked segments that have produced great earthquakes in historical times, separated by a 150-km-long creeping zone. The San Andreas Fault Observatory at Depth (SAFOD) is a scientific borehole located northwest of Parkfield, California, near the southern end of the creeping zone. Core was recovered from across the actively deforming San Andreas fault at a vertical depth of 2.7 km (ref. 1). Here we report laboratory strength measurements of these fault core materials at in situ conditions, demonstrating that at this locality and this depth the San Andreas fault is profoundly weak (coefficient of friction, 0.15) owing to the presence of the smectite clay mineral saponite, which is one of the weakest phyllosilicates known. This Mg-rich clay is the low-temperature product of metasomatic reactions between the quartzofeldspathic wall rocks and serpentinite blocks in the fault2, 3. These findings provide strong evidence that deformation of the mechanically unusual creeping portions of the San Andreas fault system is controlled by the presence of weak minerals rather than by high fluid pressure or other proposed mechanisms1. The combination of these measurements of fault core strength with borehole observations1, 4, 5 yields a self-consistent picture of the stress state of the San Andreas fault at the SAFOD site, in which the fault is intrinsically weak in an otherwise strong crust.

  11. Low strength of deep San Andreas fault gouge from SAFOD core

    USGS Publications Warehouse

    Lockner, D.A.; Morrow, C.; Moore, D.; Hickman, S.

    2011-01-01

    The San Andreas fault accommodates 28-"34-???mm-???yr ????'1 of right lateral motion of the Pacific crustal plate northwestward past the North American plate. In California, the fault is composed of two distinct locked segments that have produced great earthquakes in historical times, separated by a 150-km-long creeping zone. The San Andreas Fault Observatory at Depth (SAFOD) is a scientific borehole located northwest of Parkfield, California, near the southern end of the creeping zone. Core was recovered from across the actively deforming San Andreas fault at a vertical depth of 2.7-???km (ref. 1). Here we report laboratory strength measurements of these fault core materials at in situ conditions, demonstrating that at this locality and this depth the San Andreas fault is profoundly weak (coefficient of friction, 0.15) owing to the presence of the smectite clay mineral saponite, which is one of the weakest phyllosilicates known. This Mg-rich clay is the low-temperature product of metasomatic reactions between the quartzofeldspathic wall rocks and serpentinite blocks in the fault. These findings provide strong evidence that deformation of the mechanically unusual creeping portions of the San Andreas fault system is controlled by the presence of weak minerals rather than by high fluid pressure or other proposed mechanisms. The combination of these measurements of fault core strength with borehole observations yields a self-consistent picture of the stress state of the San Andreas fault at the SAFOD site, in which the fault is intrinsically weak in an otherwise strong crust. ?? 2011 Macmillan Publishers Limited. All rights reserved.

  12. The portrayal of J. Marion Sims' controversial surgical legacy.

    PubMed

    Spettel, Sara; White, Mark Donald

    2011-06-01

    In the mid 1800s Dr. J. Marion Sims reported the successful repair of vesicovaginal fistulas with a technique he developed by performing multiple operations on female slaves. A venerated physician in his time, the legacy of Dr. Sims is controversial and represents a significant chapter in the mistreatment of African-Americans by the medical establishment. This review compares the modern debate surrounding his legacy with the presentation of his operation in widely consulted urological texts and journals. A literature review was performed of medical, sociological and periodical sources (1851 to the present) regarding J. Marion Sims and vesicovaginal fistula repair. During the last several decades, while the controversy around Dr. Sims' surgical development has produced a steady stream of articles in the historical and popular literature, relatively little mention is found in standard urology textbooks or journals. With increased public attention, some have debated the removal or modification of public tributes to Dr. Sims. This move has been countered by arguments against the validity of judging a 19th century physician by modern standards. While historians, ethicists and the popular press have debated Dr. Sims' legacy, medical sources have continued to portray him unquestionably as a great figure in medical history. This division keeps the medical profession uninformed and detached from the public debate on his legacy and, thus, the larger issues of ethical treatment of surgical patients. Copyright © 2011 American Urological Association Education and Research, Inc. Published by Elsevier Inc. All rights reserved.

  13. 75 FR 27383 - CSX Transportation, Inc.-Abandonment Exemption-in Marion County, IN.

    Federal Register 2010, 2011, 2012, 2013, 2014

    2010-05-14

    ... Running Track, in Indianapolis, Marion County, Ind. The line traverses United States Postal Service Zip... interim trail use. Any request for a public use condition under 49 CFR 1152.28 or for trail use/rail banking under 49 CFR 1152.29 will be due no later than June 3, 2010. Each trail use request must be...

  14. Volatile Composition and Odour-Activity Value of Thornless 'Black Diamond' and 'Marion' Blackberries

    USDA-ARS?s Scientific Manuscript database

    'Black Diamond' is a recently developed thornless blackberry cultivar with large fruit size, high yield, and good processed fruit quality that has rapidly become an industry standard. The flavour of 'Black Diamond' fruit is not the same as 'Marion', which is regarded by the industry as having the id...

  15. 78 FR 11733 - Georgia Southwestern Railroad, Inc.-Discontinuance of Service Exemption-in Chattahoochee, Marion...

    Federal Register 2010, 2011, 2012, 2013, 2014

    2013-02-19

    ... discontinue service, not to abandon the line, trail use/rail banking and public use conditions are not... No. AB 290 (Sub-No. 344X)] Georgia Southwestern Railroad, Inc.--Discontinuance of Service Exemption...--Discontinuance of Service Exemption--in Chattahoochee, Marion, and Schley Counties, GA Central of Georgia...

  16. Hydrologic and water quality monitoring on Turkey Creek watershed, Francis Marion National Forest, SC

    Treesearch

    D.M. Amatya; T.J. Callahan; A. Radecki-Pawlik; P. Drewes; C. Trettin; W.F. Hansen

    2008-01-01

    The re-initiation of a 7,260 ha forested watershed study on Turkey Creek, a 3rd order stream, within the Francis Marion National forest in South Carolina, completes the development of a multi-scale hydrology and ecosystem monitoring framework in the Atlantic Coastal Plain. Hydrology and water quality monitoring began on the Santee Experimental...

  17. Earthquake Swarm Along the San Andreas Fault near Palmdale, Southern California, 1976 to 1977.

    PubMed

    McNally, K C; Kanamori, H; Pechmann, J C; Fuis, G

    1978-09-01

    Between November 1976 and November 1977 a swarm of small earthquakes (local magnitude Andreas fault near Palmdale, California. This swarm was the first observed along this section of the San Andreas since cataloging of instrumental data began in 1932. The activity followed partial subsidence of the 35-centimeter vertical crustal uplift known as the Palmdale bulge along this "locked" section of the San Andreas, which last broke in the great (surface-wave magnitude = 8(1/4)+) 1857 Fort Tejon earthquake. The swarm events exhibit characteristics previously observed for some foreshock sequences, such as tight clustering of hypocenters and time-dependent rotations of stress axes inferred from focal mechanisms. However, because of our present lack of understanding of the processes that precede earthquake faulting, the implications of the swarm for future large earthquakes on the San Andreas fault are unknown.

  18. Earthquake swarm along the San Andreas fault near Palmdale, Southern California, 1976 to 1977

    USGS Publications Warehouse

    Mcnally, K.C.; Kanamori, H.; Pechmann, J.C.; Fuis, G.

    1978-01-01

    Between November 1976 and November 1977 a swarm of small earthquakes (local magnitude ??? 3) occurred on or near the San Andreas fault near Palmdale, California. This swarm was the first observed along this section of the San Andreas since cataloging of instrumental data began in 1932. The activity followed partial subsidence of the 35-centimeter vertical crustal uplift known as the Palmdale bulge along this "locked" section of the San Andreas, which last broke in the great (surface-wave magnitude = 81/4+) 1857 Fort Tejon earthquake. The swarm events exhibit characteristics previously observed for some foreshock sequences, such as tight clustering of hypocenters and time-dependent rotations of stress axes inferred from focal mechanisms. However, because of our present lack of understanding of the processes that precede earthquake faulting, the implications of the swarm for future large earthquakes on the San Andreas fault are unknown. Copyright ?? 1978 AAAS.

  19. Vertical tectonic deformation associated with the San Andreas fault zone offshore of San Francisco, California

    USGS Publications Warehouse

    Ryan, H.F.; Parsons, T.; Sliter, R.W.

    2008-01-01

    A new fault map of the shelf offshore of San Francisco, California shows that faulting occurs as a distributed shear zone that involves many fault strands with the principal displacement taken up by the San Andreas fault and the eastern strand of the San Gregorio fault zone. Structures associated with the offshore faulting show compressive deformation near where the San Andreas fault goes offshore, but deformation becomes extensional several km to the north off of the Golden Gate. Our new fault map serves as the basis for a 3-D finite element model that shows that the block between the San Andreas and San Gregorio fault zone is subsiding at a long-term rate of about 0.2-0.3??mm/yr, with the maximum subsidence occurring northwest of the Golden Gate in the area of a mapped transtensional basin. Although the long-term rates of vertical displacement primarily show subsidence, the model of coseismic deformation associated with the 1906 San Francisco earthquake indicates that uplift on the order of 10-15??cm occurred in the block northeast of the San Andreas fault. Since 1906, 5-6??cm of regional subsidence has occurred in that block. One implication of our model is that the transfer of slip from the San Andreas fault to a fault 5??km to the east, the Golden Gate fault, is not required for the area offshore of San Francisco to be in extension. This has implications for both the deposition of thick Pliocene-Pleistocene sediments (the Merced Formation) observed east of the San Andreas fault, and the age of the Peninsula segment of the San Andreas fault.

  20. 75 FR 39460 - Prevailing Rate Systems; Redefinition of the Chicago, IL; Fort Wayne-Marion, IN; Indianapolis, IN...

    Federal Register 2010, 2011, 2012, 2013, 2014

    2010-07-09

    .... Joseph Steuben Wabash White Whitley Ohio: Allen Defiance Fulton Henry Mercer Paulding Putnam Van Wert Williams INDIANAPOLIS Survey Area Indiana: Boone Hamilton Hancock Hendricks Johnson Marion Morgan Shelby... Vermillion Vigo Warren * * * * * OHIO * * * * * Cleveland Survey Area Ohio: Cuyahoga Geauga Lake Medina Area...

  1. Deep permeability of the San Andreas Fault from San Andreas Fault Observatory at Depth (SAFOD) core samples

    USGS Publications Warehouse

    Morrow, Carolyn A.; Lockner, David A.; Moore, Diane E.; Hickman, Stephen H.

    2014-01-01

    The San Andreas Fault Observatory at Depth (SAFOD) scientific borehole near Parkfield, California crosses two actively creeping shear zones at a depth of 2.7 km. Core samples retrieved from these active strands consist of a foliated, Mg-clay-rich gouge containing porphyroclasts of serpentinite and sedimentary rock. The adjacent damage zone and country rocks are comprised of variably deformed, fine-grained sandstones, siltstones, and mudstones. We conducted laboratory tests to measure the permeability of representative samples from each structural unit at effective confining pressures, Pe up to the maximum estimated in situ Pe of 120 MPa. Permeability values of intact samples adjacent to the creeping strands ranged from 10−18 to 10−21 m2 at Pe = 10 MPa and decreased with applied confining pressure to 10−20–10−22 m2 at 120 MPa. Values for intact foliated gouge samples (10−21–6 × 10−23 m2 over the same pressure range) were distinctly lower than those for the surrounding rocks due to their fine-grained, clay-rich character. Permeability of both intact and crushed-and-sieved foliated gouge measured during shearing at Pe ≥ 70 MPa ranged from 2 to 4 × 10−22 m2 in the direction perpendicular to shearing and was largely insensitive to shear displacement out to a maximum displacement of 10 mm. The weak, actively-deforming foliated gouge zones have ultra-low permeability, making the active strands of the San Andreas Fault effective barriers to cross-fault fluid flow. The low matrix permeability of the San Andreas Fault creeping zones and adjacent rock combined with observations of abundant fractures in the core over a range of scales suggests that fluid flow outside of the actively-deforming gouge zones is probably fracture dominated.

  2. Overview of the Southern San Andreas Fault Model

    USGS Publications Warehouse

    Weldon, Ray J.; Biasi, Glenn P.; Wills, Chris J.; Dawson, Timothy E.

    2008-01-01

    This appendix summarizes the data and methodology used to generate the source model for the southern San Andreas fault. It is organized into three sections, 1) a section by section review of the geological data in the format of past Working Groups, 2) an overview of the rupture model, and 3) a manuscript by Biasi and Weldon (in review Bulletin of the Seismological Society of America) that describes the correlation methodology that was used to help develop the ?geologic insight? model. The goal of the Biasi and Weldon methodology is to quantify the insight that went into developing all A faults; as such it is in concept consistent with all other A faults but applied in a more quantitative way. The most rapidly slipping fault and the only known source of M~8 earthquakes in southern California is the San Andreas fault. As such it plays a special role in the seismic hazard of California, and has received special attention in the current Working Group. The underlying philosophy of the current Working Group is to model the recurrence behavior of large, rapidly slipping faults like the San Andreas from observed data on the size, distribution and timing of past earthquakes with as few assumptions about underlying recurrence behavior as possible. In addition, we wish to carry the uncertainties in the data and the range of reasonable extrapolations from the data to the final model. To accomplish this for the Southern San Andreas fault we have developed an objective method to combine all of the observations of size, timing, and distribution of past earthquakes into a comprehensive set of earthquake scenarios that each represent a possible history of earthquakes for the past ~1400 years. The scenarios are then ranked according to their overall consistency with the data and then the frequencies of all of the ruptures permitted by the current Working Group?s segmentation model are calculated. We also present 30-yr conditional probabilities by segment and compare to previous

  3. Abrupt along-strike change in tectonic style: San Andreas fault zone, San Francisco Peninsula

    USGS Publications Warehouse

    Zoback, M.L.; Jachens, R.C.; Olson, J.A.

    1999-01-01

    Seismicity and high-resolution aeromagnetic data are used to define an abrupt change from compressional to extensional tectonism within a 10- to 15-km-wide zone along the San Andreas fault on the San Francisco Peninsula and offshore from the Golden Gate. This 100-km-long section of the San Andreas fault includes the hypocenter of the Mw = 7.8 1906 San Francisco earthquake as well as the highest level of persistent microseismicity along that ???470-km-long rupture. We define two distinct zones of deformation along this stretch of the fault using well-constrained relocations of all post-1969 earthquakes based a joint one-dimensional velocity/hypocenter inversion and a redetermination of focal mechanisms. The southern zone is characterized by thrust- and reverse-faulting focal mechanisms with NE trending P axes that indicate "fault-normal" compression in 7- to 10-km-wide zones of deformation on both sides of the San Andreas fault. A 1- to 2-km-wide vertical zone beneath the surface trace of the San Andreas is characterized by its almost complete lack of seismicity. The compressional deformation is consistent with the young, high topography of the Santa Cruz Mountains/Coast Ranges as the San Andreas fault makes a broad restraining left bend (???10??) through the southernmost peninsula. A zone of seismic quiescence ???15 km long separates this compressional zone to the south from a zone of combined normal-faulting and strike-slip-faulting focal mechanisms (including a ML = 5.3 earthquake in 1957) on the northernmost peninsula and offshore on the Golden Gate platform. Both linear pseudo-gravity gradients, calculated from the aeromagnetic data, and seismic reflection data indicate that the San Andreas fault makes an abrupt ???3-km right step less than 5 km offshore in this northern zone. A similar right-stepping (dilatational) geometry is also observed for the subparallel San Gregorio fault offshore. Persistent seismicity and extensional tectonism occur within the San

  4. 78 FR 59913 - Revision of the Land Management Plan for the Francis Marion National Forest

    Federal Register 2010, 2011, 2012, 2013, 2014

    2013-09-30

    ... our Web site. We are inviting the public to help us develop a preliminary ``need for change'' and a... revision of the Francis Marion NF land management plan will be posted on the following Web site at www.fs... preliminary ``need for change'' and a proposed action will be announced on the Web site cited above. It is...

  5. Miocene History of the East Antarctic Ice-sheet Inferred from the Eustatic and Paleoceanographic Record of The Marion Plateau, Northeastern Australia (ODP Leg 194)

    NASA Astrophysics Data System (ADS)

    John, C. M.; Browning, E.; Lowery, C.; Leckie, R. M.; Karner, G. D.; Schouten, S.

    2012-12-01

    The East Antarctic Ice Sheet (EAIS) had a major influence on Cenozoic eustasy and paleoceanography. Reconstructing changes in ice volume and climate associated with the EAIS is thus critical to a better understanding of climate dynamics, but this has proven difficult to do using only sedimentary records from Antarctica because ice sheets tend to erase evidences of their own history. It is possible, however, to gain information about the dynamics of the EAIS by reconstructing glacio-eustasy and regional paleoceanographic changes away from Antarctica. We present a record of carbonate deposition spanning the uppermost Oligocene to upper Miocene on the Marion Plateau of Northeastern Australia (ODP Leg 194). The Marion Plateau is ideally located for our study on a tectonically quiescent margin, and it is a sensitive recorder of Antarctic paleoenvironmental changes due to its geographical position in the southern hemisphere. We conducted a multi-disciplinary study involving sedimentology, sequence stratigraphy, calcareous nannofossil and foraminifer micropaleontology, and organic geochemistry. Eleven lower and middle Miocene sequence boundaries can be recognized on the Marion Plateau based on facies and seismic reflection lines, and each sequence boundary (sea-level fall) is associated with a phase of growth of the EAIS (as evidenced by positive shifts in δ18O, termed "Mi Events"). The oldest boundary recovered is 23.16 Ma old. By combining backstripping and δ18O estimates, the amplitude of four of the Miocene eustatic falls are constrained to 27±1 m at 16.5 Ma (Mi2), 27±1 m at 15.6 Ma (Mi2a), 33±3 m at 14.8 Ma (Mi3a), and 59±6 m at 13.6 Ma (Mi3). The amplitude of the eustatic drop associated with event Mi3 suggests a major growth phase of the EAIS in the middle Miocene, which is supported by other deep-sea records and several recent studies from continental Antarctica. Furthermore, nannofossil assemblages show four main clusters corresponding to changes in

  6. Loading of the San Andreas fault by flood-induced rupture of faults beneath the Salton Sea

    USGS Publications Warehouse

    Brothers, Daniel; Kilb, Debi; Luttrell, Karen; Driscoll, Neal W.; Kent, Graham

    2011-01-01

    The southern San Andreas fault has not experienced a large earthquake for approximately 300 years, yet the previous five earthquakes occurred at ~180-year intervals. Large strike-slip faults are often segmented by lateral stepover zones. Movement on smaller faults within a stepover zone could perturb the main fault segments and potentially trigger a large earthquake. The southern San Andreas fault terminates in an extensional stepover zone beneath the Salton Sea—a lake that has experienced periodic flooding and desiccation since the late Holocene. Here we reconstruct the magnitude and timing of fault activity beneath the Salton Sea over several earthquake cycles. We observe coincident timing between flooding events, stepover fault displacement and ruptures on the San Andreas fault. Using Coulomb stress models, we show that the combined effect of lake loading, stepover fault movement and increased pore pressure could increase stress on the southern San Andreas fault to levels sufficient to induce failure. We conclude that rupture of the stepover faults, caused by periodic flooding of the palaeo-Salton Sea and by tectonic forcing, had the potential to trigger earthquake rupture on the southern San Andreas fault. Extensional stepover zones are highly susceptible to rapid stress loading and thus the Salton Sea may be a nucleation point for large ruptures on the southern San Andreas fault.

  7. The medical ethics of the 'father of gynaecology', Dr J Marion Sims.

    PubMed Central

    Ojanuga, D

    1993-01-01

    Vesico-vaginal fistula (VVF) was a common ailment among American women in the 19th century. Prior to that time, no successful surgery had been developed for the cure of this condition until Dr J Marion Sims perfected a successful surgical technique in 1849. Dr Sims used female slaves as research subjects over a four-year period of experimentation (1845-1849). This paper discusses the controversy surrounding his use of powerless women and whether his actions were acceptable during that historical period. PMID:8459435

  8. Fall food habits of wood ducks from Lake Marion, South Carolina

    USGS Publications Warehouse

    McGilvrey, F.B.

    1966-01-01

    A total of 108 stomachs of wood ducks (Aix sponsa) collected from hunters on the upper end of Lake Marion, South Carolina, between November 29 and December 6, 1961, were examined for information on food habits. Six plants made up over 98 percent of the total volume. Five were tree fruits: water and pin oak (Quercus nigra and Q. palustris), baldcypress (Taxodium distichum), sweetgum (Liquidambar styraciflua), and water hickory (Carya aquatica). The sixth important food was corn (Zea mays). In areas being managed for wood ducks and timber, therefore, these tree species should not be removed.

  9. Seguenziidae (Gastropoda: Vetigastropoda) from SE Brazil collected by the Marion Dufresne (MD55) expedition .

    PubMed

    Salvador, Rodrigo B; Cavallari, Daniel C; Simone, Luiz R L

    2014-10-30

    The present work deals with the vetigastropods of the family Seguenziidae collected by the Marion Dufresne (MD55) expedition in SE Brazil, reporting the occurrence of eight species. The following species have their geographical range extended: Ancistrobasis costulata, Carenzia carinata, Carenzia trispinosa, Hadroconus altus, Seguenzia elegans and Seguenzia formosa. Two new species, Halystina umberlee sp. nov. and Seguenzia triteia sp. nov., are described.

  10. Coulomb Stress Accumulation along the San Andreas Fault System

    NASA Technical Reports Server (NTRS)

    Smith, Bridget; Sandwell, David

    2003-01-01

    Stress accumulation rates along the primary segments of the San Andreas Fault system are computed using a three-dimensional (3-D) elastic half-space model with realistic fault geometry. The model is developed in the Fourier domain by solving for the response of an elastic half-space due to a point vector body force and analytically integrating the force from a locking depth to infinite depth. This approach is then applied to the San Andreas Fault system using published slip rates along 18 major fault strands of the fault zone. GPS-derived horizontal velocity measurements spanning the entire 1700 x 200 km region are then used to solve for apparent locking depth along each primary fault segment. This simple model fits remarkably well (2.43 mm/yr RMS misfit), although some discrepancies occur in the Eastern California Shear Zone. The model also predicts vertical uplift and subsidence rates that are in agreement with independent geologic and geodetic estimates. In addition, shear and normal stresses along the major fault strands are used to compute Coulomb stress accumulation rate. As a result, we find earthquake recurrence intervals along the San Andreas Fault system to be inversely proportional to Coulomb stress accumulation rate, in agreement with typical coseismic stress drops of 1 - 10 MPa. This 3-D deformation model can ultimately be extended to include both time-dependent forcing and viscoelastic response.

  11. Viscoelastic coupling model of the San Andreas fault along the big bend, southern California

    USGS Publications Warehouse

    Savage, J.C.; Lisowski, M.

    1997-01-01

    The big bend segment of the San Andreas fault is the 300-km-long segment in southern California that strikes about N65??W, roughly 25?? counterclockwise from the local tangent to the small circle about the Pacific-North America pole of rotation. The broad distribution of deformation of trilateration networks along this segment implies a locking depth of at least 25 km as interpreted by the conventional model of strain accumulation (continuous slip on the fault below the locking depth at the rate of relative plate motion), whereas the observed seismicity and laboratory data on fault strength suggest that the locking depth should be no greater than 10 to 15 km. The discrepancy is explained by the viscoelastic coupling model which accounts for the viscoelastic response of the lower crust. Thus the broad distribution of deformation observed across the big bend segment can be largely associated with the San Andreas fault itself, not subsidiary faults distributed throughout the region. The Working Group on California Earthquake Probabilities [1995] in using geodetic data to estimate the seismic risk in southern California has assumed that strain accumulated off the San Andreas fault is released by earthquakes located off the San Andreas fault. Thus they count the San Andreas contribution to total seismic moment accumulation more than once, leading to an overestimate of the seismicity for magnitude 6 and greater earthquakes in their Type C zones.

  12. Correlation of clayey gouge in a surface exposure of serpentinite in the San Andreas Fault with gouge from the San Andreas Fault Observatory at Depth (SAFOD)

    NASA Astrophysics Data System (ADS)

    Moore, Diane E.; Rymer, Michael J.

    2012-05-01

    Magnesium-rich clayey gouge similar to that comprising the two actively creeping strands of the San Andreas Fault in drill core from the San Andreas Fault Observatory at Depth (SAFOD) has been identified in a nearby outcrop of serpentinite within the fault zone at Nelson Creek. Each occurrence of the gouge consists of porphyroclasts of serpentinite and sedimentary rocks dispersed in a fine-grained, foliated matrix of Mg-rich smectitic clays. The clay minerals in all three gouges are interpreted to be the product of fluid-assisted, shear-enhanced reactions between quartzofeldspathic wall rocks and serpentinite that was tectonically entrained in the fault from a source in the Coast Range Ophiolite. We infer that the gouge at Nelson Creek connects to one or both of the gouge zones in the SAFOD core, and that similar gouge may occur at depths in between. The special significance of the outcrop is that it preserves the early stages of mineral reactions that are greatly advanced at depth, and it confirms the involvement of serpentinite and the Mg-rich phyllosilicate minerals that replace it in promoting creep along the central San Andreas Fault.

  13. Fine-scale delineation of the location of and relative ground shaking within the San Andreas Fault zone at San Andreas Lake, San Mateo County, California

    USGS Publications Warehouse

    Catchings, R.D.; Rymer, M.J.; Goldman, M.R.; Prentice, C.S.; Sickler, R.R.

    2013-01-01

    The San Francisco Public Utilities Commission is seismically retrofitting the water delivery system at San Andreas Lake, San Mateo County, California, where the reservoir intake system crosses the San Andreas Fault (SAF). The near-surface fault location and geometry are important considerations in the retrofit effort. Because the SAF trends through highly distorted Franciscan mélange and beneath much of the reservoir, the exact trace of the 1906 surface rupture is difficult to determine from surface mapping at San Andreas Lake. Based on surface mapping, it also is unclear if there are additional fault splays that extend northeast or southwest of the main surface rupture. To better understand the fault structure at San Andreas Lake, the U.S. Geological Survey acquired a series of seismic imaging profiles across the SAF at San Andreas Lake in 2008, 2009, and 2011, when the lake level was near historical lows and the surface traces of the SAF were exposed for the first time in decades. We used multiple seismic methods to locate the main 1906 rupture zone and fault splays within about 100 meters northeast of the main rupture zone. Our seismic observations are internally consistent, and our seismic indicators of faulting generally correlate with fault locations inferred from surface mapping. We also tested the accuracy of our seismic methods by comparing our seismically located faults with surface ruptures mapped by Schussler (1906) immediately after the April 18, 1906 San Francisco earthquake of approximate magnitude 7.9; our seismically determined fault locations were highly accurate. Near the reservoir intake facility at San Andreas Lake, our seismic data indicate the main 1906 surface rupture zone consists of at least three near-surface fault traces. Movement on multiple fault traces can have appreciable engineering significance because, unlike movement on a single strike-slip fault trace, differential movement on multiple fault traces may exert compressive and

  14. Migrating tremors illuminate complex deformation beneath the seismogenic San Andreas fault

    USGS Publications Warehouse

    Shelly, David R.

    2010-01-01

    The San Andreas fault is one of the most extensively studied faults in the world, yet its physical character and deformation mode beneath the relatively shallow earthquake-generating portion remain largely unconstrained. Tectonic ‘non-volcanic’ tremor, a recently discovered seismic signal probably generated by shear slip on the deep extension of some major faults, can provide new insight into the deep fate of such faults, including that of the San Andreas fault near Parkfield, California. Here I examine continuous seismic data from mid-2001 to 2008, identifying tremor and decomposing the signal into different families of activity based on the shape and timing of the waveforms at multiple stations. This approach allows differentiation between activities from nearby patches of the deep fault and begins to unveil rich and complex patterns of tremor occurrence. I find that tremor exhibits nearly continuous migration, with the most extensive episodes propagating more than 20 kilometres along fault strike at rates of 15–80 kilometres per hour. This suggests that the San Andreas fault remains a localized through-going structure, at least to the base of the crust, in this area. Tremor rates and recurrence behaviour changed markedly in the wake of the 2004 magnitude-6.0 Parkfield earthquake, but these changes were far from uniform within the tremor zone, probably reflecting heterogeneous fault properties and static and dynamic stresses decaying away from the rupture. The systematic recurrence of tremor demonstrated here suggests the potential to monitor detailed time-varying deformation on this portion of the deep San Andreas fault, deformation which unsteadily loads the shallower zone that last ruptured in the 1857 magnitude-7.9 Fort Tejon earthquake.

  15. δ18O and Marion Plateau backstripping: Combining two approaches to constrain late middle Miocene eustatic amplitude

    NASA Astrophysics Data System (ADS)

    John, Cédric M.; Karner, Garry D.; Mutti, Maria

    2004-09-01

    δ18Obenthic values from Leg 194 Ocean Drilling Program Sites 1192 and 1195 (drilled on the Marion Plateau) were combined with deep-sea values to reconstruct the magnitude range of the late middle Miocene sea-level fall (13.6 11.4 Ma). In parallel, an estimate for the late middle Miocene sea-level fall was calculated from the stratigraphic relationship identified during Leg 194 and the structural relief of carbonate platforms that form the Marion Plateau. Corrections for thermal subsidence induced by Late Cretaceous rifting, flexural sediment loading, and sediment compaction were taken into account. The response of the lithosphere to sediment loading was considered for a range of effective elastic thicknesses (10 < Te < 40 km). By overlapping the sea-level range of both the deep-sea isotopes and the results from the backstripping analysis, we demonstrate that the amplitude of the late middle Miocene sea-level fall was 45 68 m (56.5 ± 11.5 m). Including an estimate for sea-level variation using the δ18Obenthic results from the subtropical Marion Plateau, the range of sea-level fall is tightly constrained between 45 and 55 m (50.0 ± 5.0 m). This result is the first precise quantitative estimate for the amplitude of the late middle Miocene eustatic fall that sidesteps the errors inherent in using benthic foraminifera assemblages to predict paleo water depth. The estimate also includes an error analysis for the flexural response of the lithosphere to both water and sediment loads. Our result implies that the extent of ice buildup in the Miocene was larger than previously estimated, and conversely that the amount of cooling associated with this event was less important.

  16. Gender and Hyper-Linear History in the Representation of the Female Australian Primary School Teacher in "Marion" (ABCTV, 1974)

    ERIC Educational Resources Information Center

    May, Josephine

    2018-01-01

    Building on the author's previous work on Australian national cinema and schooling, this article explores the representation of the female primary school teacher in the television mini-series entitled "Marion" (Australian Broadcasting Commission, 1974). Using narrative analysis, it argues that this representation is disruptive of…

  17. Sediment inflow, outflow and deposition for Lakes Marion and Moultrie, South Carolina, October 1983-March 1985

    USGS Publications Warehouse

    Cooney, T.W.

    1988-01-01

    In 1941 a Coastal Plain reach of the Santee River was impounded to form Lake Marion and diverted into a diked-off part of the Cooper River basin to form Lake Moultrie. Rates of sediment inflow and outflow of the lakes were determined by the U.S. Geological Survey for the periods July 1966 - June 1968 and October 1983 - March 1985. Total sediment discharge was estimated for two inflow stations and continuous streamflow monitors and automatic suspended-sediment samplers were used for computation of suspended-sediment discharge. Bedload discharge was computed by the modified Einstein procedure. Suspended-sediment discharge was monitored at three outflow stations, with the suspended-sediment concentration measured on a weekly basis. During the 1983-1985 study, mean annual suspended-sediment inflow to Lakes Marion and Moultrie was estimated to be 722,000 tons, and the outflow was estimated at 175,000 tons, for a trap efficiency of 76% and a deposition rate of about 547,000 tons/year. This is about 33% less than the deposition rate determined during the 1966-68 study. The deposition rate for suspended and bedload sediment during the 1983 - 1985 study was about 650,000 tons/year. (USGS)

  18. Change in failure stress on the southern San Andreas fault system caused by the 1992 magnitude = 7.4 Landers earthquake

    USGS Publications Warehouse

    Stein, R.S.; King, G.C.P.; Lin, J.

    1992-01-01

    The 28 June Landers earthquake brought the San Andreas fault significantly closer to failure near San Bernardino, a site that has not sustained a large shock since 1812. Stress also increased on the San Jacinto fault near San Bernardino and on the San Andreas fault southeast of Palm Springs. Unless creep or moderate earthquakes relieve these stress changes, the next great earthquake on the southern San Andreas fault is likely to be advanced by one to two decades. In contrast, stress on the San Andreas north of Los Angeles dropped, potentially delaying the next great earthquake there by 2 to 10 years.

  19. San Andreas fault geometry in the Parkfield, California, region

    USGS Publications Warehouse

    Simpson, R.W.; Barall, M.; Langbein, J.; Murray, J.R.; Rymer, M.J.

    2006-01-01

    In map view, aftershocks of the 2004 Parkfield earthquake lie along a line that forms a straighter connection between San Andreas fault segments north and south of the Parkfield reach than does the mapped trace of the fault itself. A straightedge laid on a geologic map of Central California reveals a ???50-km-long asymmetric northeastward warp in the Parkfield reach of the fault. The warp tapers gradually as it joins the straight, creeping segment of the San Andreas to the north-west, but bends abruptly across Cholame Valley at its southeast end to join the straight, locked segment that last ruptured in 1857. We speculate that the San Andreas fault surface near Parkfield has been deflected in its upper ???6 km by nonelastic behavior of upper crustal rock units. These units and the fault surface itself are warped during periods between large 1857-type earthquakes by the presence of the 1857-locked segment to the south, which buttresses intermittent coseismic and continuous aseismic slip on the Parkfield reach. Because of nonelastic behavior, the warping is not completely undone when an 1857-type event occurs, and the upper portion of the three-dimensional fault surface is slowly ratcheted into an increasingly prominent bulge. Ultimately, the fault surface probably becomes too deformed for strike-slip motion, and a new, more vertical connection to the Earth's surface takes over, perhaps along the Southwest Fracture Zone. When this happens a wedge of material currently west of the main trace will be stranded on the east side of the new main trace.

  20. The eponymous Dr James Marion Sims MD, LLD (1813-1883).

    PubMed

    West, M J; Irvine, L M

    2015-02-01

    Dr James Marion Sims was born in 1813 in Lancaster County, South Carolina. It was while pioneering numerous surgical procedures in Alabama that in 1849 he achieved the outstanding landmark in medical history of successfully, and consistently, repairing vesicovaginal fistulae. Sims soon developed a reputation as a fine surgeon, with new operations and techniques, using novel surgical instruments and his innovative approaches frequently published. Moving to New York City in 1853, he further established hospitals devoted entirely to women's health. Sims was controversial, with flamboyant descriptions of self-confident success, yet they were tempered with sober reflection of failure and loss. Today we remain with the Sims speculum and Sims position, eponymous tributes to his accomplishments as the 'Father of Gynaecology'. © The Author(s) 2013 Reprints and permissions: sagepub.co.uk/journalsPermissions.nav.

  1. M ≥ 7.0 earthquake recurrence on the San Andreas fault from a stress renewal model

    USGS Publications Warehouse

    Parsons, Thomas E.

    2006-01-01

     Forecasting M ≥ 7.0 San Andreas fault earthquakes requires an assessment of their expected frequency. I used a three-dimensional finite element model of California to calculate volumetric static stress drops from scenario M ≥ 7.0 earthquakes on three San Andreas fault sections. The ratio of stress drop to tectonic stressing rate derived from geodetic displacements yielded recovery times at points throughout the model volume. Under a renewal model, stress recovery times on ruptured fault planes can be a proxy for earthquake recurrence. I show curves of magnitude versus stress recovery time for three San Andreas fault sections. When stress recovery times were converted to expected M ≥ 7.0 earthquake frequencies, they fit Gutenberg-Richter relationships well matched to observed regional rates of M ≤ 6.0 earthquakes. Thus a stress-balanced model permits large earthquake Gutenberg-Richter behavior on an individual fault segment, though it does not require it. Modeled slip magnitudes and their expected frequencies were consistent with those observed at the Wrightwood paleoseismic site if strict time predictability does not apply to the San Andreas fault.

  2. Topographically driven groundwater flow and the San Andreas heat flow paradox revisited

    USGS Publications Warehouse

    Saffer, D.M.; Bekins, B.A.; Hickman, S.

    2003-01-01

    Evidence for a weak San Andreas Fault includes (1) borehole heat flow measurements that show no evidence for a frictionally generated heat flow anomaly and (2) the inferred orientation of ??1 nearly perpendicular to the fault trace. Interpretations of the stress orientation data remain controversial, at least in close proximity to the fault, leading some researchers to hypothesize that the San Andreas Fault is, in fact, strong and that its thermal signature may be removed or redistributed by topographically driven groundwater flow in areas of rugged topography, such as typify the San Andreas Fault system. To evaluate this scenario, we use a steady state, two-dimensional model of coupled heat and fluid flow within cross sections oriented perpendicular to the fault and to the primary regional topography. Our results show that existing heat flow data near Parkfield, California, do not readily discriminate between the expected thermal signature of a strong fault and that of a weak fault. In contrast, for a wide range of groundwater flow scenarios in the Mojave Desert, models that include frictional heat generation along a strong fault are inconsistent with existing heat flow data, suggesting that the San Andreas Fault at this location is indeed weak. In both areas, comparison of modeling results and heat flow data suggest that advective redistribution of heat is minimal. The robust results for the Mojave region demonstrate that topographically driven groundwater flow, at least in two dimensions, is inadequate to obscure the frictionally generated heat flow anomaly from a strong fault. However, our results do not preclude the possibility of transient advective heat transport associated with earthquakes.

  3. Constraints on the stress state of the San Andreas Fault with analysis based on core and cuttings from San Andreas Fault Observatory at Depth (SAFOD) drilling phases 1 and 2

    USGS Publications Warehouse

    Tembe, S.; Lockner, D.; Wong, T.-F.

    2009-01-01

    Analysis of field data has led different investigators to conclude that the San Andreas Fault (SAF) has either anomalously low frictional sliding strength (?? 0.6). Arguments for the apparent weakness of the SAF generally hinge on conceptual models involving intrinsically weak gouge or elevated pore pressure within the fault zone. Some models assert that weak gouge and/or high pore pressure exist under static conditions while others consider strength loss or fluid pressure increase due to rapid coseismic fault slip. The present paper is composed of three parts. First, we develop generalized equations, based on and consistent with the Rice (1992) fault zone model to relate stress orientation and magnitude to depth-dependent coefficient of friction and pore pressure. Second, we present temperature-and pressure-dependent friction measurements from wet illite-rich fault gouge extracted from San Andreas Fault Observatory at Depth (SAFOD) phase 1 core samples and from weak minerals associated with the San Andreas Fault. Third, we reevaluate the state of stress on the San Andreas Fault in light of new constraints imposed by SAFOD borehole data. Pure talc (?????0.1) had the lowest strength considered and was sufficiently weak to satisfy weak fault heat flow and stress orientation constraints with hydrostatic pore pressure. Other fault gouges showed a systematic increase in strength with increasing temperature and pressure. In this case, heat flow and stress orientation constraints would require elevated pore pressure and, in some cases, fault zone pore pressure in excess of vertical stress. Copyright 2009 by the American Geophysical Union.

  4. On simultaneous tilt and creep observations on the San Andreas Fault

    USGS Publications Warehouse

    Johnston, M.J.S.; McHugh, S.; Burford, S.

    1976-01-01

    THE installation of an array of tiltmeters along the San Andreas Fault 1 has provided an excellent opportunity to study the amplitude and spatial scale of the tilt fields associated with fault creep. We report here preliminary results from, and some implications of, a search for interrelated surface tilts and creep event observations at four pairs of tiltmeters and creepmeters along an active 20-km stretch of the San Andreas Fault. We have observed clear creep-related tilts above the instrument resolution (10 -8 rad) only on a tiltmeter less than 0.5 km from the fault. The tilt events always preceded surface creep observations by 2-12 min, and were not purely transient in character. ?? 1975 Nature Publishing Group.

  5. NEW HOUSE EVALUATION OF POTENTIAL BUILDING DESIGN AND CONSTRUCTION FOR THE CONTROL OF RADON IN MARION AND ALACHUA COUNTIES, FLORIDA

    EPA Science Inventory

    The report gives results of a new house evaluation of potential building design and construction for the control of radon in Marion and Alachua Counties, Florida. During the project, which began in March 1992, the construction of 14 new houses, built in accordance with the Draft ...

  6. Structure of the San Andreas fault zone at SAFOD from a seismic refraction survey

    USGS Publications Warehouse

    Hole, J.A.; Ryberg, T.; Fuis, G.S.; Bleibinhaus, F.; Sharma, A.K.

    2006-01-01

    Refraction traveltimes from a 46-km long seismic survey across the San Andreas Fault were inverted to obtain two-dimensional velocity structure of the upper crust near the SAFOD drilling project. The model contains strong vertical and lateral velocity variations from <2 km/s to ???6 km/s. The Salinian terrane west of the San Andreas Fault has much higher velocity than the Franciscan terrane east of the fault. Salinian basement deepens from 0.8 km subsurface at SAFOD to ???2.5 km subsurface 20 km to the southwest. A strong reflection and subtle velocity contrast suggest a steeply dipping fault separating the Franciscan terrane from the Great Valley Sequence. A low-velocity wedge of Cenozoic sedimentary rocks lies immediately southwest of the San Andreas Fault. This body is bounded by a steep fault just northeast of SAFOD and approaches the depth of the shallowest earthquakes. Multiple active and inactive fault strands complicate structure near SAFOD. Copyright 2006 by the American Geophysical Union.

  7. A critical evaluation of crustal dehydration as the cause of an overpressured and weak San Andreas Fault

    USGS Publications Warehouse

    Fulton, P.M.; Saffer, D.M.; Bekins, B.A.

    2009-01-01

    Many plate boundary faults, including the San Andreas Fault, appear to slip at unexpectedly low shear stress. One long-standing explanation for a "weak" San Andreas Fault is that fluid release by dehydration reactions during regional metamorphism generates elevated fluid pressures that are localized within the fault, reducing the effective normal stress. We evaluate this hypothesis by calculating realistic fluid production rates for the San Andreas Fault system, and incorporating them into 2-D fluid flow models. Our results show that for a wide range of permeability distributions, fluid sources from crustal dehydration are too small and short-lived to generate, sustain, or localize fluid pressures in the fault sufficient to explain its apparent mechanical weakness. This suggests that alternative mechanisms, possibly acting locally within the fault zone, such as shear compaction or thermal pressurization, may be necessary to explain a weak San Andreas Fault. More generally, our results demonstrate the difficulty of localizing large fluid pressures generated by regional processes within near-vertical fault zones. ?? 2009 Elsevier B.V.

  8. Earthquake geology and paleoseismology of major strands of the San Andreas fault system: Chapter 38

    USGS Publications Warehouse

    Rockwell, Thomas; Scharer, Katherine M.; Dawson, Timothy E.

    2016-01-01

    The San Andreas fault system in California is one of the best-studied faults in the world, both in terms of the long-term geologic history and paleoseismic study of past surface ruptures. In this paper, we focus on the Quaternary to historic data that have been collected from the major strands of the San Andreas fault system, both on the San Andreas Fault itself, and the major subparallel strands that comprise the plate boundary, including the Calaveras-Hayward- Rogers Creek-Maacama fault zone and the Concord-Green Valley-Bartlett Springs fault zone in northern California, and the San Jacinto and Elsinore faults in southern California. The majority of the relative motion between the Pacific and North American lithospheric plates is accommodated by these faults, with the San Andreas slipping at about 34 mm/yr in central California, decreasing to about 20 mm/yr in northern California north of its juncture with the Calaveras and Concord faults. The Calaveras-Hayward-Rogers Creek-Maacama fault zone exhibits a slip rate of 10-15 mm/yr, whereas the rate along the Concord-Green Valley-Bartlett Springs fault zone is lower at about 5 mm/yr. In southern California, the San Andreas exhibits a slip rate of about 35 mm/yr along the Mojave section, decreasing to as low as 10-15 mm/yr along its juncture with the San Jacinto fault, and about 20 mm/yr in the Coachella Valley. The San Jacinto and Elsinore fault zones exhibit rates of about 15 and 5 mm/yr, respectively. The average recurrence interval for surface-rupturing earthquakes along individual elements of the San Andreas fault system range from 100-500 years and is consistent with slip rate at those sites: higher slip rates produce more frequent or larger earthquakes. There is also evidence of short-term variations in strain release (slip rate) along various fault sections, as expressed as “flurries” or clusters of earthquakes as well as periods of relatively fewer surface ruptures in these relatively short records. This

  9. San Andreas Fault, Southern California, Shaded Relief, Wrapped Color as Height

    NASA Image and Video Library

    2000-02-17

    This topographic map acquired by NASA Shuttle Radar Topography Mission SRTM from data collected on February 16, 2000 vividly displays California famous San Andreas Fault along the southwestern edge of the Mojave Desert, Calif.

  10. San Andreas Fault, Southern California , Radar Image, Wrapped Color as Height

    NASA Image and Video Library

    2000-02-17

    This topographic map acquired by NASA Shuttle Radar Topography Mission SRTM from data collected on February 16, 2000 vividly displays California famous San Andreas Fault along the southwestern edge of the Mojave Desert, Calif.

  11. Late Holocene slip rate of the San Andreas fault and its accommodation by creep and moderate-magnitude earthquakes at Parkfield, California

    USGS Publications Warehouse

    Toke, N.A.; Arrowsmith, J.R.; Rymer, M.J.; Landgraf, A.; Haddad, D.E.; Busch, M.; Coyan, J.; Hannah, A.

    2011-01-01

    Investigation of a right-laterally offset channel at the Miller's Field paleoseismic site yields a late Holocene slip rate of 26.2 +6.4/-4.3 mm/yr (1??) for the main trace of the San Andreas fault at Park-field, California. This is the first well-documented geologic slip rate between the Carrizo and creeping sections of the San Andreas fault. This rate is lower than Holocene measurements along the Carrizo Plain and rates implied by far-field geodetic measurements (~35 mm/yr). However, the rate is consistent with historical slip rates, measured to the northwest, along the creeping section of the San Andreas fault (<30 mm/yr). The paleoseismic exposures at the Miller's Field site reveal a pervasive fabric of clay shear bands, oriented clockwise oblique to the San Andreas fault strike and extending into the upper-most stratigraphy. This fabric is consistent with dextral aseismic creep and observations of surface slip from the 28 September 2004 M6 Parkfield earthquake. Together, this slip rate and deformation fabric suggest that the historically observed San Andreas fault slip behavior along the Parkfield section has persisted for at least a millennium, and that significant slip is accommodated by structures in a zone beyond the main San Andreas fault trace. ?? 2011 Geological Society of America.

  12. Tectonic history of the north portion of the San Andreas fault system, California, inferred from gravity and magnetic anomalies

    USGS Publications Warehouse

    Griscom, A.; Jachens, R.C.

    1989-01-01

    Geologic and geophysical data for the San Andreas fault system north of San Francisco suggest that the eastern boundary of the Pacific plate migrated eastward from its presumed original position at the base of the continental slope to its present position along the San Andreas transform fault by means of a series of eastward jumps of the Mendocino triple junction. These eastward jumps total a distance of about 150 km since 29 Ma. Correlation of right-laterally displaced gravity and magnetic anomalies that now have components at San Francisco and on the shelf north of Point Arena indicates that the presently active strand of the San Andreas fault north of the San Francisco peninsula formed recently at about 5 Ma when the triple junction jumped eastward a minimum of 100 km to its present location at the north end of the San Andreas fault. -from Authors

  13. Impact of a Large San Andreas Fault Earthquake on Tall Buildings in Southern California

    NASA Astrophysics Data System (ADS)

    Krishnan, S.; Ji, C.; Komatitsch, D.; Tromp, J.

    2004-12-01

    In 1857, an earthquake of magnitude 7.9 occurred on the San Andreas fault, starting at Parkfield and rupturing in a southeasterly direction for more than 300~km. Such a unilateral rupture produces significant directivity toward the San Fernando and Los Angeles basins. The strong shaking in the basins due to this earthquake would have had a significant long-period content (2--8~s). If such motions were to happen today, they could have a serious impact on tall buildings in Southern California. In order to study the effects of large San Andreas fault earthquakes on tall buildings in Southern California, we use the finite source of the magnitude 7.9 2001 Denali fault earthquake in Alaska and map it onto the San Andreas fault with the rupture originating at Parkfield and proceeding southward over a distance of 290~km. Using the SPECFEM3D spectral element seismic wave propagation code, we simulate a Denali-like earthquake on the San Andreas fault and compute ground motions at sites located on a grid with a 2.5--5.0~km spacing in the greater Southern California region. We subsequently analyze 3D structural models of an existing tall steel building designed in 1984 as well as one designed according to the current building code (Uniform Building Code, 1997) subjected to the computed ground motion. We use a sophisticated nonlinear building analysis program, FRAME3D, that has the ability to simulate damage in buildings due to three-component ground motion. We summarize the performance of these structural models on contour maps of carefully selected structural performance indices. This study could benefit the city in laying out emergency response strategies in the event of an earthquake on the San Andreas fault, in undertaking appropriate retrofit measures for tall buildings, and in formulating zoning regulations for new construction. In addition, the study would provide risk data associated with existing and new construction to insurance companies, real estate developers, and

  14. Tropical Storm Andrea June 6, 2013

    NASA Image and Video Library

    2017-12-08

    This image from the MODIS instrument aboard NASA's Terra satellite shows tropical storm Andrea on June 6, 2013, at 2:45 p.m. EDT, as the system was making landfall in the big bend area of Florida. Credit: NASA Goddard's MODIS Rapid Response Team NASA image use policy. NASA Goddard Space Flight Center enables NASA’s mission through four scientific endeavors: Earth Science, Heliophysics, Solar System Exploration, and Astrophysics. Goddard plays a leading role in NASA’s accomplishments by contributing compelling scientific knowledge to advance the Agency’s mission. Follow us on Twitter Like us on Facebook Find us on Instagram

  15. Tropical Storm Andrea June 7, 2013

    NASA Image and Video Library

    2017-12-08

    This image of tropical storm Andrea was assembled from data collected by NOAA's GOES-14 satellite at 8:31 a.m. EDT on June 7, when the storm's center was about 35 miles north-northwest of Charleston, S.C. Credit: NASA/NOAA GOES Project NASA image use policy. NASA Goddard Space Flight Center enables NASA’s mission through four scientific endeavors: Earth Science, Heliophysics, Solar System Exploration, and Astrophysics. Goddard plays a leading role in NASA’s accomplishments by contributing compelling scientific knowledge to advance the Agency’s mission. Follow us on Twitter Like us on Facebook Find us on Instagram

  16. Thermal regime of the San Andreas fault near Parkfield, California

    USGS Publications Warehouse

    Sass, J.H.; Williams, C.F.; Lachenbruch, A.H.; Galanis, S.P.; Grubb, F.V.

    1997-01-01

    Knowledge of the temperature variation with depth near the San Andreas fault is vital to understanding the physical processes that occur within the fault zone during earthquakes and creep events. Parkfield is near the southern end of the Coast Ranges segment of the San Andreas fault. This segment has higher mean heat flow than the Cape Mendocino segment to the northwest or the Mojave segment to the southeast. Boreholes were drilled specifically for the U.S. Geological Survey's Parkfield earthquake prediction experiment or converted from other uses at 25 sites within a few kilometers of the fault near Parkfield. These holes, which range in depth from 150 to over 1500 m, were intended mainly for the deployment of volumetric strain meters, water-level recorders, and other downhole instruments. Temperature profiles were obtained from all the holes, and heat flow values were estimated from 17 of them. For a number of reasons, including a paucity of thermal conductivity data and rugged local topography, the accuracy of individual determinations was not sufficiently high to document local variations in heat flow. Values range from 54 to 92 mW m-2, with mean and 95% confidence limits of 74 ?? 4 mW m-2. This mean is slightly lower than the mean (83 ?? 3) of 39 previously published values from the central Coast Ranges, but it is consistent with the overall pattern of elevated heat flow in the Coast Ranges, and it is transitional to the mean of 68 ?? 2 mW m-2 that characterizes the Mojave segment of the San Andreas fault immediately to the south. The lack of a heat flow peak near the fault underscores the absence of a frictional thermal anomaly and provides additional support for a very small resolved shear stress parallel to the San Andreas fault and the nearly fault-normal maximum compressive stress observed in this region. Estimates of subsurface thermal conditions indicate that the seismic-aseismic transition for the Parkfield segment corresponds to temperatures in the

  17. Paleomagnetic reorientation of San Andreas Fault Observatory at Depth (SAFOD) core

    USGS Publications Warehouse

    Pares, J.M.; Schleicher, A.M.; van der Pluijm, B.A.; Hickman, S.

    2008-01-01

    We present a protocol for using paleomagnetic analysis to determine the absolute orientation of core recovered from the SAFOD borehole. Our approach is based on determining the direction of the primary remanent magnetization of a spot core recovered from the Great Valley Sequence during SAFOD Phase 2 and comparing its direction to the expected reference field direction for the Late Cretaceous in North America. Both thermal and alternating field demagnetization provide equally resolved magnetization, possibly residing in magnetite, that allow reorientation. Because compositionally similar siltstones and fine-grained sandstones were encountered in the San Andreas Fault Zone during Stage 2 rotary drilling, we expect that paleomagnetic reorientation will yield reliable core orientations for continuous core acquired from directly within and adjacent to the San Andreas Fault during SAFOD Phase 3, which will be key to interpretation of spatial properties of these rocks. Copyright 2008 by the American Geophysical Union.

  18. Assessment of the Old Red Rock Indian Line Sycamore Tree, Lake Red Rock, Marion County, Iowa

    DTIC Science & Technology

    1992-01-01

    miles, when reduced to a straight line , from the junction of the White Breast and Des Moines (Stiles 1911:4). George W. Harrison was instructed to...AD-A255 372 Assessment of the Old Red Rock Indian Line Sycamore Tree, Lake Red Rock, Marion County, Iowa DACW25-92-M-0414 by Leah D. Rogers Project...portion of tree 22 9. Map showing location of Red Rock line within treaty cession area of 23 1842 10. Portion of 1844 map showing incorrect placement of

  19. Heat flow and energetics of the San Andreas fault zone.

    USGS Publications Warehouse

    Lachenbruch, A.H.; Sass, J.H.

    1980-01-01

    Approximately 100 heat flow measurements in the San Andreas fault zone indicate 1) there is no evidence for local frictional heating of the main fault trace at any latitude over a 1000-km length from Cape Mendocino to San Bernardino, 2) average heat flow is high (ca.2 HFU, ca.80 mW m-2) throughout the 550-km segment of the Coast Ranges that encloses the San Andreas fault zone in central California; this broad anomaly falls off rapidly toward the Great Valley to the east, and over a 200-km distance toward the Mendocino Triple Junction to the northwest. As others have pointed out, a local conductive heat flow anomaly would be detectable unless the frictional resistance allocated to heat production on the main trace were less than 100 bars. Frictional work allocated to surface energy of new fractures is probably unimportant, and hydrologic convection is not likely to invalidate the conduction assumption, since the heat discharge by thermal springs near the fault is negligible. -Authors

  20. Scientific drilling into the San Andreas Fault Zone - an overview of SAFOD's first five years

    USGS Publications Warehouse

    Zoback, Mark; Hickman, Stephen; Ellsworth, William; ,

    2011-01-01

    The San Andreas Fault Observatory at Depth (SAFOD) was drilled to study the physical and chemical processes controlling faulting and earthquake generation along an active, plate-bounding fault at depth. SAFOD is located near Parkfield, California and penetrates a section of the fault that is moving due to a combination of repeating microearthquakes and fault creep. Geophysical logs define the San Andreas Fault Zone to be relatively broad (~200 m), containing several discrete zones only 2–3 m wide that exhibit very low P- and S-wave velocities and low resistivity. Two of these zones have progressively deformed the cemented casing at measured depths of 3192 m and 3302 m. Cores from both deforming zones contain a pervasively sheared, cohesionless, foliated fault gouge that coincides with casing deformation and explains the observed extremely low seismic velocities and resistivity. These cores are being now extensively tested in laboratories around the world, and their composition, deformation mechanisms, physical properties, and rheological behavior are studied. Downhole measurements show that within 200 m (maximum) of the active fault trace, the direction of maximum horizontal stress remains at a high angle to the San Andreas Fault, consistent with other measurements. The results from the SAFOD Main Hole, together with the stress state determined in the Pilot Hole, are consistent with a strong crust/weak fault model of the San Andreas. Seismic instrumentation has been deployed to study physics of faulting—earthquake nucleation, propagation, and arrest—in order to test how laboratory-derived concepts scale up to earthquakes occurring in nature.

  1. Dipping San Andreas and Hayward faults revealed beneath San Francisco Bay, California

    USGS Publications Warehouse

    Parsons, T.; Hart, P.E.

    1999-01-01

    The San Francisco Bay area is crossed by several right-lateral strike-slip faults of the San Andreas fault zone. Fault-plane reflections reveal that two of these faults, the San Andreas and Hayward, dip toward each other below seismogenic depths at 60?? and 70??, respectively, and persist to the base of the crust. Previously, a horizontal detachment linking the two faults in the lower crust beneath San Francisco Bay was proposed. The only near-vertical-incidence reflection data available prior to the most recent experiment in 1997 were recorded parallel to the major fault structures. When the new reflection data recorded orthogonal to the faults are compared with the older data, the highest, amplitude reflections show clear variations in moveout with recording azimuth. In addition, reflection times consistently increase with distance from the faults. If the reflectors were horizontal, reflection moveout would be independent of azimuth, and reflection times would be independent of distance from the faults. The best-fit solution from three-dimensional traveltime modeling is a pair of high-angle dipping surfaces. The close correspondence of these dipping structures with the San Andreas and Hayward faults leads us to conclude that they are the faults beneath seismogenic depths. If the faults retain their observed dips, they would converge into a single zone in the upper mantle -45 km beneath the surface, although we can only observe them in the crust.

  2. Neogene contraction between the San Andreas fault and the Santa Clara Valley, San Francisco Bay region, California

    USGS Publications Warehouse

    McLaughlin, R.J.; Langenheim, V.E.; Schmidt, K.M.; Jachens, R.C.; Stanley, R.G.; Jayko, A.S.; McDougall, K.A.; Tinsley, J.C.; Valin, Z.C.

    1999-01-01

    In the southern San Francisco Bay region of California, oblique dextral reverse faults that verge northeastward from the San Andreas fault experienced triggered slip during the 1989 M7.1 Loma Prieta earthquake. The role of these range-front thrusts in the evolution of the San Andreas fault system and the future seismic hazard that they may pose to the urban Santa Clara Valley are poorly understood. Based on recent geologic mapping and geophysical investigations, we propose that the range-front thrust system evolved in conjunction with development of the San Andreas fault system. In the early Miocene, the region was dominated by a system of northwestwardly propagating, basin-bounding, transtensional faults. Beginning as early as middle Miocene time, however, the transtensional faulting was superseded by transpressional NE-stepping thrust and reverse faults of the range-front thrust system. Age constraints on the thrust faults indicate that the locus of contraction has focused on the Monte Vista, Shannon, and Berrocal faults since about 4.8 Ma. Fault slip and fold reconstructions suggest that crustal shortening between the San Andreas fault and the Santa Clara Valley within this time frame is ~21%, amounting to as much as 3.2 km at a rate of 0.6 mm/yr. Rates probably have not remained constant; average rates appear to have been much lower in the past few 100 ka. The distribution of coseismic surface contraction during the Loma Prieta earthquake, active seismicity, late Pleistocene to Holocene fluvial terrace warping, and geodetic data further suggest that the active range-front thrust system includes blind thrusts. Critical unresolved issues include information on the near-surface locations of buried thrusts, the timing of recent thrust earthquake events, and their recurrence in relation to earthquakes on the San Andreas fault.

  3. Reflections on Comparative Education: Telling Tales in Honor of Andreas Kazamias

    ERIC Educational Resources Information Center

    Cowen, Robert

    2018-01-01

    Emphasizing the important role of "history" within comparative education is the classic way, much celebrated in the writings of Andreas Kazamias, to treat this theme. This article uses a different perspective. The argument is that "comparative education" and "history" use two words as professional identifiers of a way…

  4. Base and precious metal occurrences along the San Andreas Fault, Point Delgada, California

    USGS Publications Warehouse

    McLaughlin, Robert J.; Sorg, D.H.; Ohlin, H.N.; Heropoulos, Chris

    1979-01-01

    Previously unrecognized veins containing lead, zinc, and copper sulfide minerals at Point Delgada, Calif., are associated with late Mesozoic(?) and Tertiary volcanic and sedimentary rocks of the Franciscan assemblage. Sulfide minerals include pyrite, sphalerite, galena, and minor chalcopyrite, and galena-rich samples contain substantial amounts of silver. These minerals occur in a quartz-carbonate gangue along northeast-trending faults and fractures that exhibit (left?) lateral and vertical slip. The sense of fault movement and the northeasterly strike are consistent with predicted conjugate fault sets of the present San Andreas fault system. The sulfide mineralization is younger than the Franciscan rocks of Point Delgada and King Range, and it may have accompanied or postdated the inception of San Andreas faulting. Mineralization largely preceded uplift, the formation of a marine terrace, and the emplacement of landslide-related debris-flow breccias that overlie the mineralized rocks and truncate the sulfide veins. These field relations indicate that the sulfide mineralization and inception of San Andreas faulting were clearly more recent than the early Miocene and that the mineralization could be younger than about 1.2 m.y. The sulfide veins at Point Delgada may be of economic significance. However, prior to any exploitation of the occurrence, economic and environmental conflicts of interest involving private land ownership, the Shelter Cove home development, and proximity of the coast must be resolved.

  5. Kinematic evolution of the junction of the San Andreas, Garlock, and Big Pine faults, California

    USGS Publications Warehouse

    Bohannon, Robert G.; Howell, David G.

    1982-01-01

    If the San Andreas fault with about 300 km of right slip, the Carlock fault with about 60 km of left slip, and the Big Pine fault with about 15 km of left slip are considered to have been contemporaneously active, a space problem at their high-angle junctions becomes apparent. Large crustal masses converge in the area of the junctions as a result of the simultaneous large displacements on the faults. We present here a model in which an early straight north-northwest–trending San Andreas deforms to its present bent configuration in response to a westward displacement of crust north of the Garlock fault. During this deformation, the crust north of the Garlock in the vicinity of the junction undergoes north-south shortening, while the fault junction migrates along the trace of the San Andreas fault to the southeast relative to its original position. As a result of this migration, the Mojave area is displaced to the east relative to the original junction position. We suggest a similar history in mirror image for the Big Pine fault and the areas of crust adjacent to it.

  6. Oil production from the Ste. Genevieve limestone in the exchange area, Marion County, Illinois

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Stevenson, D.L.

    1969-01-01

    The 4 townships making up the SE. quarter of Marion County, Illinois, contain 7 relatively small oil pools. The major portion of the oil produced from these pools was trapped in oolitic limestones and sandstones of the Ste. Genevieve Formation. The oil appears to be accumulated in stratigraphic traps, as no significant structural closure is evident. A reconstruction of the geologic history of the area, aided by the technique of trend-surface fitting, suggests that structural folding was, however, an important factor in trapping the oil. Subsequent tilting of the area has removed the closure of the structures, but the oilmore » did not escape because of permeability pinch outs of the reservoir beds.« less

  7. The San Andreas fault in the San Francisco Bay region, California: Structure and kinematics of a Young plate boundary

    USGS Publications Warehouse

    Jachens, R.C.; Zoback, M.L.

    1999-01-01

    Recently acquired high-resolution aeromagnetic data delineate offset and/or truncated magnetic rock bodies of the Franciscan Complex that define the location and structure of, and total offset across, the San Andreas fault in the San Francisco Bay region. Two distinctive magnetic anomalies caused by ultramafic rocks and metabasalts east of, and truncated at, the San Andreas fault have clear counterparts west of the fault that indicate a total right-lateral offset of only 22 km on the Peninsula segment, the active strand that ruptured in 1906. The location of the Peninsula segment is well defined magnetically on the northern peninsula where it goes offshore, and can be traced along strike an additional ~6 km to the northwest. Just offshore from Lake Merced, the inferred fault trace steps right (northeast) 3 km onto a nearly parallel strand that can be traced magnetically northwest more than 20 km as the linear northeast edge of a magnetic block bounded by the San Andreas fault, the Pilarcitos fault, and the San Gregorio-Hosgri fault zone. This right-stepping strand, the Golden Gate segment, joins the eastern mapped trace of the San Andreas fault at Bolinas Lagoon and projects back onshore to the southeast near Lake Merced. Inversion of detailed gravity data on the San Francisco Peninsula reveals a 3 km wide basin situated between the two strands of the San Andreas fault, floored by Franciscan basement and filled with Plio-Quaternary sedimentary deposits of the Merced and Colma formations. The basin, ~1 km deep at the coast, narrows and becomes thinner to the southeast along the fault over a distance of ~12 km. The length, width, and location of the basin between the two strands are consistent with a pull-apart basin formed behind the right step in the right-lateral strike-slip San Andreas fault system and currently moving southeast with the North American plate. Slight nonparallelism of the two strands bounding the basin (implying a small component of convergence

  8. Evaluation of hypotheses for right-lateral displacement of Neogene strata along the San Andreas Fault between Parkfield and Maricopa, California

    USGS Publications Warehouse

    Stanley, Richard G.; Barron, John A.; Powell, Charles L.

    2017-12-22

    We used geological field studies and diatom biostratigraphy to test a published hypothesis that Neogene marine siliceous strata in the Maricopa and Parkfield areas, located on opposite sides of the San Andreas Fault, were formerly contiguous and then were displaced by about 80–130 kilometers (km) of right-lateral slip along the fault. In the Maricopa area on the northeast side of the San Andreas Fault, the upper Miocene Bitterwater Creek Shale consists of hard, siliceous shale with dolomitic concretions and turbidite sandstone interbeds. Diatom assemblages indicate that the Bitterwater Creek Shale was deposited about 8.0–6.7 million years before present (Ma) at the same time as the uppermost part of the Monterey Formation in parts of coastal California. In the Parkfield area on the southwest side of the San Andreas Fault, the upper Miocene Pancho Rico Formation consists of soft to indurated mudstone and siltstone and fossiliferous, bioturbated sandstone. Diatom assemblages from the Pancho Rico indicate deposition about 6.7–5.7 Ma (latest Miocene), younger than the Bitterwater Creek Shale and at about the same time as parts of the Sisquoc Formation and Purisima Formation in coastal California. Our results show that the Bitterwater Creek Shale and Pancho Rico Formation are lithologically unlike and of different ages and therefore do not constitute a cross-fault tie that can be used to estimate rightlateral displacement along the San Andreas Fault.In the Maricopa area northeast of the San Andreas Fault, the Bitterwater Creek Shale overlies conglomeratic fan-delta deposits of the upper Miocene Santa Margarita Formation, which in turn overlie siliceous shale of the Miocene Monterey Formation from which we obtained a diatom assemblage dated at about 10.0–9.3 Ma. Previous investigations noted that the Santa Margarita Formation in the Maricopa area contains granitic and metamorphic clasts derived from sources in the northern Gabilan Range, on the opposite side of

  9. The response of creeping parts of the San Andreas fault to earthquakes on nearby faults: Two examples

    USGS Publications Warehouse

    Simpson, R.W.; Schulz, S.S.; Dietz, L.D.; Burford, R.O.

    1988-01-01

    Rates of shallow slip on creeping sections of the San Andreas fault have been perturbed on a number of occasions by earthquakes occurring on nearby faults. One example of such perturbations occurred during the 26 January 1986 magnitude 5.3 Tres Pinos earthquake located about 10 km southeast of Hollister, California. Seven creepmeters on the San Andreas fault showed creep steps either during or soon after the shock. Both left-lateral (LL) and right-lateral (RL) steps were observed. A rectangular dislocation in an elastic half-space was used to model the coseismic fault offset at the hypocenter. For a model based on the preliminary focal mechanism, the predicted changes in static shear stress on the plane of the San Andreas fault agreed in sense (LL or RL) with the observed slip directions at all seven meters; for a model based on a refined focal mechanism, six of the seven meters showed the correct sense of motion. Two possible explanations for such coseismic and postseismic steps are (1) that slip was triggered by the earthquake shaking or (2) that slip occurred in response to the changes in static stress fields accompanying the earthquake. In the Tres Pinos example, the observed steps may have been of both the triggered and responsive kinds. A second example is provided by the 2 May 1983 magnitude 6.7 Coalinga earthquake, which profoundly altered slip rates at five creepmeters on the San Andreas fault for a period of months to years. The XMM1 meter 9 km northwest of Parkfield, California recorded LL creep for more than a year after the event. To simulate the temporal behavior of the XMM1 meter and to view the stress perturbation provided by the Coalinga earthquake in the context of steady-state deformation on the San Andreas fault, a simple time-evolving dislocation model was constructed. The model was driven by a single long vertical dislocation below 15 km in depth, that was forced to slip at 35 mm/yr in a RL sense. A dislocation element placed in the

  10. Retention time and flow patterns in Lake Marion, South Carolina, 1984

    USGS Publications Warehouse

    Patterson, G.G.; Harvey, R.M.

    1995-01-01

    In 1984, six dye tracer tests were made on Lake Marion to determine flow patterns and retention times under conditions of high and low flow. During the high-flow tests, with an average inflow of about 29,000 cubic feet per second, the approximate travel time through the lake for the peak tracer concentration was 14 days. The retention time was about 20 days. During the low-flow tests, with an average inflow of about 9,000 cubic feet per second, the approximate travel time was 41 days, and the retention time was about 60 days. The primary factors controlling movement of water in the lake are lake inflow and outflow. The tracer cloud moved consistently downstream, slowing as the lake widened. Flow patterns in most of the coves, and in some areas along the northeastern shore, are influenced more by tributary inflow than by factors attributable to water from the main body of the lake.

  11. Lead and cadmium in the blood of nine species of seabirds, Marion Island, South Africa.

    PubMed

    Summers, Carly F; Bowerman, William W; Parsons, Nola; Chao, Wayne Y; Bridges, William C

    2014-10-01

    Levels of lead (Pb) and cadmium (Cd) were investigated as potential stressors in nine species of breeding seabirds on Marion Island, South Africa. The majority of blood Pb levels (95 %) were below background exposure levels. Species was a significant factor in ranked means analysis for mean blood Pb levels. Fewer individual blood Cd levels (<60 %) were within background exposure levels and species was not significant. Elevated levels of Cd have been documented in other seabird species without apparent outward effects, which suggests that seabirds may be adapted to high cadmium environments, particularly from their diets. Overall, the results suggest Pb and Cd are not primary causes for concern in these seabirds.

  12. Geophysical Surveys of the San Andreas and Crystal Springs Reservoir System Including Seismic-Reflection Profiles and Swath Bathymetry, San Mateo County, California

    USGS Publications Warehouse

    Finlayson, David P.; Triezenberg, Peter J.; Hart, Patrick E.

    2010-01-01

    This report describes geophysical data acquired by the U.S. Geological Survey (USGS) in San Andreas Reservoir and Upper and Lower Crystal Springs Reservoirs, San Mateo County, California, as part of an effort to refine knowledge of the location of traces of the San Andreas Fault within the reservoir system and to provide improved reservoir bathymetry for estimates of reservoir water volume. The surveys were conducted by the Western Coastal and Marine Geology (WCMG) Team of the USGS for the San Francisco Public Utilities Commission (SFPUC). The data were acquired in three separate surveys: (1) in June 2007, personnel from WCMG completed a three-day survey of San Andreas Reservoir, collecting approximately 50 km of high-resolution Chirp subbottom seismic-reflection data; (2) in November 2007, WCMG conducted a swath-bathymetry survey of San Andreas reservoir; and finally (3) in April 2008, WCMG conducted a swath-bathymetry survey of both the upper and lower Crystal Springs Reservoir system. Top of PageFor more information, contact David Finlayson.

  13. ["... I shall never forget the gift by which you established yourself as friend in my life!" The letters of Lou Andreas-Salomé to Max Eitingon (1911-1933)].

    PubMed

    Weber, Inge

    2015-01-01

    The correspondence between Andreas-Salomé and the Eitingons draws attention to their long-standing relation. The letters contained among the Eitingon papers in Jerusalem (81 items) were complemented by the much smaller set (5 items) held by the Lou Andreas-Salomé Archives in Göttingen. The material highlights for the first time Eitingon's role in securing Andreas-Salomé's access to the Berlin psychoanalytic association and for her entering psychoanalytic practice. In the 20s the relation between Andreas-Salomé and Mirra Eitingon intensified, based on their common Russian background. Several aspects featured in the letters are discussed in appendixes: the role of Russian language and habits; Max Nachmansohn, an analysand of Andreas-Salomé; her literary gift to Freud's 70th birthday; the dealing with fees in psychoanalysis.

  14. Andrea’s disease (angiomegaly): a currently well-defined nosological entitys.

    PubMed

    Taurone, S; Spoletini, M; Di Matteo, F M; Mele, R; Tromba, L; Grippaudo, F R; Minni, A; Artico, M

    2017-01-01

    In 1997 D’Andrea et al. described a new nosological entity the characteristics of which consisted of lengthening, dilation and tortuosity of blood vessels, arteries or veins, less prominent, but also less circumscribed than an aneurysm. This condition does not necessarily imply specific aneurysm formation although aneurysms at multiple sites are a frequent observation. The term used by authors for angiomegaly of the venous system was venomegaly and the analogous condition of the arterial system was termed arteriomegaly. Although tortuosity and dilation of arteries and veins have been widely reported, suggesting a systemic disorder which affects the structural integrity of all vessels, most papers dealing with this intriguing condition did not describe any alterations in the components of vessel walls. In the present paper, the authors describe a well-defined condition, D’Andrea’s Disease (or DD, in this article), analyzing its salient morphological and clinical features and clarifying this pathological condition as a distinct and now well-defined nosological entity.

  15. Flood-inundation maps for the Mississinewa River at Marion, Indiana, 2013

    USGS Publications Warehouse

    Coon, William F.

    2014-01-01

    Digital flood-inundation maps for a 9-mile (mi) reach of the Mississinewa River from 0.75 mi upstream from the Pennsylvania Street bridge in Marion, Indiana, to 0.2 mi downstream from State Route 15 were created by the U.S. Geological Survey (USGS) in cooperation with the Indiana Office of Community and Rural Affairs. The flood inundation maps, which can be accessed through the USGS Flood Inundation Mapping Science Web site at http://water.usgs.gov/osw/flood_inundation, depict estimates of the areal extent and depth of flooding corresponding to selected water levels (stages) at the USGS streamgage on the Mississinewa River at Marion (station number 03326500). Near-real-time stages at this streamgage may be obtained on the Internet from the USGS National Water Information System at http://waterdata.usgs.gov/ or the National Weather Service (NWS) Advanced Hydrologic Prediction Service at http://water.weather.gov/ahps/, which also forecasts flood hydrographs at this site. Flood profiles were computed for the stream reach by means of a one-dimensional step-backwater model. The model was calibrated by using the current stage-discharge relation at the Mississinewa River streamgage, in combination with water-surface profiles from historic floods and from the current (2002) flood-insurance study for Grant County, Indiana. The hydraulic model was then used to compute seven water-surface profiles for flood stages at 1-fo (ft) intervals referenced to the streamgage datum and ranging from 10 ft, which is near bankfull, to 16 ft, which is between the water levels associated with the estimated 10- and 2-percent annual exceedance probability floods (floods with recurrence interval between 10 and 50 years) and equals the “major flood stage” as defined by the NWS. The simulated water-surface profiles were then combined with a Geographic Information System digital elevation model (derived from light detection and ranging (lidar) data having a 0.98 ft vertical accuracy and 4.9 ft

  16. Northern California LIDAR Data: A Tool for Mapping the San Andreas Fault and Pleistocene Marine Terraces in Heavily Vegetated Terrain

    NASA Astrophysics Data System (ADS)

    Prentice, C. S.; Crosby, C. J.; Harding, D. J.; Haugerud, R. A.; Merritts, D. J.; Gardner, T. W.; Koehler, R. D.; Baldwin, J. N.

    2003-12-01

    Recent acquisition of airborne LIDAR (also known as ALSM) data covering approximately 418 square kilometers of coastal northern California provides a powerful new tool for mapping geomorphic features related to the San Andreas Fault and coastal uplift. LIDAR data has been previously used in the Puget Lowland region of Washington to identify and map Holocene faults and uplifted shorelines concealed under dense vegetation (Haugerud et al., 2003; see http://pugetsoundlidar.org). Our effort represents the first use of LIDAR data for this purpose along the San Andreas Fault. This data set is the result of a collaborative effort between NASA Solid Earth and Natural Hazards Program, Goddard Space Flight Center, Stennis Space Center, USGS, and TerraPoint, LLC. The coverage extends from near Fort Ross, California, in Sonoma County, along the coast northward to the town of Mendocino, in Mendocino County, and as far inland as about 1-3 km east of the San Andreas Fault. The survey area includes about 70 km of the northern San Andreas Fault under dense redwood forest, and Pleistocene coastal marine terraces both north and south of the fault. The average data density is two laser pulses per square meter, with up to four LIDAR returns per pulse. Returns are classified as ground or vegetation, allowing construction of both canopy-top and bare-earth DEMs with 1.8m grid spacing. Vertical accuracy is better than 20 cm RMSE, confirmed by a network of ground-control points established using high-precision GPS surveying. We are using hillshade images generated from the bare-earth DEMs to begin detailed mapping of geomorphic features associated with San Andreas Fault traces, such as scarps, offset streams, linear valleys, shutter ridges, and sag ponds. In addition, we are using these data in conjunction with field mapping and interpretation of conventional 1:12,000 and 1:6000 scale aerial photographs to map and correlate marine terraces to better understand rates of coastal uplift, and

  17. Holocene Geologic Slip Rate for the Banning Strand of the Southern San Andreas Fault near San Gorgonio Pass, Southern California

    NASA Astrophysics Data System (ADS)

    Gold, P. O.; Behr, W. M.; Rood, D. H.; Kendrick, K. J.; Rockwell, T. K.; Sharp, W. D.

    2014-12-01

    We present the first Holocene geologic slip rate for the Banning strand of the southern San Andreas Fault in southern California. The southern San Andreas Fault splays into the sub-parallel Banning and Mission Creek strands in the northwestern Coachella Valley, and although it has long been surmised that the Banning strand eventually accommodates the majority of displacement and transfers it into San Gorgonio Pass, until now it has been uncertain how slip is actually partitioned between these two fault strands. Our new slip rate measurement, critically located at the northwestern end of the Banning strand, overlaps within errors with the published rate for the southern San Andreas Fault measured at Biskra Palms Oasis. This indicates that the majority of southern San Andreas Fault displacement transfers from the southeastern Mission Creek strand northwest to the Banning strand and into San Gorgonio Pass. Our result corroborates the UCERF3 hazard model, and is consistent with most previous interpretations of how slip is partitioned between the Banning and Mission Creek fault strands. To measure this slip rate, we used B4 airborne LiDAR to identify the apex of an alluvial fan offset laterally 30 ± 5 m from its source. We calculated the depositional age of the fan using 10Be in-situ cosmogenic exposure dating of 5 cobbles and a depth profile. We calculated a most probable fan age of 4.0 +2.0/-1.6 ka (1σ) by combining the inheritance-corrected cobble ages assuming Gaussian uncertainty. However, the probability density function yielded a multi-peaked distribution, which we attribute to variable 10Be inheritance in the cobbles, so we favor the depth profile age of 2.2-3.6 ka. Combined, these measurements yield a late Holocene slip rate for the Banning strand of the southern San Andreas Fault of 11.1 +3.1/-3.3 mm/yr. This slip rate does not preclude possibility that some slip transfers north along the Mission Creek strand and the Garnet Hill fault, but it does confirm

  18. Interferometric imaging of the San Andreas Fault at Parkfield Using a Massive VSP

    NASA Astrophysics Data System (ADS)

    Chavarria, J.; Goertz, A.; Karrenbach, M.; Paulsson, B.

    2006-12-01

    The San Andreas Fault Observatory at Depth (SAFOD) has the goal of investigating the physical processes controlling earthquakes in central California at Parkfield. The observatory consists of a deep well, that intersected the San Andreas Fault at seismogenic depths of ~3.2 km, where recurring microearthquake activity takes place. Previous to the drilling of Phase II, that intersected the fault at the end of Summer 2005, a long array of seismometers was deployed in the deeper part of the well to characterize the fault structure and to aid the monitoring activities of the natural seismicity. The Massive VSP array consisted of 80 three component seismometers that were deployed during April and May 2005. During this period of time we recorded active data from 13 explosions used to refine the velocity models that improved the locations of target events. In addition to this, we continuously monitored passive data from the natural earthquake activity in the area that was dominated by the Parkfield 2004 event aftershock sequence. During the recording time of this project we were able to record one of the target earthquakes of the SAFOD project as well as recently discovered non-volcanic tremor. The data recorded by this deep array of seismometers has provided important information to characterize the structure of the fault at scale that is comparable to the scale of the geologic observations. These observations, derived during drilling of the well, consist of cores and cuttings as well as geophysical logs. With the use of the borehole array of seismometers we have been able to directly correlate this lithologic information with seismic phases observed in the VSP data. This has lead to better understanding the nature of the scattering that takes place in this complex fault zone. Seismic reflections in the dataset were analyzed with Kirchhoff imaging methods to determine a partial image of the San Andreas Fault at depth. Due to the low density distribution of sources in the

  19. San Andreas Fault in the Carrizo Plain

    NASA Technical Reports Server (NTRS)

    2000-01-01

    The 1,200-kilometer (800-mile)San Andreas is the longest fault in California and one of the longest in North America. This perspective view of a portion of the fault was generated using data from the Shuttle Radar Topography Mission (SRTM), which flew on NASA's Space Shuttle last February, and an enhanced, true-color Landsat satellite image. The view shown looks southeast along the San Andreas where it cuts along the base of the mountains in the Temblor Range near Bakersfield. The fault is the distinctively linear feature to the right of the mountains. To the left of the range is a portion of the agriculturally rich San Joaquin Valley. In the background is the snow-capped peak of Mt. Pinos at an elevation of 2,692 meters (8,831 feet). The complex topography in the area is some of the most spectacular along the course of the fault. To the right of the fault is the famous Carrizo Plain. Dry conditions on the plain have helped preserve the surface trace of the fault, which is scrutinized by both amateur and professional geologists. In 1857, one of the largest earthquakes ever recorded in the United States occurred just north of the Carrizo Plain. With an estimated magnitude of 8.0, the quake severely shook buildings in Los Angeles, caused significant surface rupture along a 350-kilometer (220-mile) segment of the fault, and was felt as far away as Las Vegas, Nev. This portion of the San Andreas is an important area of study for seismologists. For visualization purposes, topographic heights displayed in this image are exaggerated two times.

    The elevation data used in this image was acquired by SRTM aboard the Space Shuttle Endeavour, launched on February 11, 2000. SRTM used the same radar instrument that comprised the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) that flew twice on Endeavour in 1994. SRTM was designed to collect three-dimensional measurements of Earth's land surface. To collect the 3-D SRTM data, engineers added a mast 60

  20. Monitoring of hydrogen along the San Andreas and Calaveras faults in central California in 1980-1984

    NASA Astrophysics Data System (ADS)

    Sato, Motoaki; Sutton, A. J.; McGee, K. A.; Russell-Robinson, Susan

    1986-11-01

    Hydrogen (H2) has been monitored continuously at 1.5-m depth at nine sites along the San Andreas and Calaveras faults in central California since December 1980. Site characteristic small noninstrumental diurnal variations were recorded during quiescent periods at most sites. Abrupt H2 changes were observed concurrently at two sites on the Calaveras fault; some of these were correlated with oscillatory fault slips. Large (1000-4000 ppm) H2 increases were recorded at some sites on the San Andreas fault between July 1982 and November 1983, which may be correlated with eleven M ≥ 5 earthquakes that occurred near Coalinga during this period. We attribute both the H2 increases and the triggering of the earthquakes to a large-scale compressive stress field within the ductile mafic crust near the plate boundary. The stress perhaps caused bulging of the base of the brittle upper crust and thus caused dilation of the San Andreas fault zone, allowing the escape of pent-up H2 generated by hydration reaction of the mafic crust. At the same time, mobile serpentinites may have squeezed into the seismogenic fault beneath the Coalinga area triggering the earthquakes.

  1. Evolving transpressional strain fields along the San Andreas fault in southern California: implications for fault branching, fault dip segmentation and strain partitioning

    NASA Astrophysics Data System (ADS)

    Bergh, Steffen; Sylvester, Arthur; Damte, Alula; Indrevær, Kjetil

    2014-05-01

    The San Andreas fault in southern California records only few large-magnitude earthquakes in historic time, and the recent activity is confined primarily on irregular and discontinuous strike-slip and thrust fault strands at shallow depths of ~5-20 km. Despite this fact, slip along the San Andreas fault is calculated to c. 35 mm/yr based on c.160 km total right lateral displacement for the southern segment of the fault in the last c. 8 Ma. Field observations also reveal complex fault strands and multiple events of deformation. The presently diffuse high-magnitude crustal movements may be explained by the deformation being largely distributed along more gently dipping reverse faults in fold-thrust belts, in contrast to regions to the north where deformation is less partitioned and localized to narrow strike-slip fault zones. In the Mecca Hills of the Salton trough transpressional deformation of an uplifted segment of the San Andreas fault in the last ca. 4.0 My is expressed by very complex fault-oblique and fault-parallel (en echelon) folding, and zones of uplift (fold-thrust belts), basement-involved reverse and strike-slip faults and accompanying multiple and pervasive cataclasis and conjugate fracturing of Miocene to Pleistocene sedimentary strata. Our structural analysis of the Mecca Hills addresses the kinematic nature of the San Andreas fault and mechanisms of uplift and strain-stress distribution along bent fault strands. The San Andreas fault and subsidiary faults define a wide spectrum of kinematic styles, from steep localized strike-slip faults, to moderate dipping faults related to oblique en echelon folds, and gently dipping faults distributed in fold-thrust belt domains. Therefore, the San Andreas fault is not a through-going, steep strike-slip crustal structure, which is commonly the basis for crustal modeling and earthquake rupture models. The fault trace was steep initially, but was later multiphase deformed/modified by oblique en echelon folding

  2. Monitoring microearthquakes with the San Andreas fault observatory at depth

    USGS Publications Warehouse

    Oye, V.; Ellsworth, W.L.

    2007-01-01

    In 2005, the San Andreas Fault Observatory at Depth (SAFOD) was drilled through the San Andreas Fault zone at a depth of about 3.1 km. The borehole has subsequently been instrumented with high-frequency geophones in order to better constrain locations and source processes of nearby microearthquakes that will be targeted in the upcoming phase of SAFOD. The microseismic monitoring software MIMO, developed by NORSAR, has been installed at SAFOD to provide near-real time locations and magnitude estimates using the high sampling rate (4000 Hz) waveform data. To improve the detection and location accuracy, we incorporate data from the nearby, shallow borehole (???250 m) seismometers of the High Resolution Seismic Network (HRSN). The event association algorithm of the MIMO software incorporates HRSN detections provided by the USGS real time earthworm software. The concept of the new event association is based on the generalized beam forming, primarily used in array seismology. The method requires the pre-computation of theoretical travel times in a 3D grid of potential microearthquake locations to the seismometers of the current station network. By minimizing the differences between theoretical and observed detection times an event is associated and the location accuracy is significantly improved.

  3. ["... here I am entirely among patients now..": the psychoanalytical practice of Lou Andreas-Salomé].

    PubMed

    Klemann, Manfred

    2005-01-01

    The aim of this article is to disprove the widespread prejudice depicting Andreas-Salomé merely as a femme fatale, or companion of a few famous contemporaries (Nietzsche, Rilke, and Freud), while suppressing her original intellectual and clinical-practical achievement as a psychoanalyst. An evaluation of both published and hitherto unpublished sources clearly confirms the broad and thorough foundations of her psychoanalytical training in theory as well as in practice. Between 1913 and 1933 Andreas-Salomé conducted a relatively large number of analyses, discussed some of them with Freud in a kind of "supervision" by correspondence and published several articles on central psychoanalytical issues. So far, however, many psychoanalysts seem to have been unaware of her status as a former accomplished colleague.

  4. The San Andreas Fault in the San Francisco Bay area, California: a geology fieldtrip guidebook to selected stops on public lands

    USGS Publications Warehouse

    Stoffer, Philip W.

    2005-01-01

    This guidebook contains a series of geology fieldtrips with selected destinations along the San Andreas Fault in part of the region that experienced surface rupture during the Great San Francisco Earthquake of 1906. Introductory materials present general information about the San Andreas Fault System, landscape features, and ecological factors associated with faults in the South Bay, Santa Cruz Mountains, the San Francisco Peninsula, and the Point Reyes National Seashore regions. Trip stops include roadside areas and recommended hikes along regional faults and to nearby geologic and landscape features that provide opportunities to make casual observations about the geologic history and landscape evolution. Destinations include the sites along the San Andreas and Calaveras faults in the San Juan Bautista and Hollister region. Stops on public land along the San Andreas Fault in the Santa Cruz Mountains in Santa Clara and Santa Cruz counties include in the Loma Prieta summit area, Forest of Nicene Marks State Park, Lexington County Park, Sanborn County Park, Castle Rock State Park, and the Mid Peninsula Open Space Preserve. Destinations on the San Francisco Peninsula and along the coast in San Mateo County include the Crystal Springs Reservoir area, Mussel Rock Park, and parts of Golden Gate National Recreation Area, with additional stops associated with the San Gregorio Fault system at Montara State Beach, the James F. Fitzgerald Preserve, and at Half Moon Bay. Field trip destinations in the Point Reyes National Seashore and vicinity provide information about geology and character of the San Andreas Fault system north of San Francisco.

  5. Andreas Vesalius on the anatomy and function of the lower thoracic vertebrae.

    PubMed

    Biesbrouck, Maurits; Vanden Berghe, Alex

    2016-04-01

    Some remarkable statements made by Andreas Vesalius (1514-1564) in his principal work De Humani Corporis Fabrica (1543) about the anatomy and function of the lower thoracic vertebrae are discussed in the light of information from the literature. Their accuracy is evaluated on the basis of several pieces of anatomical evidence and clinical cases.

  6. Geophysical Characterization of Groundwater-Fault Dynamics at San Andreas Oasis

    NASA Astrophysics Data System (ADS)

    Faherty, D.; Polet, J.; Osborn, S. G.

    2017-12-01

    The San Andreas Oasis has historically provided a reliable source of fresh water near the northeast margin of the Salton Sea, although since the recent completion of the Coachella Canal Lining Project and persistent drought in California, surface water at the site has begun to disappear. This may be an effect of the canal lining, however, the controls on groundwater are complicated by the presence of the Hidden Springs Fault (HSF), a northeast dipping normal fault that trends near the San Andreas Oasis. Its surface expression is apparent as a lineation against which all plant growth terminates, suggesting that it may form a partial barrier to subsurface groundwater flow. Numerous environmental studies have detailed the chemical evolution of waters resources at San Andreas Spring, although there remains a knowledge gap on the HSF and its relation to groundwater at the site. To better constrain flow paths and characterize groundwater-fault interactions, we have employed resistivity surveys near the surface trace of the HSF to generate profiles of lateral and depth-dependent variations in resistivity. The survey design is comprised of lines installed in Wenner Arrays, using an IRIS Syscal Kid, with 24 electrodes, at a maximum electrode spacing of 5 meters. In addition, we have gathered constraints on the geometry of the HSF using a combination of ground-based magnetic and gravity profiles, conducted with a GEM walking Proton Precession magnetometer and a Lacoste & Romberg gravimeter. Seventeen gravity measurements were acquired across the surface trace of the fault. Preliminary resistivity results depict a shallow conductor localized at the oasis and discontinuous across the HSF. Magnetic data reveal a large contrast in subsurface magnetic susceptibility that appears coincident with the surface trace and trend of the HSF, while gravity data suggests a shallow, relatively high density anomaly centered near the oasis. These data also hint at a second, previously

  7. Deformation rates across the San Andreas Fault system, central California determined by geology and geodesy

    NASA Astrophysics Data System (ADS)

    Titus, Sarah J.

    The San Andreas fault system is a transpressional plate boundary characterized by sub-parallel dextral strike-slip faults separating internally deformed crustal blocks in central California. Both geodetic and geologic tools were used to understand the short- and long-term partitioning of deformation in both the crust and the lithospheric mantle across the plate boundary system. GPS data indicate that the short-term discrete deformation rate is ˜28 mm/yr for the central creeping segment of the San Andreas fault and increases to 33 mm/yr at +/-35 km from the fault. This gradient in deformation rates is interpreted to reflect elastic locking of the creeping segment at depth, distributed off-fault deformation, or some combination of these two mechanisms. These short-term fault-parallel deformation rates are slower than the expected geologic slip rate and the relative plate motion rate. Structural analysis of folds and transpressional kinematic modeling were used to quantify long-term distributed deformation adjacent to the Rinconada fault. Folding accommodates approximately 5 km of wrench deformation, which translates to a deformation rate of ˜1 mm/yr since the start of the Pliocene. Integration with discrete offset on the Rinconada fault indicates that this portion of the San Andreas fault system is approximately 80% strike-slip partitioned. This kinematic fold model can be applied to the entire San Andreas fault system and may explain some of the across-fault gradient in deformation rates recorded by the geodetic data. Petrologic examination of mantle xenoliths from the Coyote Lake basalt near the Calaveras fault was used to link crustal plate boundary deformation at the surface with models for the accommodation of deformation in the lithospheric mantle. Seismic anisotropy calculations based on xenolith petrofabrics suggest that an anisotropic mantle layer thickness of 35-85 km is required to explain the observed shear wave splitting delay times in central

  8. Correlation between deep fluids, tremor and creep along the central San Andreas fault

    USGS Publications Warehouse

    Becken, M.; Ritter, O.; Bedrosian, P.A.; Weckmann, U.

    2011-01-01

    The seismicity pattern along the San Andreas fault near Parkfield and Cholame, California, varies distinctly over a length of only fifty kilometres. Within the brittle crust, the presence of frictionally weak minerals, fault-weakening high fluid pressures and chemical weakening are considered possible causes of an anomalously weak fault northwest of Parkfield. Non-volcanic tremor from lower-crustal and upper-mantle depths is most pronounced about thirty kilometres southeast of Parkfield and is thought to be associated with high pore-fluid pressures at depth. Here we present geophysical evidence of fluids migrating into the creeping section of the San Andreas fault that seem to originate in the region of the uppermost mantle that also stimulates tremor, and evidence that along-strike variations in tremor activity and amplitude are related to strength variations in the lower crust and upper mantle. Interconnected fluids can explain a deep zone of anomalously low electrical resistivity that has been imaged by magnetotelluric data southwest of the Parkfield-Cholame segment. Near Cholame, where fluids seem to be trapped below a high-resistivity cap, tremor concentrates adjacent to the inferred fluids within a mechanically strong zone of high resistivity. By contrast, subvertical zones of low resistivity breach the entire crust near the drill hole of the San Andreas Fault Observatory at Depth, northwest of Parkfield, and imply pathways for deep fluids into the eastern fault block, coincident with a mechanically weak crust and the lower tremor amplitudes in the lower crust. Fluid influx to the fault system is consistent with hypotheses of fault-weakening high fluid pressures in the brittle crust.

  9. Photomosaics and logs of trenches on the San Andreas Fault, Thousand Palms Oasis, California

    USGS Publications Warehouse

    Fumal, Thomas E.; Frost, William T.; Garvin, Christopher; Hamilton, John C.; Jaasma, Monique; Rymer, Michael J.

    2004-01-01

    We present photomosaics and logs of the walls of trenches excavated for a paleoseismic study at Thousand Palms Oasis (Fig. 1). The site is located on the Mission Creek strand of the San Andreas fault zone, one of two major active strands of the fault in the Indio Hills along the northeast margin of the Coachella Valley (Fig. 2). The Coachella Valley section is the most poorly understood major part of the San Andreas fault with regard to slip rate and timing of past large-magnitude earthquakes, and therefore earthquake hazard. No large earthquakes have occurred for more than three centuries, the longest elapsed time for any part of the southern San Andreas fault. In spite of this, the Working Group on California Earthquake Probabilities (1995) assigned the lowest 30-year conditional probability on the southern San Andreas fault to the Coachella Valley. Models of the behavior of this part of the fault, however, have been based on very limited geologic data. The Thousand Palms Oasis is an attractive location for paleoseismic study primarily because of the well-bedded late Holocene sedimentary deposits with abundant layers of organic matter for radiocarbon dating necessary to constrain the timing of large prehistoric earthquakes. Previous attempts to develop a chronology of paleoearthquakes for the region have been hindered by the scarcity of in-situ 14C-dateable material for age control in this desert environment. Also, the fault in the vicinity of Thousand Palms Oasis consists of a single trace that is well expressed, both geomorphically and as a vegetation lineament (Figs. 2, 3). Results of our investigations are discussed in Fumal et al. (2002) and indicate that four and probably five surface-rupturing earthquakes occurred along this part of the fault during the past 1200 years. The average recurrence time for these earthquakes is 215 ± 25 years, although interevent times may have been as short as a few decades or as long as 400 years. Thus, although the elapsed

  10. Andreas Vesalius 500 years - A Renaissance that revolutionized cardiovascular knowledge

    PubMed Central

    Mesquita, Evandro Tinoco; de Souza Júnior, Celso Vale; Ferreira, Thiago Reigado

    2015-01-01

    The history of medicine and cardiology is marked by some geniuses who dared in thinking, research, teaching and transmitting scientific knowledge, and the Italian Andreas Vesalius one of these brilliant masters. His main scientific work "De Humani Corporis Fabrica" is not only a landmark study of human anatomy but also an artistic work of high aesthetic quality published in 1543. In the year 2014 we celebrated 500 years since the birth of the brilliant professor of Padua University, who with his courage and sense of observation changed the understanding of cardiovascular anatomy and founded a school to date in innovative education and research of anatomy. By identifying "the anatomical errors" present in Galen's book and speech, he challenged the dogmas of the Catholic Church, the academic world and the doctors of his time. However, the accuracy of his findings and his innovative way to disseminate them among his students and colleagues was essential so that his contributions are considered by many the landmark of modern medicine. His death is still surrounded by mysteries having different hypotheses, but a certainty, suffered sanctions of the Catholic Church for the spread of their ideas. The cardiologists, cardiovascular surgeons, interventional cardiologists, electrophysiologists and cardiovascular imaginologists must know the legacy of genius Andreas Vesalius that changed the paradigm of human anatomy. PMID:26107459

  11. Andreas Vesalius 500 years--A Renaissance that revolutionized cardiovascular knowledge.

    PubMed

    Mesquita, Evandro Tinoco; Souza Júnior, Celso Vale de; Ferreira, Thiago Reigado

    2015-01-01

    The history of medicine and cardiology is marked by some geniuses who dared in thinking, research, teaching and transmitting scientific knowledge, and the Italian Andreas Vesalius one of these brilliant masters. His main scientific work "De Humani Corporis Fabrica" is not only a landmark study of human anatomy but also an artistic work of high aesthetic quality published in 1543. In the year 2014 we celebrated 500 years since the birth of the brilliant professor of Padua University, who with his courage and sense of observation changed the understanding of cardiovascular anatomy and founded a school to date in innovative education and research of anatomy. By identifying "the anatomical errors" present in Galen's book and speech, he challenged the dogmas of the Catholic Church, the academic world and the doctors of his time. However, the accuracy of his findings and his innovative way to disseminate them among his students and colleagues was essential so that his contributions are considered by many the landmark of modern medicine. His death is still surrounded by mysteries having different hypotheses, but a certainty, suffered sanctions of the Catholic Church for the spread of their ideas. The cardiologists, cardiovascular surgeons, interventional cardiologists, electrophysiologists and cardiovascular imaginologists must know the legacy of genius Andreas Vesalius that changed the paradigm of human anatomy.

  12. Cataclastic rocks of the San Gabriel fault—an expression of deformation at deeper crustal levels in the San Andreas fault zone

    NASA Astrophysics Data System (ADS)

    Anderson, J. Lawford; Osborne, Robert H.; Palmer, Donald F.

    1983-10-01

    The San Gabriel fault, a deeply eroded late Oligocene to middle Pliocene precursor to the San Andreas, was chosen for petrologic study to provide information regarding intrafault material representative of deeper crustal levels. Cataclastic rocks exposed along the present trace of the San Andreas in this area are exclusively a variety of fault gouge that is essentially a rock flour with a quartz, feldspar, biotite, chlorite, amphibole, epidote, and Fe-Ti oxide mineralogy representing the milled-down equivalent of the original rock (Anderson and Osborne, 1979; Anderson et al., 1980). Likewise, fault gouge and associated breccia are common along the San Gabriel fault, but only where the zone of cataclasis is several tens of meters wide. At several localities, the zone is extremely narrow (several centimeters), and the cataclastic rock type is cataclasite, a dark, aphanitic, and highly comminuted and indurated rock. The cataclastic rocks along the San Gabriel fault exhibit more comminution than that observed for gouge along the San Andreas. The average grain diameter for the San Andreas gouge ranges from 0.01 to 0.06 mm. For the San Gabriel cataclastic rocks, it ranges from 0.0001 to 0.007 mm. Whereas the San Andreas gouge remains particulate to the smallest grain-size, the ultra-fine grain matrix of the San Gabriel cataclasite is composed of a mosaic of equidimensional, interlocking grains. The cataclastic rocks along the San Gabriel fault also show more mineralogiec changes compared to gouge from the San Andreas fault. At the expense of biotite, amphibole, and feldspar, there is some growth of new albite, chlorite, sericite, laumontite, analcime, mordenite (?), and calcite. The highest grade of metamorphism is laumontite-chlorite zone (zeolite facies). Mineral assemblages and constrained uplift rates allow temperature and depth estimates of 200 ± 30° C and 2-5 km, thus suggesting an approximate geothermal gradient of ~50°C/km. Such elevated temperatures imply a

  13. The Eastern California Shear Zone as the northward extension of the southern San Andreas Fault

    USGS Publications Warehouse

    Thatcher, Wayne R.; Savage, James C.; Simpson, Robert W.

    2016-01-01

    Cluster analysis offers an agnostic way to organize and explore features of the current GPS velocity field without reference to geologic information or physical models using information only contained in the velocity field itself. We have used cluster analysis of the Southern California Global Positioning System (GPS) velocity field to determine the partitioning of Pacific-North America relative motion onto major regional faults. Our results indicate the large-scale kinematics of the region is best described with two boundaries of high velocity gradient, one centered on the Coachella section of the San Andreas Fault and the Eastern California Shear Zone and the other defined by the San Jacinto Fault south of Cajon Pass and the San Andreas Fault farther north. The ~120 km long strand of the San Andreas between Cajon Pass and Coachella Valley (often termed the San Bernardino and San Gorgonio sections) is thus currently of secondary importance and carries lesser amounts of slip over most or all of its length. We show these first order results are present in maps of the smoothed GPS velocity field itself. They are also generally consistent with currently available, loosely bounded geologic and geodetic fault slip rate estimates that alone do not provide useful constraints on the large-scale partitioning we show here. Our analysis does not preclude the existence of smaller blocks and more block boundaries in Southern California. However, attempts to identify smaller blocks along and adjacent to the San Gorgonio section were not successful.

  14. The Eastern California Shear Zone as the northward extension of the southern San Andreas Fault

    NASA Astrophysics Data System (ADS)

    Thatcher, W.; Savage, J. C.; Simpson, R. W.

    2016-04-01

    Cluster analysis offers an agnostic way to organize and explore features of the current GPS velocity field without reference to geologic information or physical models using information only contained in the velocity field itself. We have used cluster analysis of the Southern California Global Positioning System (GPS) velocity field to determine the partitioning of Pacific-North America relative motion onto major regional faults. Our results indicate the large-scale kinematics of the region is best described with two boundaries of high velocity gradient, one centered on the Coachella section of the San Andreas Fault and the Eastern California Shear Zone and the other defined by the San Jacinto Fault south of Cajon Pass and the San Andreas Fault farther north. The ~120 km long strand of the San Andreas between Cajon Pass and Coachella Valley (often termed the San Bernardino and San Gorgonio sections) is thus currently of secondary importance and carries lesser amounts of slip over most or all of its length. We show these first order results are present in maps of the smoothed GPS velocity field itself. They are also generally consistent with currently available, loosely bounded geologic and geodetic fault slip rate estimates that alone do not provide useful constraints on the large-scale partitioning we show here. Our analysis does not preclude the existence of smaller blocks and more block boundaries in Southern California. However, attempts to identify smaller blocks along and adjacent to the San Gorgonio section were not successful.

  15. The wister mud pot lineament: Southeastward extension or abandoned strand of the San Andreas fault?

    USGS Publications Warehouse

    Lynch, D.K.; Hudnut, K.W.

    2008-01-01

    We present the results of a survey of mud pots in the Wister Unit of the Imperial Wildlife Area. Thirty-three mud pots, pot clusters, or related geothermal vents (hundreds of pots in all) were identified, and most were found to cluster along a northwest-trending line that is more or less coincident with the postulated Sand Hills fault. An extrapolation of the trace of the San Andreas fault southeastward from its accepted terminus north of Bombay Beach very nearly coincides with the mud pot lineament and may represent a surface manifestation of the San Andreas fault southeast of the Salton Sea. Additionally, a recent survey of vents near Mullet Island in the Salton Sea revealed eight areas along a northwest-striking line where gas was bubbling up through the water and in two cases hot mud and water were being violently ejected.

  16. Chicks in Charge: Andrea Baker & Amy Daniels--Airport High School Media Center, Columbia, SC

    ERIC Educational Resources Information Center

    Library Journal, 2004

    2004-01-01

    This article briefly discusses two librarians exploration of Linux. Andrea Baker and Amy Daniels were tired of telling their students that new technology items were not in the budget. They explored Linux, which is a program that recycles older computers, installs free operating systems and free software.

  17. ESA Astronaut Andreas Mogensen and NASA astronaut Randy Bresnik during NEEMO 19 communications training with instructors

    NASA Image and Video Library

    2014-08-21

    Date: 08-21-14 Location: Bldg 36, 131 Subject: ESA Astronaut Andreas Mogensen and NASA astronaut Randy Bresnik during NEEMO 19 communications training with instructors Marcum Reagan and Barbara Janoiko Photographer: James Blair

  18. Quasi-periodic recurrence of large earthquakes on the southern San Andreas fault

    USGS Publications Warehouse

    Scharer, Katherine M.; Biasi, Glenn P.; Weldon, Ray J.; Fumal, Tom E.

    2010-01-01

    It has been 153 yr since the last large earthquake on the southern San Andreas fault (California, United States), but the average interseismic interval is only ~100 yr. If the recurrence of large earthquakes is periodic, rather than random or clustered, the length of this period is notable and would generally increase the risk estimated in probabilistic seismic hazard analyses. Unfortunately, robust characterization of a distribution describing earthquake recurrence on a single fault is limited by the brevity of most earthquake records. Here we use statistical tests on a 3000 yr combined record of 29 ground-rupturing earthquakes from Wrightwood, California. We show that earthquake recurrence there is more regular than expected from a Poisson distribution and is not clustered, leading us to conclude that recurrence is quasi-periodic. The observation of unimodal time dependence is persistent across an observationally based sensitivity analysis that critically examines alternative interpretations of the geologic record. The results support formal forecast efforts that use renewal models to estimate probabilities of future earthquakes on the southern San Andreas fault. Only four intervals (15%) from the record are longer than the present open interval, highlighting the current hazard posed by this fault.

  19. Structure of the 1906 near-surface rupture zone of the San Andreas Fault, San Francisco Peninsula segment, near Woodside, California

    USGS Publications Warehouse

    Rosa, C.M.; Catchings, R.D.; Rymer, M.J.; Grove, Karen; Goldman, M.R.

    2016-07-08

    High-resolution seismic-reflection and refraction images of the 1906 surface rupture zone of the San Andreas Fault near Woodside, California reveal evidence for one or more additional near-surface (within about 3 meters [m] depth) fault strands within about 25 m of the 1906 surface rupture. The 1906 surface rupture above the groundwater table (vadose zone) has been observed in paleoseismic trenches that coincide with our seismic profile and is seismically characterized by a discrete zone of low P-wave velocities (Vp), low S-wave velocities (Vs), high Vp/Vs ratios, and high Poisson’s ratios. A second near-surface fault strand, located about 17 m to the southwest of the 1906 surface rupture, is inferred by similar seismic anomalies. Between these two near-surface fault strands and below 5 m depth, we observed a near-vertical fault strand characterized by a zone of high Vp, low Vs, high Vp/Vs ratios, and high Poisson’s ratios on refraction tomography images and near-vertical diffractions on seismic-reflection images. This prominent subsurface zone of seismic anomalies is laterally offset from the 1906 surface rupture by about 8 m and likely represents the active main (long-term) strand of the San Andreas Fault at 5 to 10 m depth. Geometries of the near-surface and subsurface (about 5 to 10 m depth) fault zone suggest that the 1906 surface rupture dips southwestward to join the main strand of the San Andreas Fault at about 5 to 10 m below the surface. The 1906 surface rupture forms a prominent groundwater barrier in the upper 3 to 5 m, but our interpreted secondary near-surface fault strand to the southwest forms a weaker barrier, suggesting that there has been less or less-recent near-surface slip on that strand. At about 6 m depth, the main strand of the San Andreas Fault consists of water-saturated blue clay (collected from a hand-augered borehole), which is similar to deeply weathered serpentinite observed within the main strand of the San Andreas Fault at

  20. Response of deformation patterns to reorganizations of the southern San Andreas fault system since ca. 1.5 Ma

    NASA Astrophysics Data System (ADS)

    Cooke, M. L.; Fattaruso, L.; Dorsey, R. J.; Housen, B. A.

    2015-12-01

    Between ~1.5 and 1.1 Ma, the southern San Andreas fault system underwent a major reorganization that included initiation of the San Jacinto fault and termination of slip on the extensional West Salton detachment fault. The southern San Andreas fault itself has also evolved since this time, with several shifts in activity among fault strands within San Gorgonio Pass. We use three-dimensional mechanical Boundary Element Method models to investigate the impact of these changes to the fault network on deformation patterns. A series of snapshot models of the succession of active fault geometries explore the role of fault interaction and tectonic loading in abandonment of the West Salton detachment fault, initiation of the San Jacinto fault, and shifts in activity of the San Andreas fault. Interpreted changes to uplift patterns are well matched by model results. These results support the idea that growth of the San Jacinto fault led to increased uplift rates in the San Gabriel Mountains and decreased uplift rates in the San Bernardino Mountains. Comparison of model results for vertical axis rotation to data from paleomagnetic studies reveals a good match to local rotation patterns in the Mecca Hills and Borrego Badlands. We explore the mechanical efficiency at each step in the evolution, and find an overall trend toward increased efficiency through time. Strain energy density patterns are used to identify regions of off-fault deformation and potential incipient faulting. These patterns support the notion of north-to-south propagation of the San Jacinto fault during its initiation. The results of the present-day model are compared with microseismicity focal mechanisms to provide additional insight into the patterns of off-fault deformation within the southern San Andreas fault system.

  1. San Andreas fault geometry at Desert Hot Springs, California, and its effects on earthquake hazards and groundwater

    USGS Publications Warehouse

    Catchings, R.D.; Rymer, M.J.; Goldman, M.R.; Gandhok, G.

    2009-01-01

    The Mission Creek and Banning faults are two of the principal strands of the San Andreas fault zone in the northern Coachella Valley of southern California. Structural characteristics of the faults affect both regional earthquake hazards and local groundwater resources. We use seismic, gravity, and geological data to characterize the San Andreas fault zone in the vicinity of Desert Hot Springs. Seismic images of the upper 500 m of the Mission Creek fault at Desert Hot Springs show multiple fault strands distributed over a 500 m wide zone, with concentrated faulting within a central 200 m wide area of the fault zone. High-velocity (up to 5000 m=sec) rocks on the northeast side of the fault are juxtaposed against a low-velocity (6.0) earthquakes in the area (in 1948 and 1986) occurred at or near the depths (~10 to 12 km) of the merged (San Andreas) fault. Large-magnitude earthquakes that nucleate at or below the merged fault will likely generate strong shaking from guided waves along both fault zones and from amplified seismic waves in the low-velocity basin between the two fault zones. The Mission Creek fault zone is a groundwater barrier with the top of the water table varying by 60 m in depth and the aquifer varying by about 50 m in thickness across a 200 m wide zone of concentrated faulting.

  2. Miocene climate as recorded on slope carbonates : examples from Malta (Central Mediterranean) and Northeastern Australia (Marion Plateau, ODP LEG 194)

    NASA Astrophysics Data System (ADS)

    John, Cédric Michaël

    2003-08-01

    This study investigated the slope carbonates of two Miocene carbonate systems: the Maltese Islands (in the Central Mediterranean) and the Marion Plateau (Northeastern Australia, drilled during ODP Leg 194). The aim of the study was to trace the impact of the Miocene cooling steps (events Mi1-Mi6) in these carbonate systems, especially the Mi3 event, which took place around 13.6 Ma and deeply impacted the marine oxygen isotope record. This event also profoundly impacted oceanographic and climatic patterns, eventually leading to the establishment of the modern ice-house world. In particular, East Antarctica became ice covered at that period. The rational behind the present study was to investigate the impact that this event had on shallow water systems in order to complement the deep-sea record and hence acquire a more global perspective on Miocene climate change. The Maltese Islands were investigated for trends in bulk-rock carbon and oxygen isotopes, as well as bulk-rock mineralogy, clay minerals analysis and organic geochemisty. Results showed that the mid Miocene cooling event deeply impacted sedimentation at that location by changing sedimentation from carbonate to clay-rich sediments. Moreover, it was discovered that each phase of Antarctic glaciation, not just the major mid Miocene event, resulted in higher terrigenous input on Malta. Mass accumulation rates revealed that this was linked to increased runoff during periods when Antarctica was glaciated, and thus that the carbonate sediments were “diluted” by clay-rich sediments. The model subsequently developed to explain this implies feedback from Antarctic glaciations creating cold, dense air masses that push the ITCZ Northward, thus increasing precipitation on the North African subcontinent. Increased precipitation (or stronger African monsoon) accelerated continental weathering and runoff, thus bringing more terrigenous sediment to the paleo-location of the slope sediments of Malta. Spectral

  3. Timing of large earthquakes since A.D. 800 on the Mission Creek strand of the San Andreas fault zone at Thousand Palms Oasis, near Palm Springs, California

    USGS Publications Warehouse

    Fumal, T.E.; Rymer, M.J.; Seitz, G.G.

    2002-01-01

    Paleoseismic investigations across the Mission Creek strand of the San Andreas fault at Thousand Palms Oasis indicate that four and probably five surface-rupturing earthquakes occurred during the past 1200 years. Calendar age estimates for these earthquakes are based on a chronological model that incorporates radio-carbon dates from 18 in situ burn layers and stratigraphic ordering constraints. These five earthquakes occurred in about A.D. 825 (770-890) (mean, 95% range), A.D. 982 (840-1150), A.D. 1231 (1170-1290), A.D. 1502 (1450-1555), and after a date in the range of A.D. 1520-1680. The most recent surface-rupturing earthquake at Thousand Palms is likely the same as the A.D. 1676 ?? 35 event at Indio reported by Sieh and Williams (1990). Each of the past five earthquakes recorded on the San Andreas fault in the Coachella Valley strongly overlaps in time with an event at the Wrightwood paleoseismic site, about 120 km northwest of Thousand Palms Oasis. Correlation of events between these two sites suggests that at least the southernmost 200 km of the San Andreas fault zone may have ruptured in each earthquake. The average repeat time for surface-rupturing earthquakes on the San Andreas fault in the Coachella Valley is 215 ?? 25 years, whereas the elapsed time since the most recent event is 326 ?? 35 years. This suggests the southernmost San Andreas fault zone likely is very near failure. The Thousand Palms Oasis site is underlain by a series of six channels cut and filled since about A.D. 800 that cross the fault at high angles. A channel margin about 900 years old is offset right laterally 2.0 ?? 0.5 m, indicating a slip rate of 4 ?? 2 mm/yr. This slip rate is low relative to geodetic and other geologic slip rate estimates (26 ?? 2 mm/yr and about 23-35 mm/yr, respectively) on the southernmost San Andreas fault zone, possibly because (1) the site is located in a small step-over in the fault trace and so the rate is not be representative of the Mission Creek fault

  4. Availability of water from the Outwash Aquifer, Marion County, Indiana

    USGS Publications Warehouse

    Smith, B.S.

    1983-01-01

    The outwash aquifer in Marion County, Indiana is a continuous, unconfined sand and gravel deposit containing isolated boulder, till, silt, and clay deposits along the White River, Fall Creek, and Eagle Creek. Flow in the aquifer is from the boundaries of the aquifer with the Tipton till plain toward the streams and major pumping centers in the aquifer. A two-dimensional, finite-difference model of the outwash aquifer was calibrated to water levels of October 6 to 10, 1980 and used to estimate availability of water in the aquifer. A drawdown limit of 50-percent saturated thickness applied to 78 simulated-pumping wells assumed to be 1 foot in diameter produced 97 cubic feet per second from the outwash aquifer. Streamflow reductions caused by 97 cubic feet per second simulated pumpage and constant-flux boundaries were estimated to be 85 cubic feet per second in the White River and 12 cubic feet per second in Fall Creek. In comparison, the 7-day, 10-year low flows were 83 cubic feet per second in the White River near Nora and 23 cubic feet per second in Fall Creek at Millersville. Simulated pumpage of 115 cubic feet per second and constant-flux boundaries produced streamflow reductions of 101 cubic feet per second on the White River and 13 cubic feet per second on Fall Creek. (USGS)

  5. The stress shadow effect: a mechanical analysis of the evenly-spaced parallel strike-slip faults in the San Andreas fault system

    NASA Astrophysics Data System (ADS)

    Zuza, A. V.; Yin, A.; Lin, J. C.

    2015-12-01

    Parallel evenly-spaced strike-slip faults are prominent in the southern San Andreas fault system, as well as other settings along plate boundaries (e.g., the Alpine fault) and within continental interiors (e.g., the North Anatolian, central Asian, and northern Tibetan faults). In southern California, the parallel San Jacinto, Elsinore, Rose Canyon, and San Clemente faults to the west of the San Andreas are regularly spaced at ~40 km. In the Eastern California Shear Zone, east of the San Andreas, faults are spaced at ~15 km. These characteristic spacings provide unique mechanical constraints on how the faults interact. Despite the common occurrence of parallel strike-slip faults, the fundamental questions of how and why these fault systems form remain unanswered. We address this issue by using the stress shadow concept of Lachenbruch (1961)—developed to explain extensional joints by using the stress-free condition on the crack surface—to present a mechanical analysis of the formation of parallel strike-slip faults that relates fault spacing and brittle-crust thickness to fault strength, crustal strength, and the crustal stress state. We discuss three independent models: (1) a fracture mechanics model, (2) an empirical stress-rise function model embedded in a plastic medium, and (3) an elastic-plate model. The assumptions and predictions of these models are quantitatively tested using scaled analogue sandbox experiments that show that strike-slip fault spacing is linearly related to the brittle-crust thickness. We derive constraints on the mechanical properties of the southern San Andreas strike-slip faults and fault-bounded crust (e.g., local fault strength and crustal/regional stress) given the observed fault spacing and brittle-crust thickness, which is obtained by defining the base of the seismogenic zone with high-resolution earthquake data. Our models allow direct comparison of the parallel faults in the southern San Andreas system with other similar strike

  6. Geomorphology, denudation rates, and stream channel profiles reveal patterns of mountain building adjacent to the San Andreas fault in northern California, USA

    USGS Publications Warehouse

    DeLong, Stephen B.; Hilley, George E.; Prentice, Carol S.; Crosby, Christopher J.; Yokelson, Intan N.

    2017-01-01

    Relative horizontal motion along strike-slip faults can build mountains when motion is oblique to the trend of the strike-slip boundary. The resulting contraction and uplift pose off-fault seismic hazards, which are often difficult to detect because of the poor vertical resolution of satellite geodesy and difficulty of locating offset datable landforms in active mountain ranges. Sparse geomorphic markers, topographic analyses, and measurement of denudation allow us to map spatiotemporal patterns of uplift along the northern San Andreas fault. Between Jenner and Mendocino, California, emergent marine terraces found southwest of the San Andreas fault record late Pleistocene uplift rates between 0.20 and 0.45 mm yr–1 along much of the coast. However, on the northeast side of the San Andreas fault, a zone of rapid uplift (0.6–1.0 mm yr–1) exists adjacent to the San Andreas fault, but rates decay northeastward as the coast becomes more distant from the San Andreas fault. A newly dated 4.5 Ma shallow-marine deposit located at ∼500 m above sea level (masl) adjacent to the San Andreas fault is warped down to just 150 masl 15 km northeast of the San Andreas fault, and it is exposed at just 60–110 masl to the west of the fault. Landscape denudation rates calculated from abundance of cosmogenic radionuclides in fluvial sediment northeast of, and adjacent to, the San Andreas fault are 0.16–0.29 mm yr–1, but they are only 0.03–0.07 mm yr–1 west of the fault. Basin-average channel steepness and the denudation rates can be used to infer the erosive properties of the underlying bedrock. Calibrated erosion rates can then be estimated across the entire landscape using the spatial distribution of channel steepness with these erosive properties. The lower-elevation areas of this landscape that show high channel steepness (and hence calibrated erosion rate) are distinct from higher-elevation areas with systematically lower channel steepness and denudation rates

  7. Perspective View, San Andreas Fault

    NASA Technical Reports Server (NTRS)

    2000-01-01

    The prominent linear feature straight down the center of this perspective view is California's famous San Andreas Fault. The image, created with data from NASA's Shuttle Radar Topography Mission (SRTM), will be used by geologists studying fault dynamics and landforms resulting from active tectonics. This segment of the fault lies west of the city of Palmdale, Calif., about 100 kilometers (about 60 miles) northwest of Los Angeles. The fault is the active tectonic boundary between the North American plate on the right, and the Pacific plate on the left. Relative to each other, the Pacific plate is moving away from the viewer and the North American plate is moving toward the viewer along what geologists call a right lateral strike-slip fault. Two large mountain ranges are visible, the San Gabriel Mountains on the left and the Tehachapi Mountains in the upper right. Another fault, the Garlock Fault lies at the base of the Tehachapis; the San Andreas and the Garlock Faults meet in the center distance near the town of Gorman. In the distance, over the Tehachapi Mountains is California's Central Valley. Along the foothills in the right hand part of the image is the Antelope Valley, including the Antelope Valley California Poppy Reserve. The data used to create this image were acquired by SRTM aboard the Space Shuttle Endeavour, launched on February 11, 2000.

    This type of display adds the important dimension of elevation to the study of land use and environmental processes as observed in satellite images. The perspective view was created by draping a Landsat satellite image over an SRTM elevation model. Topography is exaggerated 1.5 times vertically. The Landsat image was provided by the United States Geological Survey's Earth Resources Observations Systems (EROS) Data Center, Sioux Falls, South Dakota.

    SRTM uses the same radar instrument that comprised the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) that flew twice on the Space

  8. Paleoseismic investigations in the Santa Cruz mountains, California: Implications for recurrence of large-magnitude earthquakes on the San Andreas fault

    USGS Publications Warehouse

    Schwartz, D.P.; Pantosti, D.; Okumura, K.; Powers, T.J.; Hamilton, J.C.

    1998-01-01

    Trenching, microgeomorphic mapping, and tree ring analysis provide information on timing of paleoearthquakes and behavior of the San Andreas fault in the Santa Cruz mountains. At the Grizzly Flat site alluvial units dated at 1640-1659 A.D., 1679-1894 A.D., 1668-1893 A.D., and the present ground surface are displaced by a single event. This was the 1906 surface rupture. Combined trench dates and tree ring analysis suggest that the penultimate event occurred in the mid-1600s, possibly in an interval as narrow as 1632-1659 A.D. There is no direct evidence in the trenches for the 1838 or 1865 earthquakes, which have been proposed as occurring on this part of the fault zone. In a minimum time of about 340 years only one large surface faulting event (1906) occurred at Grizzly Flat, in contrast to previous recurrence estimates of 95-110 years for the Santa Cruz mountains segment. Comparison with dates of the penultimate San Andreas earthquake at sites north of San Francisco suggests that the San Andreas fault between Point Arena and the Santa Cruz mountains may have failed either as a sequence of closely timed earthquakes on adjacent segments or as a single long rupture similar in length to the 1906 rupture around the mid-1600s. The 1906 coseismic geodetic slip and the late Holocene geologic slip rate on the San Francisco peninsula and southward are about 50-70% and 70% of their values north of San Francisco, respectively. The slip gradient along the 1906 rupture section of the San Andreas reflects partitioning of plate boundary slip onto the San Gregorio, Sargent, and other faults south of the Golden Gate. If a mid-1600s event ruptured the same section of the fault that failed in 1906, it supports the concept that long strike-slip faults can contain master rupture segments that repeat in both length and slip distribution. Recognition of a persistent slip rate gradient along the northern San Andreas fault and the concept of a master segment remove the requirement that

  9. Andreas Vesalius (1514-1564) - an unfinished life.

    PubMed

    Ambrose, Charles T

    2014-01-01

    The fame of Andreas Vesalius (1514-1564) rests on his anatomy text, De humani corporis fabrica, regarded as a seminal book in modern medicine. It was compiled while he taught anatomy at Padua, 1537-1543. Some of his findings challenged Galen's writings of the 2c AD, and caused De fabrica to be rejected immediately by classically trained anatomists. At age 29, Vesalius abandoned his studies and over the next two decades served as physician to Emperor Charles V of the Holy Roman Empire (HRE) and later to King Philip II of Spain in Madrid. In 1564, he sought to resume teaching anatomy in Padua, but release from royal service obliged him first to make a pilgrimage to Palestine. During the return voyage to Venice, he became ill and was put ashore alone on an Ionian island Zakynthos, where he died days later at age 50.

  10. Did J. Marion Sims deliberately addict his first fistula patients to opium?

    PubMed

    Wall, L Lewis

    2007-07-01

    American surgeon J. Marion Sims (1813-83) is regarded by many modern authors as a controversial figure because he carried out a series of experimental surgeries on enslaved African American women between 1846 and 1849 in an attempt to cure them of vesicovaginal fistulas, which they had all developed as a result of prolonged obstructed labor. He operated on one woman, Anarcha Westcott, thirty times before he successfully closed her fistula. Sims performed these fistula repair operations without benefit of anesthesia but gave these women substantial doses of opium afterwards. Several modern writers have alleged that Sims did this in order to addict them to the drug and thereby to enhance his control over them. This article examines the controversy surrounding Sims' use of postoperative opium in these enslaved surgical patients. The evidence suggests that although these women were probably tolerant to the doses of opium that he used, there is no evidence that he deliberately tried to addict them to this drug. Sims' use of postoperative opium appears to have been well supported by the therapeutic practices of his day, and the regimen that he used was enthusiastically supported by many contemporary surgeons.

  11. Major refit of R/V MARION DUFRESNE and giant sediment corer improvements

    NASA Astrophysics Data System (ADS)

    Leau, Hélène; Réaud, Yvan

    2015-04-01

    The french Research Vessel MARION DUFRESNE is equipped with a unique sediment coring facility, called CALYPSO, developed initially by Yvon BALUT at the French Polar Institute, Paul-Emile Victor (IPEV) that operates the vessel 217 days per year in all oceans. The CALYPSO sediment corer retrieves routinely 50 m long undisturbed sediment cores in any water depths, and presently holds the worldwide record of the longest core ever retrieved, that is 64.5 m. This vessel is then a fantastic opportunity for the paleoceanographic community to carry out expeditions at sea. Over the last 20 years, many international IMAGES coring expeditions were organized in all the ocean basins around the world on board the R/V MARION DUFRESNE. More than 1500 cores were retrieved, leading to major advances in the paleoceanography and paleoclimatology of the Late Quaternary. The vessel will celebrate her 20th anniversary in 2015 and will undergo a major refit on hull & machineries, public spaces, as well as scientific equipment. The coring capacity is currently being developed to further improve - The length of the retrievable core, with an objective of 75 m long core in routine - The quality of the sediment un-disturbance with a specially designed coring cable with controlled minimum elasticity - The safety of the operations at sea - The quality control of the operations with a suite of sensors and software allowing a detailed monitoring of the coring operation - The time requested for each operation - The environment data collection, in the same time as the coring operations The detailed description of the upgrades will be presented. They consist in a new suite of acoustic sensors that will be integrated on board the vessel during the 4 months ship yard stay from April to July 2015, amongst which a KONSBERG EM122 multibeam echo-sounder and a SBP 120-3 sub-bottom profiler, both mounted on a gondola fitted under the hull of the vessel. This equipment will allow the highest quality images of

  12. Peter Andreas Hansen and the astronomical community - a first investigation of the Hansen papers. (German Title: Peter Andreas Hansen und die astronomische Gemeinschaft - eine erste Auswertung des Hansen-Nachlasses. )

    NASA Astrophysics Data System (ADS)

    Schwarz, Oliver; Strumpf, Manfred

    The literary assets of Peter Andreas Hansen are deposited in the Staatsarchiv Hamburg, the Forschungs- und Landesbibliothek Gotha and the Thüringer Staatsarchiv Gotha. They were never systematically investigated. We present here some results of a first evaluation. It was possible to reconstruct the historical events with regard to the maintenance of the Astronomische Nachrichten and the Altona observatory in 1854. Hansen was a successful teacher for many young astronomers. His way of stimulating the evolution of astronomy followed Zach's tradition.

  13. San Andreas fault zone drilling project: scientific objectives and technological challenges

    USGS Publications Warehouse

    Hickman, Stephen; Younker, Leland; Zobeck, Mark; Cooper, George; ,

    1994-01-01

    We are leading a new international initiative to conduct scientific drilling within the San Andreas fault zone at depths of up to 10 km. This project is motivated by the need to understand the physical and chemical processes operating within the fault zone and to answer fundamental questions about earthquake generation along major plate-boundary faults. Through an integrated program of coring, fluid sampling, in-situ and laboratory experimentation and long-term monitoring, we hope to provide fundamental constraints on the structure, composition, mechanical behavior and physical state of the San Andreas fault system at depths comparable to the nucleation zones of great earthquakes. The drilling, sampling and observational requirements needed to ensure the success of this project are stringent. These include: 1) drilling stable vertical holes to depths of about 9 km in fractured rock at temperatures of up to 300??C; 2) continuous coring of inclined holes branched off these vertical boreholes to intersect the fault at depths of 3, 6 and 9 km; 3) conducting sophisticated borehole geophysical measurements and fluid/rock sampling at high temperatures and pressures; and 4) instrumenting some or all of these inclined core holes for continuous monitoring of seismicity and a broad range of physical and chemical properties over periods of up to several decades. For all of these tasks, because of the overpressured clay-rich formations anticipated within the fault zone at depth, we expect to encounter difficult drilling, coring and hole-completion conditions in the regions of greatest scientific interest.

  14. San Andreas fault zone drilling project: scientific objectives and technological challenges

    USGS Publications Warehouse

    Hickman, S.H.; Younker, L.W.; Zoback, M.D.

    1995-01-01

    We are leading a new international initiative to conduct scientific drilling within the San Andreas fault zone at depths of up to 10 km. This project is motivated by the need to understand the physical and chemical processes operating within the fault zone and to answer fundamental questions about earthquake generation along major plate-boundary faults. Through a comprehensive program of coring, fluid sampling, downhole measurements, laboratory experimentation, and long-term monitoring, we hope to obtain critical information on the structure, composition, mechanical behavior and physical state of the San Andreas fault system at depths comparable to the nucleation zones of great earthquakes. The drilling, sampling and observational requirements needed to ensure the success of this project are stringent. These include: 1) drilling stable vertical holes to depths of about 9 km in fractured rock at temperatures of up to 300°C; 2) continuous coring and completion of inclined holes branched off these vertical boreholes to intersect the fault at depths of 3, 6, and 9 km; 3) conducting sophisticated borehole geophysical measurements and fluid/rock sampling at high temperatures and pressures; and 4) instrumenting some or all of these inclined core holes for continuous monitoring of earthquake activity, fluid pressure, deformation and other parameters for periods of up to several decades. For all of these tasks, because of the overpressured clay-rich formations anticipated within the fault zone at depth, we expect to encounter difficult drilling, coring and hole-completion conditions in the region of greatest scientific interest.

  15. Dating offset fans along the Mojave section of the San Andreas fault using cosmogenic 26Al and 10Be

    USGS Publications Warehouse

    Matmon, A.; Schwartz, D.P.; Finkel, R.; Clemmens, S.; Hanks, T.

    2005-01-01

    Analysis of cosmogenic 10Be and 26Al in samples collected from exposed boulders (n = 20) and from buried sediment (n = 3) from offset fans along the San Andreas fault near Little Rock, California, yielded ages, ranging from 16 to 413 ka, which increase with distance from their source at the mouth of Little Rock Creek. In order to determine the age of the relatively younger fans, the erosion rate of the boulders and the cosmogenic nuclide inheritance from exposure prior to deposition in the fan were established. Cosmogenic nuclide inheritance values that range between 8.5 ?? 103 and 196 ?? 103 atoms 10Be g-1 quartz were determined by measuring the concentrations and ratios of 10Be and 26Al in boulders (n = 10) and fine sediment (n = 7) at the outlet of the present active stream. Boulder erosion rate, ranging between 17 and 160 mm k.y.-1, was estimated by measuring 10Be and 26Al concentrations in nearby bedrock outcrops (n = 8). Since the boulders on the fans represent the most resistant rocks in this environment, we used the lowest rate for the age calculations. Monte Carlo simulations were used to determine ages of 16 ?? 5 and 29 ?? 7 ka for the two younger fan surfaces. Older fans (older than 100 ka) were dated by analyzing 10Be and 26Al concentrations in buried sand samples. The ages of the three oldest fans range between 227 ?? 242 and 413 ?? 185 ka. Although fan age determinations are accompanied by large uncertainties, the results of this study show a clear trend of increasing fan ages with increasing distance from the source near Little Rock Creek and provide a long-term slip rate along this section of the San Andreas fault. Slip rate along the Mojave section of the San Andreas fault for the past 413 k.y. can be determined in several ways. The average slip rate calculated from the individual fan ages is 4.2 ?? 0.9 cm yr-1. A linear regression through the data points implies a slip rate of 3.7 ?? 1.0 cm yr-1. A most probable slip rate of 3.0 ?? 1.0 cm yr-1 is

  16. The outlook of physician histories: J. Marion Sims and 'The Discovery of Anaesthesia'.

    PubMed

    Rosenbloom, Julia M; Schonberger, Robert B

    2015-12-01

    The fact that doctors have a long tradition of writing medical history to interpret and direct their profession is well established. But readers (particularly modern physician readers) can also understand physician-authored histories as offering commentary and analysis of the world beyond medicine. In this essay, we offer a reading (perhaps a modern one) of J. Marion Sims's 1877 article, 'The Discovery of Anaesthesia' which exemplifies the stance of looking both inward and outward from the medical field. We begin by discussing Sims, including the complicated legacy he left as a physician. Next, we review late 19th-century history with a focus on Reconstruction. Finally, we show how the modern reader can use Sims's article both to trace the first use of ether and nitrous oxide for surgical anaesthesia and to provide a window into the 19th-century medical profession and the post-Civil War period. Through this study, we hope to show how to read both medicine and the world around it in physician histories. Published by the BMJ Publishing Group Limited. For permission to use (where not already granted under a licence) please go to http://www.bmj.com/company/products-services/rights-and-licensing/

  17. Holocene slip rates along the San Andreas Fault System in the San Gorgonio Pass and implications for large earthquakes in southern California

    NASA Astrophysics Data System (ADS)

    Heermance, Richard V.; Yule, Doug

    2017-06-01

    The San Gorgonio Pass (SGP) in southern California contains a 40 km long region of structural complexity where the San Andreas Fault (SAF) bifurcates into a series of oblique-slip faults with unknown slip history. We combine new 10Be exposure ages (Qt4: 8600 (+2100, -2200) and Qt3: 5700 (+1400, -1900) years B.P.) and a radiocarbon age (1260 ± 60 years B.P.) from late Holocene terraces with scarp displacement of these surfaces to document a Holocene slip rate of 5.7 (+2.7, -1.5) mm/yr combined across two faults. Our preferred slip rate is 37-49% of the average slip rates along the SAF outside the SGP (i.e., Coachella Valley and San Bernardino sections) and implies that strain is transferred off the SAF in this area. Earthquakes here most likely occur in very large, throughgoing SAF events at a lower recurrence than elsewhere on the SAF, so that only approximately one third of SAF ruptures penetrate or originate in the pass.Plain Language SummaryHow large are earthquakes on the southern San <span class="hlt">Andreas</span> Fault? The answer to this question depends on whether or not the earthquake is contained only along individual fault sections, such as the Coachella Valley section north of Palm Springs, or the rupture crosses multiple sections including the area through the San Gorgonio Pass. We have determined the age and offset of faulted stream deposits within the San Gorgonio Pass to document slip rates of these faults over the last 10,000 years. Our results indicate a long-term slip rate of 6 mm/yr, which is almost 1/2 of the rates east and west of this area. These new rates, combined with faulted geomorphic surfaces, imply that large magnitude earthquakes must occasionally rupture a 300 km length of the San <span class="hlt">Andreas</span> Fault from the Salton Sea to the Mojave Desert. Although many ( 65%) earthquakes along the southern San <span class="hlt">Andreas</span> Fault likely do not rupture through the pass, our new results suggest that large >Mw 7.5 earthquakes are possible</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.bssaonline.org/content/84/3/799.abstract','USGSPUBS'); return false;" href="http://www.bssaonline.org/content/84/3/799.abstract"><span>Continuous borehole strain in the San <span class="hlt">Andreas</span> fault zone before, during, and after the 28 June 1992, MW 7.3 Landers, California, earthquake</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Johnston, M.J.S.; Linde, A.T.; Agnew, D.C.</p> <p>1994-01-01</p> <p>High-precision strain was observed with a borehole dilational strainmeter in the Devil's Punchbowl during the 11:58 UT 28 June 1992 MW 7.3 Landers earthquake and the large Big Bear aftershock (MW 6.3). The strainmeter is installed at a depth of 176 m in the fault zone approximately midway between the surface traces of the San <span class="hlt">Andreas</span> and Punchbowl faults and is about 100 km from the 85-km-long Landers rupture. We have questioned whether unusual amplified strains indicating precursive slip or high fault compliance occurred on the faults ruptured by the Landers earthquake, or in the San <span class="hlt">Andreas</span> fault zone before and during the earthquake, whether static offsets for both the Landers and Big Bear earthquakes agree with expectation from geodetic and seismologic models of the ruptures and with observations from a nearby two-color geodimeter network, and whether postseismic behavior indicated continued slip on the Landers rupture or local triggered slip on the San <span class="hlt">Andreas</span>. We show that the strain observed during the earthquake at this instrument shows no apparent amplification effects. There are no indications of precursive strain in these strain data due to either local slip on the San <span class="hlt">Andreas</span> or precursive slip on the eventual Landers rupture. The observations are generally consistent with models of the earthquake in which fault geometry and slip have the same form as that determined by either inversion of the seismic data or inversion of geodetically determined ground displacements produced by the earthquake. Finally, there are some indications of minor postseismic behavior, particularly during the month following the earthquake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/fs/1999/fs105-99/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/fs/1999/fs105-99/"><span>Hydrogeologic investigations by the U.S. Geological Survey at the former Fort Benjamin Harrison, <span class="hlt">Marion</span> County, Indiana</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Risch, Martin R.</p> <p>1999-01-01</p> <p>As part of the U.S. Department of Defense Base Realignment and Closure process, the former Fort Benjamin Harrison in <span class="hlt">Marion</span> County, Indiana (called 'Fort Harrison' in this fact sheet), was placed on the Base Closure List in 1991. Property disposal and reuse activities began when Fort Harrison was decommissioned in 1995; work continues through 1999. Fort Harrison was located on approximately 2,500 acres about 10 miles northeast of downtown Indianapolis, Ind., in the City of Lawrence (fig. 1). Since 1903, the installation served as a major training facility that at times included schools, a hospital, and Army Finance and Soldier Support Centers. In 1996, the Army leased 1,700 acres of woodland and recreational facilities to the Indiana Department of Natural Resources as Fort Harrison State Park. Another 550 acres became privately owned for industrial, commercial, and residential purposes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740030066&hterms=plate+tectonics&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dplate%2Btectonics','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740030066&hterms=plate+tectonics&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dplate%2Btectonics"><span>The San <span class="hlt">Andreas</span> fault experiment. [gross tectonic plates relative velocity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smith, D. E.; Vonbun, F. O.</p> <p>1973-01-01</p> <p>A plan was developed during 1971 to determine gross tectonic plate motions along the San <span class="hlt">Andreas</span> Fault System in California. Knowledge of the gross motion along the total fault system is an essential component in the construction of realistic deformation models of fault regions. Such mathematical models will be used in the future for studies which will eventually lead to prediction of major earthquakes. The main purpose of the experiment described is the determination of the relative velocity of the North American and the Pacific Plates. This motion being so extremely small, cannot be measured directly but can be deduced from distance measurements between points on opposite sites of the plate boundary taken over a number of years.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2004/1389/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2004/1389/"><span>GPS-aided inertial technology and navigation-based photogrammetry for aerial mapping the San <span class="hlt">Andreas</span> fault system</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sanchez, Richard D.; Hudnut, Kenneth W.</p> <p>2004-01-01</p> <p>Aerial mapping of the San <span class="hlt">Andreas</span> Fault System can be realized more efficiently and rapidly without ground control and conventional aerotriangulation. This is achieved by the direct geopositioning of the exterior orientation of a digital imaging sensor by use of an integrated Global Positioning System (GPS) receiver and an Inertial Navigation System (INS). A crucial issue to this particular type of aerial mapping is the accuracy, scale, consistency, and speed achievable by such a system. To address these questions, an Applanix Digital Sensor System (DSS) was used to examine its potential for near real-time mapping. Large segments of vegetation along the San <span class="hlt">Andreas</span> and Cucamonga faults near the foothills of the San Bernardino and San Gabriel Mountains were burned to the ground in the California wildfires of October-November 2003. A 175 km corridor through what once was a thickly vegetated and hidden fault surface was chosen for this study. Both faults pose a major hazard to the greater Los Angeles metropolitan area and a near real-time mapping system could provide information vital to a post-disaster response.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20033046','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20033046"><span>Tremor-tide correlations and near-lithostatic pore pressure on the deep San <span class="hlt">Andreas</span> fault.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thomas, Amanda M; Nadeau, Robert M; Bürgmann, Roland</p> <p>2009-12-24</p> <p>Since its initial discovery nearly a decade ago, non-volcanic tremor has provided information about a region of the Earth that was previously thought incapable of generating seismic radiation. A thorough explanation of the geologic process responsible for tremor generation has, however, yet to be determined. Owing to their location at the plate interface, temporal correlation with geodetically measured slow-slip events and dominant shear wave energy, tremor observations in southwest Japan have been interpreted as a superposition of many low-frequency earthquakes that represent slip on a fault surface. Fluids may also be fundamental to the failure process in subduction zone environments, as teleseismic and tidal modulation of tremor in Cascadia and Japan and high Poisson ratios in both source regions are indicative of pressurized pore fluids. Here we identify a robust correlation between extremely small, tidally induced shear stress parallel to the San <span class="hlt">Andreas</span> fault and non-volcanic tremor activity near Parkfield, California. We suggest that this tremor represents shear failure on a critically stressed fault in the presence of near-lithostatic pore pressure. There are a number of similarities between tremor in subduction zone environments, such as Cascadia and Japan, and tremor on the deep San <span class="hlt">Andreas</span> transform, suggesting that the results presented here may also be applicable in other tectonic settings.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Tectp.693..474F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Tectp.693..474F"><span>Response of deformation patterns to reorganization of the southern San <span class="hlt">Andreas</span> fault system since ca. 1.5 Ma</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fattaruso, Laura A.; Cooke, Michele L.; Dorsey, Rebecca J.; Housen, Bernard A.</p> <p>2016-12-01</p> <p>Between 1.5 and 1.1 Ma, the southern San <span class="hlt">Andreas</span> fault system underwent a major reorganization that included initiation of the San Jacinto fault zone and termination of slip on the extensional West Salton detachment fault. The southern San <span class="hlt">Andreas</span> fault itself has also evolved since this time, with several shifts in activity among fault strands within San Gorgonio Pass. We use three-dimensional mechanical Boundary Element Method models to investigate the impact of these changes to the fault network on deformation patterns. A series of snapshot models of the succession of active fault geometries explore the role of fault interaction and tectonic loading in abandonment of the West Salton detachment fault, initiation of the San Jacinto fault zone, and shifts in activity of the San <span class="hlt">Andreas</span> fault. Interpreted changes to uplift patterns are well matched by model results. These results support the idea that initiation and growth of the San Jacinto fault zone led to increased uplift rates in the San Gabriel Mountains and decreased uplift rates in the San Bernardino Mountains. Comparison of model results for vertical-axis rotation to data from paleomagnetic studies reveals a good match to local rotation patterns in the Mecca Hills and Borrego Badlands. We explore the mechanical efficiency at each step in the modeled fault evolution, and find an overall trend toward increased efficiency through time. Strain energy density patterns are used to identify regions of incipient faulting, and support the notion of north-to-south propagation of the San Jacinto fault during its initiation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998BeAst...1..141S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998BeAst...1..141S"><span>Peter <span class="hlt">Andreas</span> Hansen und die astronomische Gemeinschaft - eine erste Auswertung des Hansen-Nachlasses.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schwarz, O.; Strumpf, M.</p> <p></p> <p>The literary assets of Peter <span class="hlt">Andreas</span> Hansen are deposited in the Staatsarchiv Hamburg, the Forschungs- und Landesbibliothek Gotha and the Thüringer Staatsarchiv Gotha. They were never systematically investigated. The authors present here some results of a first evaluation. It was possible to reconstruct the historical events with regard to the maintenance of the Astronomische Nachrichten and the Altona observatory in 1854. Hansen was a successful teacher for many young astronomers. His way of stimulating the evolution of astronomy followed Zach's tradition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19920010493','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19920010493"><span>Crustal deformation along the San <span class="hlt">Andreas</span>, California</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Li, Victor C.</p> <p>1992-01-01</p> <p>The goal is to achieve a better understanding of the regional and local deformation and crustal straining processes in western North America, particularly the effects of the San <span class="hlt">Andreas</span> and nearby faults on the spatial and temporal crustal deformation behavior. Construction of theoretical models based on the mechanics of coupled elastic plate, viscoelastic foundation and large scale crack mechanics provide a rational basis for the interpretation of seismic and aseismic anomalies and expedite efforts in forecasting the stability of plate boundary deformation. Special focus is placed on the three dimensional time dependent surface deformation due to localized slippage in a elastic layer coupled to a visco-elastic substrate. The numerical analysis is based on a 3-D boundary element technique. Extension to visco-elastic coupling demands the derivation of 3-D time dependent Green's function. This method was applied to analyze the viscoelastic surface displacements due to a dislocated embedded patch. Surface uplift as a function of time and position are obtained. Comparisons between surface uplift for long and short dislocated patches are made.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23752981','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23752981"><span>Rhazes in the renaissance of <span class="hlt">Andreas</span> Vesalius.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Compier, Abdul Haq</p> <p>2012-01-01</p> <p><span class="hlt">Andreas</span> Vesalius' (1514-64) first publication was a Paraphrasis of the ninth book of the Liber ad Almansorem, written by the Arab-Persian physician and alchemist Rhazes (854-925). The role of Rhazes in Vesalius' oeuvre has thus far been much disregarded. The different ways Rhazes recurs reveal an intellectual evolution in Vesalius' work. In the Paraphrasis, Vesalius subjects Rhazes to the authority of Galen in the context of the early sixteenth-century humanist campaign for the substitution of Arab influences by Greek 'originals'. Over the years Vesalius continues his work on Rhazes, but his approach becomes more internationalistic. Ultimately, Vesalius criticises Galen while expressing sympathy for the Arab author. This may be the more significant as Rhazes could have influenced Vesalius in the act of criticising Galen - critical discussions of Galen were available to Vesalius in Latin translations of Rhazes's Liber Continens. Although Vesalius never refers to the work, it is hardly possible he was unaware of it: similarities in structure, rhetoric and form between the Continens and the De humani corporis fabrica could support this hypothesis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70001345','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70001345"><span>Recurrence of seismic migrations along the central California segment of the San <span class="hlt">Andreas</span> fault system</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Wood, M.D.; Allen, S.S.</p> <p>1973-01-01</p> <p>VERIFICATIONS of tectonic concepts1 concerning seafloor spreading are emerging in a manner that has direct bearing on earthquake prediction. Although the gross pattern of worldwide seismicity contributed to the formulation of the plate tectonic hypothesis, it is the space-time characteristics of this seismicity that may contribute more toward understanding the kinematics and dynamics of the driving mechanism long speculated to originate in the mantle. If the lithosphere is composed of plates that move essentially as rigid bodies, then there should be seismic edge effects associated with this movement. It is these interplate effects, especially seismic migration patterns, that we discuss here. The unidirectional propagation at constant velocity (80 km yr-1 east to west) for earthquakes (M???7.2) on the Antblian fault for the period 1939 to 1956 (ref. 2) is one of the earliest observations of such a phenomenon. Similar studies3,4 of the Alaska Aleutian seismic zone and certain regions of the west coast of South America suggest unidirectional and recurring migrations of earthquakes (M???7.7) occur in these areas. Between these two regions along the great transform faults of the west coast of North America, there is some evidence 5 for unidirectional, constant velocity and recurrent migration of great earthquakes. The small population of earthquakes (M>7.2) in Savage's investigation5 indicates a large spatial gap along the San <span class="hlt">Andreas</span> system in central California from 1830 to 1970. Previous work on the seismicity of this gap in central California indicates that the recurrence curves remain relatively constant, independent of large earthquakes, for periods up to a century6. Recurrence intervals for earthquakes along the San <span class="hlt">Andreas</span> Fault have been calculated empirically by Wallace7 on the basis of geological evidence, surface measurements and assumptions restricted to the surficial seismic layer. Here we examine the evidence for recurrence of seismic migrations along</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70189840','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70189840"><span>Constraints on the stress state of the San <span class="hlt">Andreas</span> fault with analysis based on core and cuttings from SAFOD drilling phases I and II</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lockner, David A.; Tembe, Cheryl; Wong, Teng-fong</p> <p>2009-01-01</p> <p>Analysis of field data has led different investigators to conclude that the San <span class="hlt">Andreas</span> Fault (SAF) has either anomalously low frictional sliding strength (m < 0.2) or strength consistent with standard laboratory tests (m > 0.6). Arguments for the apparent weakness of the SAF generally hinge on conceptual models involving intrinsically weak gouge or elevated pore pressure within the fault zone. Some models assert that weak gouge and/or high pore pressure exist under static conditions while others consider strength loss or fluid pressure increase due to rapid coseismic fault slip. The present paper is composed of three parts. First, we develop generalized equations, based on and consistent with the Rice (1992) fault zone model to relate stress orientation and magnitude to depth-dependent coefficient of friction and pore pressure. Second, we present temperature- and pressure-dependent friction measurements from wet illite-rich fault gouge extracted from San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) phase 1 core samples and from weak minerals associated with the San <span class="hlt">Andreas</span> Fault. Third, we reevaluate the state of stress on the San <span class="hlt">Andreas</span> Fault in light of new constraints imposed by SAFOD borehole data. Pure talc (m0.1) had the lowest strength considered and was sufficiently weak to satisfy weak fault heat flow and stress orientation constraints with hydrostatic pore pressure. Other fault gouges showed a systematic increase in strength with increasing temperature and pressure. In this case, heat flow and stress orientation constraints would require elevated pore pressure and, in some cases, fault zone pore pressure in excess of vertical stress.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/15002023','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/15002023"><span>Aseismic Slip Events along the Southern San <span class="hlt">Andreas</span> Fault System Captured by Radar Interferometry</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Vincent, P</p> <p>2001-10-01</p> <p>A seismic slip is observed along several faults in the Salton Sea and southernmost Landers rupture zone regions using interferometric synthetic aperture radar (InSAR) data spanning different time periods between 1992 and 1997. In the southernmost Landers rupture zone, projecting south from the Pinto Mountain Fault, sharp discontinuities in the interferometric phase are observed along the sub-parallel Burnt Mountain and Eureka Peak Faults beginning three months after the Landers earthquake and is interpreted to be post-Landers after-slip. Abrupt phase offsets are also seen along the two southernmost contiguous 11 km Durmid Hill and North Shore segments of the San Andreasmore » Fault with an abrupt termination of slip near the northern end of the North Shore Segment. A sharp phase offset is seen across 20 km of the 30 km-long Superstition Hills Fault before phase decorrelation in the Imperial Valley along the southern 10 km of the fault prevents coherent imaging by InSAR. A time series of deformation interferograms suggest most of this slip occurred between 1993 and 1995 and none of it occurred between 1992 and 1993. A phase offset is also seen along a 5 km central segment of the Coyote Creek fault that forms a wedge with an adjoining northeast-southwest trending conjugate fault. Most of the slip observed on the southern San <span class="hlt">Andreas</span> and Superstition Hills Faults occurred between 1993 and 1995--no slip is observed in the 92-93 interferograms. These slip events, especially the Burnt Mountain and Eureka Peak events, are inferred to be related to stress redistribution from the June, 1992 M{sub w} = 7.3 Landers earthquake. Best-fit elastic models of the San <span class="hlt">Andreas</span> and Superstition Hills slip events suggest source mechanisms with seismic moments over three orders of magnitude larger than a maximum possible summation of seismic moments from all seismicity along each fault segment during the entire 4.8-year time interval spanned by the InSAR data. Aseismic moment releases</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1986JGR....9114080F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1986JGR....9114080F"><span>Distinctive Triassic megaporphyritic monzogranite: Evidence for only 160 km offset along the San <span class="hlt">Andreas</span> Fault, southern California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Frizzell, Virgil A., Jr.; Mattinson, James M.; Matti, Jonathan C.</p> <p>1986-12-01</p> <p>Distinctive megaporphyritic bodies of monzogranite to quartz monzonite that occur in the Mill Creek region of the San Bernardino Mountains and across the San <span class="hlt">Andreas</span> fault on Liebre Mountain share identical modal and chemical compositions, intrusive ages, and petrogenesis and similar thermal histories. Both bodies are strontium-rich and contain large potassium feldspar phenocrysts and hornblende. U-Pb determinations on zircon from both bodies indicate Triassic intrusive ages (215 Ma) and derivation, in part, from homogeneous Precambrian continental crust. U-Pb analyses on apatite and sphene and K-Ar analyses on hornblende and biotite show that the bodies suffered a Late Cretaceous thermal event (70-75 Ma). The strong similarities between the two bodies suggest that they constitute segments of a formerly continuous pluton that has been offset about 160 km by movement on the San <span class="hlt">Andreas</span> fault, about 80 km less than the generally accepted distance. Plutons having monzonitic compositions, reassembled with the megaporphyritic bodies are used as a piercing point, form a relatively coherent province within the varied suite of Mesozoic batholithic and prebatholithic rocks in southern California.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/2722366','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/2722366"><span>Injuries and risk factors for injuries from the 29 May 1982 tornado, <span class="hlt">Marion</span>, Illinois.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Duclos, P J; Ing, R T</p> <p>1989-03-01</p> <p>On 29 May 1982, a tornado struck <span class="hlt">Marion</span>, Illinois. As a result, ten people were killed, 38 were hospitalized, and 196 were examined in hospital emergency rooms. In order to describe injuries and to identify risk factors that might have increased or reduced the risk of injuries, we (1) conducted a telephone survey of all persons we could identify who were in the path of the tornado and (2) reviewed the emergency room logs and records and admission files of hospitals that provided initial care. Overall, 19.3% of people who were in the path of the tornado were injured during the impact. Only 39.0% of people examined in emergency rooms were injured as a direct result of the impact. The remainder were injured while rescuing, cleaning up, or just walking in the disaster area. Behaviours that were most protective were being in a house on the lowest floor, staying underground in a basement, protecting one's body with something like a blanket, and staying away from windows. Neither lying down nor opening doors or windows seemed to decrease the risk of being injured. Our results emphasize the need for early warnings and public education about protective behaviours.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6995558-health-hazard-evaluation-report-heta-thomson-consumer-electronics-marion-indiana','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6995558-health-hazard-evaluation-report-heta-thomson-consumer-electronics-marion-indiana"><span>Health-hazard evaluation report HETA 90-223-2211, Thomson Consumer Electronics, <span class="hlt">Marion</span>, Indiana</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Lenhart, S.W.; Driscoll, R.</p> <p>1992-05-01</p> <p>In response to a request from the Corporate Medical Consultant to Thomson Consumer Electronics (SIC-3673), <span class="hlt">Marion</span>, Indiana, a study was undertaken of an illness outbreak in workers at the facility. There were about 1900 workers at the facility, which produced television picture tubes. Production occurred over three shifts, 6 days a week. Charcoal tube sampling indicated the presence of acetone (67641) n-amyl-acetate (628637), n-butyl-acetate (123864), isoamyl-acetate (123922), toluene (108883), 1,1,1-trichloroethane (71556), and trichloroethylene (79016). No contaminants were detected in the bag samples of air collected from the in/house compressed air system. One or more symptoms were reported by 593 (82%)more » of the workers. Those most commonly reported included headache (68%), sore throat (53%), fatigue (51%), eye irritation (50%), itchy skin (47%), irritated nose (45%), dizziness (45%), unusual taste in mouth (45%), unusual smell (41%) and cough. The authors conclude that symptoms were consistent with stress related health complaints in occupational settings. Concentrations of chemicals measured in the facility would not be expected to produce the effects seen in the outbreak. The authors recommend that trichloroethylene degreasing units be replaced with equipment which uses a less toxic degreasing agent. The facility should hire a full time industrial hygienist.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AGUSM.G21A..04B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AGUSM.G21A..04B"><span>Holocene deceleration of the San <span class="hlt">Andreas</span> fault zone in San Bernardino and implications for the eastern California shear zone rate debate</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bennett, R. A.; Lavier, L.; Anderson, M. L.; Matti, J.; Powell, R. E.</p> <p>2005-05-01</p> <p>New geodetic inferences for the rate of strain accumulation on the San <span class="hlt">Andreas</span> fault associated with tectonic loading are ~20 mm/yr slower than observed Holocene surface displacement rates in the San Bernardino area, south of the fault's intersection with the San Jacinto fault zone, and north of its intersection with the eastern California shear zone (ECSZ). This displacement rate "anomaly" is significantly larger than can be easily explained by locking depth errors or earthquake cycle effects not accounted for in geodesy-constrained models for elastic loading rate. Using available time-averaged fault displacement-rates for the San <span class="hlt">Andreas</span> and San Jacinto fault zones, we estimate instantaneous time-variable displacement rates on the San <span class="hlt">Andreas</span>-San Jacinto-ECSZ fault zones, assuming that these fault zones form a closed system in the latitude band along which the fault zones overlap with one another and share in the accommodation of steady Pacific-North America relative plate motion. We find that the Holocene decrease in San <span class="hlt">Andreas</span> loading rate can be compensated by a rapid increase in loading/displacement rate within the ECSZ over the past ~5 kyrs, independent of, but consistent with geodetic and geologic constraints derived from the ECSZ itself. Based on this model, we suggest that reported differences between fast contemporary strain rates observed on faults of the ECSZ using geodesy and slow rates inferred from Quaternary geology and Holocene paleoseismology (i.e., the ECSZ rate debate) may be explained by rapid changes in the pattern and rates of strain accumulation associated with fault loading largely unrelated to postseismic stress relaxation. If so, displacement rate data sets from Holocene geology and present-day geodesy could potentially provide important new constraints on the rheology of the lower crust and upper mantle representing lithospheric behavior on time-scales of thousands of years. Moreover, the results underscore that disagreement between</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018187','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018187"><span>Strength of chrysotile-serpentinite gouge under hydrothermal conditions: Can it explain a weak San <span class="hlt">Andreas</span> fault?</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Moore, Diane E.; Lockner, D.A.; Summers, R.; Shengli, M.; Byerlee, J.D.</p> <p>1996-01-01</p> <p>Chrysotile-bearing serpentinite is a constituent of the San <span class="hlt">Andreas</span> fault zone in central and northern California. At room temperature, chrysotile gouge has a very low coefficient of friction (?? ??? 0.2), raising the possibility that under hydrothermal conditions ?? might be reduced sufficiently (to ???0.1) to explain the apparent weakness of the fault. To test this hypothesis, we measured the frictional strength of a pure chrysotile gouge at temperatures to 290??C and axial-shortening velocities as low as 0.001 ??m/s. As temperature increases to ???100??C, the strength of the chrysotile gouge decreases slightly at low velocities, but at temperatures ???200??C, it is substantially stronger and essentially independent of velocity at the lowest velocities tested. We estimate that pure chrysotile gouge at hydrostatic fluid pressure and appropriate temperatures would have shear strength averaged over a depth of 14 km of 50 MPa. Thus, on the sole basis of its strength, chrysotile cannot be the cause of a weak San <span class="hlt">Andreas</span> fault. However, chrysotile may also contribute to low fault strength by forming mineral seals that promote the development of high fluid pressures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/8209577','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/8209577"><span>[<span class="hlt">Andreas</span> Vesalius and surgery].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Van Hee, R</p> <p>1993-01-01</p> <p>By publishing De Humani Corporis Fabrica Libri Septem in 1543, Andries van Wesel (1514-1564) gave surgical science an immense impulse. The revolutionary renovation in the knowledge of man's anatomical structure changed slowly and progressively into topographical and physiological understanding of surgical diseases. At the same time, this made better aimed and more secure operations possible. Apart from the importance of this anatomical publication, <span class="hlt">Andreas</span> Vesalius also won his spurs as a surgeon. He taught surgery in Padua for many years. He was appointed court physician and surgeon at the Habsburg Court of Charles V and Philip II. He personally performed lots of operations known at the time as major ones. He not only quickly adopted the surgical innovations of his fellow-surgeon Ambroise Paré, but he even performed operations that had been forgotten during several centuries, among which thoracocentesis for pleural empyema. His clinical perspicacity in discovering the indication for some operations was staggering and was appreciated by all great monarchs of Europe in the 16th century. In his several consilia, numerous pieces of advice were given for the treatment of surgical patients. The surgical practice which Vesalius had in Brussels for many years, consequently became most successful. Many publications by Vesalius about surgery and blood-letting are well-known. His Chirurgia magna in septem Libros digesta still remains controversial; these books were published by Prospero Borgarruccio (1560) in 1568 by the Venetian editor Valgrisi. This book gives an excellent survey of surgical pathology as it was taught and treated in the 16th century. The scientific method that Vesalius used, not only in his anatomical studies but also in his surgical practice, deserves not only our full appraisal but should still be studied in our own time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3314893','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3314893"><span>Rhazes in the Renaissance of <span class="hlt">Andreas</span> Vesalius</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Compier, Abdul Haq</p> <p>2012-01-01</p> <p><span class="hlt">Andreas</span> Vesalius' (1514–64) first publication was a Paraphrasis of the ninth book of the Liber ad Almansorem, written by the Arab–Persian physician and alchemist Rhazes (854–925). The role of Rhazes in Vesalius' oeuvre has thus far been much disregarded. The different ways Rhazes recurs reveal an intellectual evolution in Vesalius' work. In the Paraphrasis, Vesalius subjects Rhazes to the authority of Galen in the context of the early sixteenth-century humanist campaign for the substitution of Arab influences by Greek ‘originals’. Over the years Vesalius continues his work on Rhazes, but his approach becomes more internationalistic. Ultimately, Vesalius criticises Galen while expressing sympathy for the Arab author. This may be the more significant as Rhazes could have influenced Vesalius in the act of criticising Galen – critical discussions of Galen were available to Vesalius in Latin translations of Rhazes's Liber Continens. Although Vesalius never refers to the work, it is hardly possible he was unaware of it: similarities in structure, rhetoric and form between the Continens and the De humani corporis fabrica could support this hypothesis. PMID:23752981</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29797278','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29797278"><span>Illness as the saturated phenomenon: the contribution of Jean-Luc <span class="hlt">Marion</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Grīnfelde, Māra</p> <p>2018-05-24</p> <p>During the last few decades, many thinkers have advocated for the importance of the phenomenological approach in developing the understanding of the lived experience of illness. In their attempts, they have referred to ideas found in the history of phenomenology, most notably, in the works of Edmund Husserl, Martin Heidegger, Maurice Merleau-Ponty and Jean-Paul Sartre. The aim of this paper is to sketch out an interpretation of illness based on a yet unexplored conceptual framework of the phenomenology of French thinker Jean-Luc <span class="hlt">Marion</span>. Focusing on concepts of the saturated phenomenon and flesh, the paper develops an interpretation of illness as the saturated phenomenon, which highlights a variety of dimensions of illness already elaborated within the phenomenology of medicine, such as the affective dimension of illness, the disruptive dimension of illness, the transformed perception of the self in illness, mineness of flesh in illness and the inexpressible and hermeneutical dimension of illness. In addition to that, the paper explores some of the consequences the proposed interpretation of illness offers regarding the nature of illness and health. It is argued that illness in its essence is very similar to the experience of other saturated phenomena, suggesting that the difference between them does not lie within the character of the affective givenness, but rather within the dynamic relationship between the affective givenness and its conceptualization. It is also shown that the experience of health is compatible with the experience of saturation and thus is not limited to the tacit and harmonious background state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70011055','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70011055"><span>Tilt precursors before earthquakes on the San <span class="hlt">Andreas</span> fault, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Johnston, M.J.S.; Mortensen, C.E.</p> <p>1974-01-01</p> <p>An array of 14 biaxial shallow-borehole tiltmeters (at 10-7 radian sensitivity) has been installed along 85 kilometers of the San <span class="hlt">Andreas</span> fault during the past year. Earthquake-related changes in tilt have been simultaneously observed on up to four independent instruments. At earthquake distances greater than 10 earthquake source dimensions, there are few clear indications of tilt change. For the four instruments with the longest records (>10 months), 26 earthquakes have occurred since July 1973 with at least one instrument closer than 10 source dimensions and 8 earthquakes with more than one instrument within that distance. Precursors in tilt direction have been observed before more than 10 earthquakes or groups of earthquakes, and no similar effect has yet been seen without the occurrence of an earthquake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45.2274K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45.2274K"><span>Multiscale Dynamics of Aseismic Slip on Central San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khoshmanesh, M.; Shirzaei, M.</p> <p>2018-03-01</p> <p>Understanding the evolution of aseismic slip enables constraining the fault's seismic budget and provides insight into dynamics of creep. Inverting the time series of surface deformation measured along the Central San <span class="hlt">Andreas</span> Fault obtained from interferometric synthetic aperture radar in combination with measurements of repeating earthquakes, we constrain the spatiotemporal distribution of creep during 1992-2010. We identify a new class of intermediate-term creep rate variations that evolve over decadal scale, releasing stress on the accelerating zone and loading adjacent decelerating patches. We further show that in short-term (<2 year period), creep avalanches, that is, isolated clusters of accelerated aseismic slip with velocities exceeding the long-term rate, govern the dynamics of creep. The statistical properties of these avalanches suggest existence of elevated pore pressure in the fault zone, consistent with laboratory experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22514000','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22514000"><span>The core competencies of James <span class="hlt">Marion</span> Sims, MD.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Straughn, J Michael; Gandy, Roy E; Rodning, Charles B</p> <p>2012-07-01</p> <p>The concept of core competencies in graduate medical education was introduced by the Accreditation Council for Graduate Medical Education of the American Medical Association to semiquantitatively assess the professional performance of students, residents, practitioners, and faculty. Many aspects of the career of J. <span class="hlt">Marion</span> Sims, MD, are exemplary of those core competencies: MEDICAL KNOWLEDGE: Author of the first American textbook related to gynecology. MEDICAL CARE: Innovator of the Sims' Vaginal Speculum, Sims' Position, Sims' Test, and vesico-/rectovaginal fistulorrhaphy; advocated abdominal exploration for penetrating wounds; performed the first cholecystostomy. PROFESSIONALISM: Served as President of the New York Academy of Medicine, the American Medical Association, and the American Gynecologic Society. INTERPERSONAL RELATIONSHIPS/COMMUNICATION: Cared for the indigent, hearthless, indentured, disenfranchised; served as consulting surgeon to the Empress Eugénie (France), the Duchess of Hamilton (Scotland), the Empress of Austria, and other royalty of the aristocratic Houses of Europe; accorded the National Order of the Legion of Honor. PRACTICE-BASED LEARNING: Introduction of silver wire sutures; adoption of the principles of asepsis/antisepsis; adoption of the principles of general anesthesia. SYSTEMS-BASED PRACTICE: Established the Woman's Hospital, New York City, New York, the predecessor of the Memorial Sloan-Kettering Center for the Treatment of Cancer and Allied Diseases; organized the Anglo-American Ambulance Corps under the patronage of Napoleon III. What led him to a life of clinical and humanitarian service? First, he was determined to succeed. His formal medical/surgical education was perhaps the best available to North Americans during that era. Second, he was courageous in experimentation and innovation, applying new developments in operative technique, asepsis/antisepsis, and general anesthesia. Third, his curiosity was not burdened by rigid</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6473057-tennol-inc-loan-guarantee-application-marion-county-tennessee-environmental-assessment','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6473057-tennol-inc-loan-guarantee-application-marion-county-tennessee-environmental-assessment"><span>Tennol, Inc. - loan-guarantee application: <span class="hlt">Marion</span> County, Tennessee. Environmental assessment</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Not Available</p> <p>1982-12-01</p> <p>Tennol, Inc., is requesting a loan guarantee to build a fuel ethanol plant having an annual capacity of 95,000 m/sup 3/ (25 x 10/sup 6/ gal) in Jasper, Tennessee. The plant will wet-mill corn (255 Gg/year) for feedstock and will burn coal (43 Gg/year) for process heat. Marketable by-products include corn gluten feed (61 Gg/year), corn gluten meal (17 Gg/year), and carbon dioxide (71 Gg/year). Dehydration of the product alcohol will be azeotropic distillation with diethyl ether, and denaturation will be with unleaded gasoline. No degradation of local or regional air quality is anticipated. A combination of wetlands, old fields,more » and bottomland hardwood forest provides excellent wildlife habitat. Over 75% of the site will remain undisturbed, and construction activity will be confined to the northern end of the site. Because no wet areas will be disturbed and all liquid effluent releases will be to the Jasper wastewater treatment plant, no impact on aquatic organisms or water quality is anticipated. Impact to terrestrial organisms will be minor, and no rare or endangered species will be affected. Water will be supplied by the town of Jasper and by onsite wells. Tennol will represent about 3% of total ground water use in <span class="hlt">Marion</span> County, but no impact is anticipated since the region has abundant groundwater resources. Although both historic and archaeological sites occur on the property, these sites lie outside the area of construction impact. Socioeconomic impact will be positive. Approximately 200 people will be employed at the peak of construction, and practically all are expected to commute. Of the 100 to 120 permanent jobs that will be created, most are expected to go to local people.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19880011370','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19880011370"><span>Geodetic Measurement of Deformation East of the San <span class="hlt">Andreas</span> Fault in Central California</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sauber, Jeanne M.; Lisowski, Michael; Solomon, Sean C.</p> <p>1988-01-01</p> <p>Triangulation and trilateration data from two geodetic networks located between the western edge of the Great Valley and the San <span class="hlt">Andreas</span> fault have been used to calculate shear strain rates in the Diablo Range and to estimate the slip rate along the Calaveras and Paicines faults in Central California. Within the Diablo Range the average shear strain rate was determined for the time period between 1962 and 1982 to be 0.15 + or - 0.08 microrad/yr, with the orientation of the most compressive strain at N 16 deg E + or - 14 deg. The orientation of the principal compressive strain predicted from the azimuth of the major structures in the region is N 25 deg E. It is inferred that the measured strain is due to compression across the folds of this area: the average shear straining corresponds to a relative shortening rate of 4.5 + or - 2.4 mm/yr. From an examination of wellbore breakout orientations and the azimuths of P-axes from earthquake focal mechanisms the inferred orientation of maximum compressive stress was found to be similar to the direction of maximum compressive strain implied by the trend of local fold structures. Results do not support the hypothesis of uniform fault-normal compression within the Coast Ranges. From trilateration measurements made between 1972 and 1987 on lines that are within 10 km of the San <span class="hlt">Andreas</span> fault, a slip rate of 10 to 12 mm/yr was calculated for the Calaveras-Paicines fault south of Hollister. The slip rate of the Paicines fault decreases to 4 mm/yr near Bitter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2563360','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2563360"><span>The medical ethics of Dr J <span class="hlt">Marion</span> Sims: a fresh look at the historical record</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wall, L L</p> <p>2006-01-01</p> <p>Vesicovaginal fistula was a catastrophic complication of childbirth among 19th century American women. The first consistently successful operation for this condition was developed by Dr J <span class="hlt">Marion</span> Sims, an Alabama surgeon who carried out a series of experimental operations on black slave women between 1845 and 1849. Numerous modern authors have attacked Sims's medical ethics, arguing that he manipulated the institution of slavery to perform ethically unacceptable human experiments on powerless, unconsenting women. This article reviews these allegations using primary historical source material and concludes that the charges that have been made against Sims are largely without merit. Sims's modern critics have discounted the enormous suffering experienced by fistula victims, have ignored the controversies that surrounded the introduction of anaesthesia into surgical practice in the middle of the 19th century, and have consistently misrepresented the historical record in their attacks on Sims. Although enslaved African American women certainly represented a “vulnerable population” in the 19th century American South, the evidence suggests that Sims's original patients were willing participants in his surgical attempts to cure their affliction—a condition for which no other viable therapy existed at that time. PMID:16731734</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16731734','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16731734"><span>The medical ethics of Dr J <span class="hlt">Marion</span> Sims: a fresh look at the historical record.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wall, L L</p> <p>2006-06-01</p> <p>Vesicovaginal fistula was a catastrophic complication of childbirth among 19th century American women. The first consistently successful operation for this condition was developed by Dr J <span class="hlt">Marion</span> Sims, an Alabama surgeon who carried out a series of experimental operations on black slave women between 1845 and 1849. Numerous modern authors have attacked Sims's medical ethics, arguing that he manipulated the institution of slavery to perform ethically unacceptable human experiments on powerless, unconsenting women. This article reviews these allegations using primary historical source material and concludes that the charges that have been made against Sims are largely without merit. Sims's modern critics have discounted the enormous suffering experienced by fistula victims, have ignored the controversies that surrounded the introduction of anaesthesia into surgical practice in the middle of the 19th century, and have consistently misrepresented the historical record in their attacks on Sims. Although enslaved African American women certainly represented a "vulnerable population" in the 19th century American South, the evidence suggests that Sims's original patients were willing participants in his surgical attempts to cure their affliction-a condition for which no other viable therapy existed at that time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70187066','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70187066"><span>Simulations of tremor-related creep reveal a weak crustal root of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Shelly, David R.; Bradley, Andrew M.; Johnson, Kaj M.</p> <p>2013-01-01</p> <p>Deep aseismic roots of faults play a critical role in transferring tectonic loads to shallower, brittle crustal faults that rupture in large earthquakes. Yet, until the recent discovery of deep tremor and creep, direct inference of the physical properties of lower-crustal fault roots has remained elusive. Observations of tremor near Parkfield, CA provide the first evidence for present-day localized slip on the deep extension of the San <span class="hlt">Andreas</span> Fault and triggered transient creep events. We develop numerical simulations of fault slip to show that the spatiotemporal evolution of triggered tremor near Parkfield is consistent with triggered fault creep governed by laboratory-derived friction laws between depths of 20–35 km on the fault. Simulated creep and observed tremor northwest of Parkfield nearly ceased for 20–30 days in response to small coseismic stress changes of order 104 Pa from the 2003 M6.5 San Simeon Earthquake. Simulated afterslip and observed tremor following the 2004 M6.0 Parkfield earthquake show a coseismically induced pulse of rapid creep and tremor lasting for 1 day followed by a longer 30 day period of sustained accelerated rates due to propagation of shallow afterslip into the lower crust. These creep responses require very low effective normal stress of ~1 MPa on the deep San <span class="hlt">Andreas</span> Fault and near-neutral-stability frictional properties expected for gabbroic lower-crustal rock.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01645&hterms=red+tide&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dred%2Btide','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01645&hterms=red+tide&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dred%2Btide"><span>The San <span class="hlt">Andreas</span> Fault and a Strike-slip Fault on Europa</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1998-01-01</p> <p><p/>The mosaic on the right of the south polar region of Jupiter's moon Europa shows the northern 290 kilometers (180 miles) of a strike-slip fault named Astypalaea Linea. The entire fault is about 810 kilometers (500 miles) long, the size of the California portion of the San <span class="hlt">Andreas</span> fault on Earth which runs from the California-Mexico border north to the San Francisco Bay. <p/>The left mosaic shows the portion of the San <span class="hlt">Andreas</span> fault near California's san Francisco Bay that has been scaled to the same size and resolution as the Europa image. Each covers an area approximately 170 by 193 kilometers(105 by 120 miles). The red line marks the once active central crack of the Europan fault (right) and the line of the San <span class="hlt">Andreas</span> fault (left). <p/>A strike-slip fault is one in which two crustal blocks move horizontally past one another, similar to two opposing lanes of traffic. The overall motion along the Europan fault seems to have followed a continuous narrow crack along the entire length of the feature, with a path resembling stepson a staircase crossing zones which have been pulled apart. The images show that about 50 kilometers (30 miles) of displacement have taken place along the fault. Opposite sides of the fault can be reconstructed like a puzzle, matching the shape of the sides as well as older individual cracks and ridges that had been broken by its movements. <p/>Bends in the Europan fault have allowed the surface to be pulled apart. This pulling-apart along the fault's bends created openings through which warmer, softer ice from below Europa's brittle ice shell surface, or frozen water from a possible subsurface ocean, could reach the surface. This upwelling of material formed large areas of new ice within the boundaries of the original fault. A similar pulling apart phenomenon can be observed in the geological trough surrounding California's Salton Sea, and in Death Valley and the Dead Sea. In those cases, the pulled apart regions can include upwelled</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25733589','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25733589"><span>Animals, Pictures, and Skeletons: <span class="hlt">Andreas</span> Vesalius's Reinvention of the Public Anatomy Lesson.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shotwell, R Allen</p> <p>2016-01-01</p> <p>In this paper, I examine the procedures used by <span class="hlt">Andreas</span> Vesalius for conducting public dissections in the early sixteenth century. I point out that in order to overcome the limitations of public anatomical demonstration noted by his predecessors, Vesalius employed several innovative strategies, including the use of animals as dissection subjects, the preparation and display of articulated skeletons, and the use of printed and hand-drawn illustrations. I suggest that the examination of these three strategies for resolving the challenges of public anatomical demonstration helps us to reinterpret Vesalius's contributions to sixteenth-century anatomy. © The Author 2015. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70100270','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70100270"><span>Steep-dip seismic imaging of the shallow San <span class="hlt">Andreas</span> Fault near Parkfield</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hole, J.A.; Catchings, R.D.; St. Clair, K.C.; Rymer, M.J.; Okaya, D.A.; Carney, B.J.</p> <p>2001-01-01</p> <p>Seismic reflection and refraction images illuminate the San <span class="hlt">Andreas</span> Fault to a depth of 1 kilometer. The prestack depth-migrated reflection image contains near-vertical reflections aligned with the active fault trace. The fault is vertical in the upper 0.5 kilometer, then dips about 70° to the southwest to at least 1 kilometer subsurface. This dip reconciles the difference between the computed locations of earthquakes and the surface fault trace. The seismic velocity cross section shows strong lateral variations. Relatively low velocity (10 to 30%), high electrical conductivity, and low density indicate a 1-kilometer-wide vertical wedge of porous sediment or fractured rock immediately southwest of the active fault trace.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70032933','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70032933"><span>Annual modulation of seismicity along the San <span class="hlt">Andreas</span> Fault near Parkfield, CA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Christiansen, L.B.; Hurwitz, S.; Ingebritsen, S.E.</p> <p>2007-01-01</p> <p>We analyze seismic data from the San <span class="hlt">Andreas</span> Fault (SAF) near Parkfield, California, to test for annual modulation in seismicity rates. We use statistical analyses to show that seismicity is modulated with an annual period in the creeping section of the fault and a semiannual period in the locked section of the fault. Although the exact mechanism for seasonal triggering is undetermined, it appears that stresses associated with the hydrologic cycle are sufficient to fracture critically stressed rocks either through pore-pressure diffusion or crustal loading/ unloading. These results shed additional light on the state of stress along the SAF, indicating that hydrologically induced stress perturbations of ???2 kPa may be sufficient to trigger earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26122490','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26122490"><span><span class="hlt">Andreas</span> Vesalius' five hundreth anniversary: initiation of the rotator cuff concept.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brinkman, Romy J; Hage, J Joris</p> <p>2015-12-01</p> <p>The rotator cuff concept refers to the four scapulohumeral muscles that stabilize and rotate the humerus relative to the scapula. To date, the first description of the rotator cuff remained unidentified. In light of the 500th birthday of <span class="hlt">Andreas</span> Vesalius (1515-1564) we searched his 1543 masterwork "Fabrica Corporis Humani Libri Septem" for references to the morphology and function of the rotator cuff muscles. Even though he distinguished three rather than four scapulohumeral muscles, Vesalius recognized the need for structures that prevent dislocation of the shoulder inherent to the morphology of the humeral caput and scapular socket. He recorded "three strong ligaments" and the "three muscles that rotate the arm" of which the tendons completely "embrace the ligaments of the joint" as such structures. Vesalius defined the rotator cuff concept avant la lettre.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.S21A2141Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.S21A2141Y"><span>Long Return Periods for Earthquakes in San Gorgonio Pass and Implications for Large Ruptures of the San <span class="hlt">Andreas</span> Fault in Southern California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yule, J.; McBurnett, P.; Ramzan, S.</p> <p>2011-12-01</p> <p>The largest discontinuity in the surface trace of the San <span class="hlt">Andreas</span> fault occurs in southern California at San Gorgonio Pass. Here, San <span class="hlt">Andreas</span> motion moves through a 20 km-wide compressive stepover on the dextral-oblique-slip thrust system known as the San Gorgonio Pass fault zone. This thrust-dominated system is thought to rupture during very large San <span class="hlt">Andreas</span> events that also involve strike-slip fault segments north and south of the Pass region. A wealth of paleoseismic data document that the San <span class="hlt">Andreas</span> fault segments on either side of the Pass, in the San Bernardino/Mojave Desert and Coachella Valley regions, rupture on average every ~100 yrs and ~200 yrs, respectively. In contrast, we report here a notably longer return period for ruptures of the San Gorgonio Pass fault zone. For example, features exposed in trenches at the Cabezon site reveal that the most recent earthquake occurred 600-700 yrs ago (this and other ages reported here are constrained by C-14 calibrated ages from charcoal). The rupture at Cabezon broke a 10 m-wide zone of east-west striking thrusts and produced a >2 m-high scarp. Slip during this event is estimated to be >4.5 m. Evidence for a penultimate event was not uncovered but presumably lies beneath ~1000 yr-old strata at the base of the trenches. In Millard Canyon, 5 km to the west of Cabezon, the San Gorgonio Pass fault zone splits into two splays. The northern splay is expressed by 2.5 ± 0.7 m and 5.0 ± 0.7 m scarps in alluvial terraces constrained to be ~1300 and ~2500 yrs old, respectively. The scarp on the younger, low terrace postdates terrace abandonment ~1300 yrs ago and probably correlates with the 600-700 yr-old event at Cabezon, though we cannot rule out that a different event produced the northern Millard scarp. Trenches excavated in the low terrace reveal growth folding and secondary faulting and clear evidence for a penultimate event ~1350-1450 yrs ago, during alluvial deposition prior to the abandonment of the low terrace</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.G11B..07P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.G11B..07P"><span>Evaluation of LiDAR Imagery as a Tool for Mapping the Northern San <span class="hlt">Andreas</span> Fault in Heavily Forested Areas of Mendocino and Sonoma Counties, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prentice, C. S.; Koehler, R. D.; Baldwin, J. N.; Harding, D. J.</p> <p>2004-12-01</p> <p>We are mapping in detail active traces of the San <span class="hlt">Andreas</span> Fault in Mendocino and Sonoma Counties in northern California, using recently acquired airborne LiDAR (also known as ALSM) data. The LiDAR data set provides a powerful new tool for mapping geomorphic features related to the San <span class="hlt">Andreas</span> Fault because it can be used to produce high-resolution images of the ground surfaces beneath the forest canopy along the 70-km-long section of the fault zone encompassed by the data. Our effort represents the first use of LiDAR data to map active fault traces in a densely vegetated region along the San <span class="hlt">Andreas</span> Fault. We are using shaded relief images generated from bare-earth DEMs to conduct detailed mapping of fault-related geomorphic features (e.g. scarps, offset streams, linear valleys, shutter ridges, and sag ponds) between Fort Ross and Point Arena. Initially, we map fault traces digitally, on-screen, based only on the geomorphology interpreted from LiDAR images. We then conduct field reconnaissance using the initial computer-based maps in order to verify and further refine our mapping. We found that field reconnaissance is of utmost importance in producing an accurate and detailed map of fault traces. Many lineaments identified as faults from the on-screen images were determined in the field to be old logging roads or other features unrelated to faulting. Also, in areas where the resolution of LiDAR data is poor, field reconnaissance, coupled with topographic maps and aerial photographs, permits a more accurate location of fault-related geomorphic features. LiDAR images are extremely valuable as a base for field mapping in this heavily forested area, and the use of LiDAR is far superior to traditional mapping techniques relying only on aerial photography and 7.5 minute USGS quadrangle topographic maps. Comparison with earlier mapping of the northern San <span class="hlt">Andreas</span> fault (Brown and Wolfe, 1972) shows that in some areas the LiDAR data allow a correction of the fault trace</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AGUFM.T24B..04E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AGUFM.T24B..04E"><span>Observing the San <span class="hlt">Andreas</span> Fault at Depth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ellsworth, W.; Hickman, S.; Zoback, M.; Davis, E.; Gee, L.; Huggins, R.; Krug, R.; Lippus, C.; Malin, P.; Neuhauser, D.; Paulsson, B.; Shalev, E.; Vajapeyam, B.; Weiland, C.; Zumberge, M.</p> <p>2005-12-01</p> <p>Extending 4 km into the Earth along a diagonal path that crosses the divide between Salinian basement accreted to the Pacific Plate and Cretaceous sediments of North America, the main hole at the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) was designed to provide a portal into the inner workings of a major plate boundary fault. The successful drilling and casing of the main hole in the summer of 2005 to a total vertical depth of 3.1 km make it possible to conduct spatially extensive and long-duration observations of active tectonic processes within the actively deforming core of the San <span class="hlt">Andreas</span> Fault. In brief, the observatory consists of retrievable seismic, deformation and environmental sensors deployed inside the casing in both the main hole (maximum temperature 135 C) and the collocated pilot hole (1.1 km depth), and a fiber optic strainmeter installed behind casing in the main hole. By using retrievable systems deployed on either wire line or rigid tubing, each hole can be used for a wide range of scientific purposes, with instrumentation that takes maximum advantage of advances in sensor technology. To meet the scientific and technical challenges of building the observatory, borehole instrumentation systems developed for use in the petroleum industry and by the academic community in other deep research boreholes have been deployed in the SAFOD pilot hole and main hole over the past year. These systems included 15Hz omni-directional and 4.5 Hz gimbaled seismometers, micro-electro-mechanical accelerometers, tiltmeters, sigma-delta digitizers, and a fiber optic interferometeric strainmeter. A 1200-m-long, 3-component 80-level clamped seismic array was also operated in the main hole for 2 weeks of recording in May of 2005, collecting continuous seismic data at 4000 sps. Some of the observational highlights include capturing one of the M 2 SAFOD target repeating earthquakes in the near-field at a distance of 420 m, with accelerations of up to 200 cm/s and a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730021606','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730021606"><span>A simulation of the San <span class="hlt">Andreas</span> fault experiment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Agreen, R. W.; Smith, D. E.</p> <p>1973-01-01</p> <p>The San <span class="hlt">Andreas</span> Fault Experiment, which employs two laser tracking systems for measuring the relative motion of two points on opposite sides of the fault, was simulated for an eight year observation period. The two tracking stations are located near San Diego on the western side of the fault and near Quincy on the eastern side; they are roughly 900 kilometers apart. Both will simultaneously track laser reflector equipped satellites as they pass near the stations. Tracking of the Beacon Explorer C Spacecraft was simulated for these two stations during August and September for eight consecutive years. An error analysis of the recovery of the relative location of Quincy from the data was made, allowing for model errors in the mass of the earth, the gravity field, solar radiation pressure, atmospheric drag, errors in the position of the San Diego site, and laser systems range biases and noise. The results of this simulation indicate that the distance of Quincy from San Diego will be determined each year with a precision of about 10 centimeters. This figure is based on the accuracy of earth models and other parameters available in 1972.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014E%26PSL.399...14C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014E%26PSL.399...14C"><span>Lithosphere-asthenosphere interactions near the San <span class="hlt">Andreas</span> fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chamberlain, C. J.; Houlié, N.; Bentham, H. L. M.; Stern, T. A.</p> <p>2014-08-01</p> <p>We decipher the strain history of the upper mantle in California through the comparison of the long-term finite strain field in the mantle and the surface strain-rate field, respectively inferred from fast polarization directions of seismic phases (SKS and SKKS), and Global Positioning System (GPS) surface velocity fields. We show that mantle strain and surface strain-rate fields are consistent in the vicinity of San <span class="hlt">Andreas</span> Fault (SAF) in California. Such an agreement suggests that the lithosphere and strong asthenosphere have been deformed coherently and steadily since >1 Ma. We find that the crustal stress field rotates (up to 40° of rotation across a 50 km distance from 50° relative to the strike of the SAF, in the near-field of SAF) from San Francisco to the Central Valley. Both observations suggest that the SAF extends to depth, likely through the entire lithosphere. From Central Valley towards the Basin and Range, the orientations of GPS strain-rates, shear wave splitting measurements and seismic stress fields diverge indicating reduced coupling or/and shallow crustal extension and/or presence of frozen anisotropy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Litho.246...48G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Litho.246...48G"><span>Melt extraction and mantle source at a Southwest Indian Ridge Dragon Bone amagmatic segment on the <span class="hlt">Marion</span> Rise</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gao, Changgui; Dick, Henry J. B.; Liu, Yang; Zhou, Huaiyang</p> <p>2016-03-01</p> <p>This paper works on the trace and major element compositions of spatially associated basalts and peridotites from the Dragon Bone amagmatic ridge segment at the eastern flank of the <span class="hlt">Marion</span> Platform on the ultraslow spreading Southwest Indian Ridge. The rare earth element compositions of basalts do not match the pre-alteration Dragon Bone peridotite compositions, but can be modeled by about 5 to 10% non-modal batch equilibrium melting from a DMM source. The Dragon Bone peridotites are clinopyroxene-poor harzburgite with average spinel Cr# 27.7. The spinel Cr# indicates a moderate degree of melting. However, CaO and Al2O3 of the peridotites are lower than other abyssal peridotites at the same Mg# and extent of melting. This requires a pyroxene-poor initial mantle source composition compared to either hypothetical primitive upper mantle or depleted MORB mantle sources. We suggest a hydrous melting of the initial Dragon Bone mantle source, as wet melting depletes pyroxene faster than dry. According to the rare earth element patterns, the Dragon Bone peridotites are divided into two groups. Heavy REE in Group 1 are extremely fractionated from middle REE, which can be modeled by 7% fractional melting in the garnet stability field and another 12.5 to 13.5% in the spinel stability field from depleted and primitive upper mantle sources, respectively. Heavy REE in Group 2 are slightly fractionated from middle REE, which can be modeled by 15 to 20% fractional melting in the spinel stability field from a depleted mantle source. Both groups show similar melting degree to other abyssal peridotites. If all the melt extraction occurred at the middle oceanic ridge where the peridotites were dredged, a normal 6 km thick oceanic crust is expected at the Dragon Bone segment. However, the Dragon Bone peridotites are exposed in an amagmatic ridge segment where only scattered pillow basalts lie on a partially serpentinized mantle pavement. Thus their depletion requires an earlier melting</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/902293','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/902293"><span>Long-term slip rate of the southern San <span class="hlt">Andreas</span> Fault, from 10Be-26Al surface exposure dating of an offset alluvial fan</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>der Woerd, J v; Klinger, Y; Sieh, K</p> <p></p> <p>We determine the long-term slip rate of the southern San <span class="hlt">Andreas</span> Fault in the southeastern Indio Hills using {sup 10}Be and {sup 26}Al isotopes to date an offset alluvial fan surface. Field mapping complemented with topographic data, air photos and satellite images allow to precisely determine piercing points across the fault zone that are used to measure an offset of 565 {+-} 80 m. A total of twenty-six quartz-rich cobbles from three different fan surfaces were collected and dated. The tight cluster of nuclide concentrations from 19 samples out of 20 from the offset fan surface implies a simple exposuremore » history, negligible prior exposure and erosion, and yield an age of 35.5 {+-} 2.5 ka. The long-term slip rate of the San <span class="hlt">Andreas</span> Fault south of Biskra Palms is thus 15.9 {+-} 3.4 mm/yr. This rate is about 10 mm/yr slower than geological (0-14 ka) and short-term geodetic estimates for this part of the San <span class="hlt">Andreas</span> Fault implying changes in slip rate or in faulting behavior. This result puts new constraints on the slip rate of the San Jacinto and on the Eastern California Shear Zone for the last 35 ka. Our study shows that more sites along the major faults of southern California need to be targeted to better constrain the slip-rates over different time scales.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70034898','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70034898"><span>Basin geometry and cumulative offsets in the Eastern Transverse Ranges, southern California: Implications for transrotational deformation along the San <span class="hlt">Andreas</span> fault system</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Langenheim, V.E.; Powell, R.E.</p> <p>2009-01-01</p> <p>The Eastern Transverse Ranges, adjacent to and southeast of the big left bend of the San <span class="hlt">Andreas</span> fault, southern California, form a crustal block that has rotated clockwise in response to dextral shear within the San <span class="hlt">Andreas</span> system. Previous studies have indicated a discrepancy between the measured magnitudes of left slip on through-going east-striking fault zones of the Eastern Transverse Ranges and those predicted by simple geometric models using paleomagnetically determined clockwise rotations of basalts distributed along the faults. To assess the magnitude and source of this discrepancy, we apply new gravity and magnetic data in combination with geologic data to better constrain cumulative fault offsets and to define basin structure for the block between the Pinto Mountain and Chiriaco fault zones. Estimates of offset from using the length of pull-apart basins developed within left-stepping strands of the sinistral faults are consistent with those derived by matching offset magnetic anomalies and bedrock patterns, indicating a cumulative offset of at most ???40 km. The upper limit of displacements constrained by the geophysical and geologic data overlaps with the lower limit of those predicted at the 95% confidence level by models of conservative slip located on margins of rigid rotating blocks and the clockwise rotation of the paleomagnetic vectors. Any discrepancy is likely resolved by internal deformation within the blocks, such as intense deformation adjacent to the San <span class="hlt">Andreas</span> fault (that can account for the absence of basins there as predicted by rigid-block models) and linkage via subsidiary faults between the main faults. ?? 2009 Geological Society of America.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70037328','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70037328"><span>Uncertainties in slip-rate estimates for the Mission Creek strand of the southern San <span class="hlt">Andreas</span> fault at Biskra Palms Oasis, southern California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Behr, W.M.; Rood, D.H.; Fletcher, K.E.; Guzman, N.; Finkel, R.; Hanks, T.C.; Hudnut, K.W.; Kendrick, K.J.; Platt, J.P.; Sharp, W.D.; Weldon, R.J.; Yule, J.D.</p> <p>2010-01-01</p> <p>This study focuses on uncertainties in estimates of the geologic slip rate along the Mission Creek strand of the southern San <span class="hlt">Andreas</span> fault where it offsets an alluvial fan (T2) at Biskra Palms Oasis in southern California. We provide new estimates of the amount of fault offset of the T2 fan based on trench excavations and new cosmogenic 10Be age determinations from the tops of 12 boulders on the fan surface. We present three alternative fan offset models: a minimum, a maximum, and a preferred offset of 660 m, 980 m, and 770 m, respectively. We assign an age of between 45 and 54 ka to the T2 fan from the 10Be data, which is significantly older than previously reported but is consistent with both the degree of soil development associated with this surface, and with ages from U-series geochronology on pedogenic carbonate from T2, described in a companion paper by Fletcher et al. (this volume). These new constraints suggest a range of slip rates between ~12 and 22 mm/yr with a preferred estimate of ~14-17 mm/yr for the Mission Creek strand of the southern San <span class="hlt">Andreas</span> fault. Previous studies suggested that the geologic and geodetic slip-rate estimates at Biskra Palms differed. We find, however, that considerable uncertainty affects both the geologic and geodetic slip-rate estimates, such that if a real discrepancy between these rates exists for the southern San <span class="hlt">Andreas</span> fault at Biskra Palms, it cannot be demonstrated with available data. ?? 2010 Geological Society of America.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AGUFM.T23E..01Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AGUFM.T23E..01Z"><span>Overview of SAFOD Phases 1 and 2: Drilling, Sampling and Measurements in the San <span class="hlt">Andreas</span> Fault Zone at Seismogenic Depth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zoback, M. D.; Hickman, S.; Ellsworth, W.</p> <p>2005-12-01</p> <p>In this talk we provide an overview of on-site drilling, sampling and downhole measurement activities associated with the first two Phases of the San <span class="hlt">Andreas</span> Fault Observatory at Depth. SAFOD is located at the transition between the creeping and locked sections of the fault, 9 km NW of Parkfield, CA. A 2.1 km deep vertical pilot hole was drilled at the site in 2002. The SAFOD main borehole was drilled vertically to a depth of 1.5 km and then deviated at an average angle of 55° to vertical, passing beneath the surface trace of the San <span class="hlt">Andreas</span> fault, 1.8 km to the NW at a depth of 3.2 km. Repeating microearthquakes on the San <span class="hlt">Andreas</span> define the main active fault trace at depth, as well as a secondary active fault about 250 m to the SW (i.e., closer to SAFOD). The hole was rotary drilled, comprehensive cuttings were obtained and a real-time analysis of gases in the drilling mud was carried out. Spot cores were obtained at three depths (at casing set points) in the shallow granite and deeper sedimentary rocks penetrated by the hole, augmented by over fifty side-wall cores. Continuous coring of the San <span class="hlt">Andreas</span> Fault Zone will be carried out in Phase 3 of the project in the summer of 2007. In addition to sampling mud gas, discrete fluid and gas samples were obtained at several depths for geochemical analysis. Real-time geophysical measurements were made while drilling through most of the San <span class="hlt">Andreas</span> Fault Zone. A suite of "open hole" geophysical measurements were also made over essentially the entire depth of the hole. Construction of the multi-component SAFOD observatory is well underway, with a seismometer and tiltmeter operating at 1 km depth in the pilot hole and a fiber-optic laser strainmeter cemented behind casing in the main hole. A seismometer deployed at depth in the hole between Phases 1 and 2 detected one of the target earthquakes. A number of surface-to-borehole seismic experiments have been carried out to characterize seismic velocities and structures at</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018183','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018183"><span>A slow earthquake sequence on the San <span class="hlt">Andreas</span> fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Linde, A.T.; Gladwin, M.T.; Johnston, M.J.S.; Gwyther, R.L.; Bilham, R.G.</p> <p>1996-01-01</p> <p>EARTHQUAKES typically release stored strain energy on timescales of the order of seconds, limited by the velocity of sound in rock. Over the past 20 years, observations and laboratory experiments have indicated that capture can also occur more slowly, with durations up to hours. Such events may be important in earthquake nucleation and in accounting for the excess of plate convergence over seismic slip in subduction zones. The detection of events with larger timescales requires near-field deformation measurements. In December 1992, two borehole strainmeters close to the San <span class="hlt">Andreas</span> fault in California recorded a slow strain event of about a week in duration, and we show here that the strain changes were produced by a slow earthquake sequence (equivalent magnitude 4.8) with complexity similar to that of regular earthquakes. The largest earthquakes associated with these slow events were small (local magnitude 3.7) and contributed negligible strain release. The importance of slow earthquakes in the seismogenic process remains an open question, but these observations extend the observed timescale for slow events by two orders of magnitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010301','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010301"><span>Use of microearthquakes in the study of the mechanics of earthquake generation along the San <span class="hlt">Andreas</span> fault in central California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Eaton, J.P.; Lee, W.H.K.; Pakiser, L.C.</p> <p>1970-01-01</p> <p>A small, dense network of independently recording portable seismograph stations was used to delineate the slip surface associated with the 1966 Parkfield-Cholame earthquake by precise three dimensional mapping of the hypocenters of its aftershocks. The aftershocks were concentrated in a very narrow vertical zone beneath or immediately adjacent to the zone of surf ace fracturing that accompanied the main shock. Focal depths ranged from less than 1 km to a maximum of 15 km. The same type of portable network was used to study microearthquakes associated with an actively creeping section of the San <span class="hlt">Andreas</span> fault south of Hollister during the summer of 1967. Microearthquake activity during the 6-week operation of this network was dominated by aftershocks of a magnitude-4 earthquake that occurred within the network near Bear Valley on July 23. Most of the aftershocks were concentrated in an equidimensional region about 2 1 2km across that contained the hypocenter of the main shock. The zone of the concentrated aftershocks was centered near the middle of the rift zone at a depth of about 3 1 2km. Hypocenters of other aftershocks outlined a 25 km long zone of activity beneath the actively creeping strand of the fault and extending from the surface to a depth of about 13 km. A continuing study of microearthquakes along the San <span class="hlt">Andreas</span>, Hayward, and Calaveras faults between Hollister and San Francisco has been under way for about 2 years. The permanent telemetered network constructed for this purpose has grown from about 30 stations in early 1968 to about 45 stations in late 1969. Microearthquakes between Hollister and San Francisco are heavily concentrated in narrow, nearly vertical zones along sections of the Sargent, San <span class="hlt">Andreas</span>, and Calaveras faults. Focal depths range from less than 1 km to about 14 km. ?? 1970.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70042474','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70042474"><span>Correlation of clayey gouge in a surface exposure of the San <span class="hlt">Andreas</span> fault with gouge at depth from SAFOD: Implications for the role of serpentinite in fault mechanics</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Moore, Diane E.; Rymer, Michael J.</p> <p>2012-01-01</p> <p>Magnesium-rich clayey gouge similar to that comprising the two actively creeping strands of the San <span class="hlt">Andreas</span> Fault in drill core from the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) has been identified in a nearby outcrop of serpentinite within the fault zone at Nelson Creek. Each occurrence of the gouge consists of porphyroclasts of serpentinite and sedimentary rocks dispersed in a fine-grained, foliated matrix of Mg-rich smectitic clays. The clay minerals in all three gouges are interpreted to be the product of fluid-assisted, shear-enhanced reactions between quartzofeldspathic wall rocks and serpentinite that was tectonically entrained in the fault from a source in the Coast Range Ophiolite. We infer that the gouge at Nelson Creek connects to one or both of the gouge zones in the SAFOD core, and that similar gouge may occur at depths in between. The special significance of the outcrop is that it preserves the early stages of mineral reactions that are greatly advanced at depth, and it confirms the involvement of serpentinite and the Mg-rich phyllosilicate minerals that replace it in promoting creep along the central San <span class="hlt">Andreas</span> Fault.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70192478','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70192478"><span>Periodic, chaotic, and doubled earthquake recurrence intervals on the deep San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Shelly, David R.</p> <p>2010-01-01</p> <p>Earthquake recurrence histories may provide clues to the timing of future events, but long intervals between large events obscure full recurrence variability. In contrast, small earthquakes occur frequently, and recurrence intervals are quantifiable on a much shorter time scale. In this work, I examine an 8.5-year sequence of more than 900 recurring low-frequency earthquake bursts composing tremor beneath the San <span class="hlt">Andreas</span> fault near Parkfield, California. These events exhibit tightly clustered recurrence intervals that, at times, oscillate between ~3 and ~6 days, but the patterns sometimes change abruptly. Although the environments of large and low-frequency earthquakes are different, these observations suggest that similar complexity might underlie sequences of large earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24173678','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24173678"><span>The Renaissance and the universal surgeon: Giovanni <span class="hlt">Andrea</span> Della Croce, a master of traumatology.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Di Matteo, Berardo; Tarabella, Vittorio; Filardo, Giuseppe; Viganò, Anna; Tomba, Patrizia; Marcacci, Maurilio</p> <p>2013-12-01</p> <p>All the medical knowledge of all time in one book, the universal and perfect manual for the Renaissance surgeon, and the man who wrote it. This paper depicts the life and works of Giovanni <span class="hlt">Andrea</span> della Croce, a 16th Century physician and surgeon, who, endowed with true spirit of Renaissance humanism, wanted to teach and share all his medical knowledge through his opus magnum, titled "Universal Surgery Complete with All the Relevant Parts for the Optimum Surgeon". An extraordinary book which truly represents a defining moment and a founding stone for traumatology, written by a lesser known historical personality, but nonetheless the Renaissance Master of Traumatology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S41C0814B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S41C0814B"><span>Tidal Sensitivity of Declustered Low Frequency Earthquake Families and Inferred Creep Episodes on the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Babb, A.; Thomas, A.; Bletery, Q.</p> <p>2017-12-01</p> <p>Low frequency earthquakes (LFEs) are detected at depths of 16-30 km on a 150 km section of the San <span class="hlt">Andreas</span> Fault centered at Parkfield, CA. The LFEs are divided into 88 families based on waveform similarity. Each family is thought to represent a brittle asperity on the fault surface that repeatedly slips during aseismic slip of the surrounding fault. LFE occurrence is irregular which allows families to be divided into continuous and episodic. In continuous families a burst of a few LFE events recurs every few days while episodic families experience essentially quiescent periods often lasting months followed by bursts of hundreds of events over a few days. The occurrence of LFEs has also been shown to be sensitive to extremely small ( 1kPa) tidal stress perturbations. However, the clustered nature of LFE occurrence could potentially bias estimates of tidal sensitivity. Here we re-evaluate the tidal sensitivity of LFE families on the deep San <span class="hlt">Andreas</span> using a declustered catalog. In this catalog LFE bursts are isolated based on the recurrence intervals between individual LFE events for each family. Preliminary analysis suggests that declustered LFE families are still highly sensitive to tidal stress perturbations, primarily right-lateral shear stress (RLSS) and to a lesser extent fault normal stress (FNS). We also find inferred creep episodes initiate preferentially during times of positive RLSS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2003/0450/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2003/0450/"><span>Photomosaics and logs of trenches on the San <span class="hlt">Andreas</span> Fault at Arano Flat near Watsonville, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fumal, Thomas E.; Heingartner, Gordon F.; Samrad, Laura; Dawson, Timothy E.; Hamilton, John C.; Baldwin, John N.</p> <p>2004-01-01</p> <p>We present photomosaics and logs of the walls of trenches excavated for a paleoseismic study at Arano Flat, one of two sites along the San <span class="hlt">Andreas</span> fault in the Santa Cruz Mountains on the Kelley-Thompson Ranch. At this location, the fault consists of a narrow zone along the northeast side of a low ridge adjacent to a possible sag pond and extends about 60-70 meters across a broad alluvial flat. This site was a part of Rancho Salsipuedes beginning in 1834 and was purchased by the present owner’s family in 1851.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26307412','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26307412"><span>Centennial ties: Harvey Cushing (1869-1939) and William Osler (1849-1919) on <span class="hlt">Andreas</span> Vesalius (1514-1564).</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Toodayan, Nadeem</p> <p>2017-08-01</p> <p><span class="hlt">Andreas</span> Vesalius is often regarded as the founding father of modern anatomical study. The quincentennial anniversary of his birth - 31 December 2014 - has been very widely commemorated, and it is the purpose of this article to contrast these celebrations with what happened during the Vesalius quatercentenary year of 1914. More specifically, we look at how Vesalius was perceived a century ago by examining his influence on two of western medicine's most iconic gentlemen - Harvey Williams Cushing (1869-1939) and his larger than life mentor, Sir William Osler (1849-1919).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.G43A0845A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.G43A0845A"><span>Investigating the creeping section of the San <span class="hlt">Andreas</span> Fault using ALOS PALSAR interferometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Agram, P. S.; Wortham, C.; Zebker, H. A.</p> <p>2010-12-01</p> <p>In recent years, time-series InSAR techniques have been used to study the temporal characteristics of various geophysical phenomena that produce surface deformation including earthquakes and magma migration in volcanoes. Conventional InSAR and time-series InSAR techniques have also been successfully used to study aseismic creep across faults in urban areas like the Northern Hayward Fault in California [1-3]. However, application of these methods to studying the time-dependent creep across the Central San <span class="hlt">Andreas</span> Fault using C-band ERS and Envisat radar satellites has resulted in limited success. While these techniques estimate the average long-term far-field deformation rates reliably, creep measurement close to the fault (< 3-4 Km) is virtually impossible due to heavy decorrelation at C-band (6cm wavelength). Shanker and Zebker (2009) [4] used the Persistent Scatterer (PS) time-series InSAR technique to estimate a time-dependent non-uniform creep signal across a section of the creeping segment of the San <span class="hlt">Andreas</span> Fault. However, the identified PS network was spatially very sparse (1 per sq. km) to study temporal characteristics of deformation of areas close to the fault. In this work, we use L-band (24cm wavelength) SAR data from the PALSAR instrument on-board the ALOS satellite, launched by Japanese Aerospace Exploration Agency (JAXA) in 2006, to study the temporal characteristics of creep across the Central San <span class="hlt">Andreas</span> Fault. The longer wavelength at L-band improves observed correlation over the entire scene which significantly increased the ground area coverage of estimated deformation in each interferogram but at the cost of decreased sensitivity of interferometric phase to surface deformation. However, noise levels in our deformation estimates can be decreased by combining information from multiple SAR acquisitions using time-series InSAR techniques. We analyze 13 SAR acquisitions spanning the time-period from March 2007 to Dec 2009 using the Short Baseline</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01644&hterms=red+tide&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dred%2Btide','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01644&hterms=red+tide&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dred%2Btide"><span>San <span class="hlt">Andreas</span>-sized Strike-slip Fault on Europa</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1998-01-01</p> <p><p/>This mosaic of the south polar region of Jupiter's moon Europa shows the northern 290 kilometers (180 miles) of a strike-slip fault named Astypalaea Linea. The entire fault is about 810 kilometers (500 miles) long, about the size of the California portion of the San <span class="hlt">Andreas</span> fault, which runs from the California-Mexico border north to the San Francisco Bay. <p/>In a strike-slip fault, two crustal blocks move horizontally past one another, similar to two opposing lanes of traffic. Overall motion along the fault seems to have followed a continuous narrow crack along the feature's entire length, with a path resembling steps on a staircase crossing zones that have been pulled apart. The images show that about 50 kilometers (30 miles) of displacement have taken place along the fault. The fault's opposite sides can be reconstructed like a puzzle, matching the shape of the sides and older, individual cracks and ridges broken by its movements. <p/> [figure removed for brevity, see original site] <p/>The red line marks the once active central crack of the fault. The black line outlines the fault zone, including material accumulated in the regions which have been pulled apart. <p/>Bends in the fault have allowed the surface to be pulled apart. This process created openings through which warmer, softer ice from below Europa's brittle ice shell surface, or frozen water from a possible subsurface ocean, could reach the surface. This upwelling of material formed large areas of new ice within the boundaries of the original fault. A similar pulling-apart phenomenon can be observed in the geological trough surrounding California's Salton Sea, in Death Valley and the Dead Sea. In those cases, the pulled-apart regions can include upwelled materials, but may be filled mostly by sedimentary and eroded material from above. <p/>One theory is that fault motion on Europa is induced by the pull of variable daily tides generated by Jupiter's gravitational tug on Europa. Tidal tension</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/833997','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/833997"><span>Vibroseis Monitoring of San <span class="hlt">Andreas</span> Fault in California</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Korneev, Valeri; Nadeau, Robert</p> <p>2004-06-11</p> <p>A unique data set of seismograms for 720 source-receiver paths has been collected as part of a controlled source Vibroseis experiment San <span class="hlt">Andreas</span> Fault (SAF) at Parkfield. In the experiment, seismic waves repeatedly illuminated the epicentral region of the expected M6 event at Parkfield from June 1987 until November 1996. For this effort, a large shear-wave vibrator was interfaced with the 3-component (3-C) borehole High-Resolution Seismic Network (HRSN), providing precisely timed collection of data for detailed studies of changes in wave propagation associated with stress and strain accumulation in the fault zone (FZ). Data collected by the borehole network weremore » examined for evidence of changes associated with the nucleation process of the anticipated M6 earthquake at Parkfield. These investigations reported significant traveltime changes in the S coda for paths crossing the fault zone southeast of the epicenter and above the rupture zone of the 1966 M6 earthquake. Analysis and modeling of these data and comparison with observed changes in creep, water level, microseismicity, slip-at-depth and propagation from characteristic repeating microearthquakes showed temporal variations in a variety of wave propagation attributes that were synchronous with changes in deformation and local seismicity patterns. Numerical modeling suggests 200 meters as an effective thickness of SAF. The observed variations can be explained by velocity 6 percent velocity variation within SAF core. Numerical modeling studies and a growing number of observations have argued for the propagation of fault-zone guided waves (FZGW) within a SAF zone that is 100 to 200 m wide at seismogenic depths and with 20 to 40 percent lower shear-wave velocity than the adjacent unfaulted rock. Guided wave amplitude tomographic inversion for SAF using microearthquakes, shows clearly that FZGW are significantly less attenuated in a well-defined region of the FZ. This region plunges to the northwest along</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.5996M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.5996M"><span>University of the seas, 15 years of oceanographic schools on board of the <span class="hlt">Marion</span> Dufresne</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malaize, Bruno; Deverchere, Jacques; Leau, Hélène; Graindorge, David</p> <p>2015-04-01</p> <p>Since the first University at Sea, proposed by two French Universities (Brest and Bordeaux) in 1999, the R/V <span class="hlt">Marion</span> Dufresne, in collaboration with the French Polar institute (IPEV), has welcome 12 oceanographic schools. The main objective of this educational and scientific program is to stimulate the potential interest of highly graduated students in scientific fields dealing with oceanography, and to broaden exchanges with foreign universities, strengthening a pool of excellence at a high international scientific level. It is a unique opportunity for the students to discover and to be involved in the work in progress of collecting scientific data on board of a ship, and to attend international research courses given by scientists involved in the cruise program. They also experience the final task of the scientific work by presenting their own training results, making posters on board, and writing a cruise report. For some University at Sea, students had also updated a daily journal, available on internet, hosted by the main institutions involved (as IPEV or EPOC, Bordeaux University). All this work is done in English, a common language to all the participants. An overview of these 15 years background experience will be presented, underlying the financial supports used, the logistic on board, as well as all the benefits acquiered by all former students, now in permanent positions in different international institutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018276','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018276"><span>A 100-year average recurrence interval for the San <span class="hlt">Andreas</span> fault at Wrightwood, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fumal, T.E.; Pezzopane, S.K.; Weldon, R.J.; Schwartz, D.P.</p> <p>1993-01-01</p> <p>Evidence for five large earthquakes during the past five centuries along the San <span class="hlt">Andreas</span> fault zone 70 kilometers northeast of Los Angeles, California, indicates that the average recurrence interval and the temporal variability are significantly smaller than previously thought. Rapid sedimentation during the past 5000 years in a 150-meter-wide structural depression has produced a greater than 21-meter-thick sequence of debris flow and stream deposits interbedded with more than 50 datable peat layers. Fault scarps, colluvial wedges, fissure infills, upward termination of ruptures, and tilted and folded deposits above listric faults provide evidence for large earthquakes that occurred in A.D. 1857, 1812, and about 1700, 1610, and 1470.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70101407','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70101407"><span>Southern San <span class="hlt">Andreas</span> Fault evaluation field activity: approaches to measuring small geomorphic offsets--challenges and recommendations for active fault studies</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Scharer, Katherine M.; Salisbury, J. Barrett; Arrowsmith, J. Ramon; Rockwell, Thomas K.</p> <p>2014-01-01</p> <p>In southern California, where fast slip rates and sparse vegetation contribute to crisp expression of faults and microtopography, field and high‐resolution topographic data (<1  m/pixel) increasingly are used to investigate the mark left by large earthquakes on the landscape (e.g., Zielke et al., 2010; Zielke et al., 2012; Salisbury, Rockwell, et al., 2012, Madden et al., 2013). These studies measure offset streams or other geomorphic features along a stretch of a fault, analyze the offset values for concentrations or trends along strike, and infer that the common magnitudes reflect successive surface‐rupturing earthquakes along that fault section. Wallace (1968) introduced the use of such offsets, and the challenges in interpreting their “unique complex history” with offsets on the Carrizo section of the San <span class="hlt">Andreas</span> fault; these were more fully mapped by Sieh (1978) and followed by similar field studies along other faults (e.g., Lindvall et al., 1989; McGill and Sieh, 1991). Results from such compilations spurred the development of classic fault behavior models, notably the characteristic earthquake and slip‐patch models, and thus constitute an important component of the long‐standing contrast between magnitude–frequency models (Schwartz and Coppersmith, 1984; Sieh, 1996; Hecker et al., 2013). The proliferation of offset datasets has led earthquake geologists to examine the methods and approaches for measuring these offsets, uncertainties associated with measurement of such features, and quality ranking schemes (Arrowsmith and Rockwell, 2012; Salisbury, Arrowsmith, et al., 2012; Gold et al., 2013; Madden et al., 2013). In light of this, the Southern San <span class="hlt">Andreas</span> Fault Evaluation (SoSAFE) project at the Southern California Earthquake Center (SCEC) organized a combined field activity and workshop (the “Fieldshop”) to measure offsets, compare techniques, and explore differences in interpretation. A thorough analysis of the measurements from the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G34A..05X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G34A..05X"><span>Crustal Deformation along San <span class="hlt">Andreas</span> Fault System revealed by GPS and Sentinel-1 InSAR</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, X.; Sandwell, D. T.</p> <p>2017-12-01</p> <p>We present a crustal deformation velocity map along the San <span class="hlt">Andreas</span> Fault System by combining measurements from Sentinel-1 Interferometric Synthetic Aperture Radar (InSAR) and Global Positioning System (GPS) velocity models (CGM V1). We assembled 5 tracks of descending Sentinel-1 InSAR data spanning 2014.11-2017.02, and produced 545 interferograms, each of which covers roughly 250km x 420km area ( 60 bursts). These interferograms are unwrapped using SNAPHU [Chen & Zebker, 2002], with the 2Npi unwrapping ambiguity corrected with a sparse recovery method. We used coherence-based small baseline subset (SBAS) method [Tong & Schmidt, 2016] together with atmospheric correction by common-point stacking [Tymofyeyeva and Fialko, 2015] to construct deformation time series [Xu et. al., 2017]. Then we project the horizontal GPS model and vertical GPS data into satellite line-of-sight directions separately. We first remove the horizontal GPS model from InSAR measurements and perform elevation-dependent atmospheric phase correction. Then we compute the discrepancy between the remaining InSAR measurements and vertical GPS data. We interpolate this discrepancy and remove it from the residual InSAR measurements. Finally, we restore the horizontal GPS model. Preliminary results show that fault creep over the San Jacinto fault, the Elsinore fault, and the San <span class="hlt">Andreas</span> creeping section is clearly resolved. During the period of drought, the Central Valley of California was subsiding at a high rate (up to 40 cm/yr), while the city of San Jose is uplifting due to recharge, with a quaternary fault acting as a ground water barrier. These findings will be reported during the meeting.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70015405','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70015405"><span>Historic surface slip along the San <span class="hlt">Andreas</span> Fault near Parkfield, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lienkaemper, J.J.; Prescott, W.H.</p> <p>1989-01-01</p> <p>The Parkfield Earthquake Prediction Experiment is focusing close attention on the 44-km-long section of the San <span class="hlt">Andreas</span> fault that last ruptured seismically in 1966 (Ms 6.0). The 20-km-long central segment of the 1966 Parkfield rupture, extending from the mainshock epicenter at Middle Mountain southeastward to Gold Hill, forms a 1- to 2-km salient northeastward away from the dominant N40??W strike. Following the 1966 earthquake afterslip, aseismic slip has been nearly constant. Moderate Parkfield earthquakes have recurred on average every 21 years since 1857, when a great earthquake (M ~ 8) ruptured at least as far north as the southern Parkfield segment. Many measurements of slip have been made near Parkfield since 1966. Nevertheless, much of the history of surface slip remained uncertain, especially the total amount associated with the 1966 event. In 1985 we measured accumulated slip on the four oldest cultural features offset by the fault along the 1966 Parkfield rupture segment. -from Authors</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70182765','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70182765"><span>Low resistivity and permeability in actively deforming shear zones on the San <span class="hlt">Andreas</span> Fault at SAFOD</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Morrow, Carolyn A.; Lockner, David A.; Hickman, Stephen H.</p> <p>2015-01-01</p> <p>The San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) scientific drillhole near Parkfield, California crosses the San <span class="hlt">Andreas</span> Fault at a depth of 2.7 km. Downhole measurements and analysis of core retrieved from Phase 3 drilling reveal two narrow, actively deforming zones of smectite-clay gouge within a roughly 200 m-wide fault damage zone of sandstones, siltstones and mudstones. Here we report electrical resistivity and permeability measurements on core samples from all of these structural units at effective confining pressures up to 120 MPa. Electrical resistivity (~10 ohm-m) and permeability (10-21 to 10-22 m2) in the actively deforming zones were one to two orders of magnitude lower than the surrounding damage zone material, consistent with broader-scale observations from the downhole resistivity and seismic velocity logs. The higher porosity of the clay gouge, 2 to 8 times greater than that in the damage zone rocks, along with surface conduction were the principal factors contributing to the observed low resistivities. The high percentage of fine-grained clay in the deforming zones also greatly reduced permeability to values low enough to create a barrier to fluid flow across the fault. Together, resistivity and permeability data can be used to assess the hydrogeologic characteristics of the fault, key to understanding fault structure and strength. The low resistivities and strength measurements of the SAFOD core are consistent with observations of low resistivity clays that are often found in the principal slip zones of other active faults making resistivity logs a valuable tool for identifying these zones.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/1990/0159/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/1990/0159/report.pdf"><span>Water-resources programs and hydrologic-information needs, <span class="hlt">Marion</span> County, Indiana, 1987</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Duwelius, R.F.</p> <p>1990-01-01</p> <p>Water resources are abundant in <span class="hlt">Marion</span> County, Indiana, and have been developed for public and industrial supply, energy generation, irrigation, and recreation. The largest water withdrawals are from surface water, and the two largest water uses are public supply and cooling water for electrical-generating plants. Water-resources programs in the county are carried out by Federal, State and local agencies to address issues of surface and groundwater availability and quality. The programs of each agency are related to the functions and goals of the agency. Although each agency has specific information needs to fulfill its functions, sometimes these needs overlap, and there are times when the same hydrologic information benefits all. Overlapping information needs and activities create opportunities for interagency coordination and cooperation. Such cooperation could lead to a savings of dollars spent on water-resources programs and could assure an improved understanding of the water resources of the county. Representatives from four agencies-- the Indiana Department of Environmental Management, the Indiana Department of Natural Resources, the Indianapolis Department of Public Works, and the U.S. Geological Survey--met four times in 1987 to describe their own water-resources programs, to identify hydrologic-information needs, and to contact other agencies with related programs. This report presents the interagency findings and is intended to further communication among water resource agencies by identifying current programs and common needs for hydrologic information. Hydrologic information needs identified by the agency representatives include more precise methods for determining the volume of water withdrawals and for determining the volume of industrial and municipal discharges to surface water. Maps of flood-prone areas need to be updated as more of the county is developed. Improved aquifer maps of the inter-till aquifers are needed, and additional observation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70011574','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70011574"><span>Observations of strain accumulation across the San <span class="hlt">Andreas</span> fault near Palmdale, California, with a two-color geodimeter</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Langbein, J.O.; Linker, M.F.; McGarr, A.; Slater, L.E.</p> <p>1982-01-01</p> <p>Two-color laser ranging measurements during a 15-month period over a geodetic network spanning the San <span class="hlt">Andreas</span> fault near Palmdale, California, indicate that the crust expands and contracts aseismically in episodes as short as 2 weeks. Shear strain parallel to the fault has accumulated monotonically since November 1980, but at a variable rate. Improvements in measurement precision and temporal resolution over those of previous geodetic studies near Palmdale have resulted in the definition of a time history of crustal deformation that is much more complex than formerly realized. Copyright ?? 1982 AAAS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880045604&hterms=sauber&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsauber','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880045604&hterms=sauber&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsauber"><span>Geodetic measurement of deformation east of the San <span class="hlt">Andreas</span> Fault in Central California</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sauber, Jeanne; Solomon, Sean C.; Lisowski, Michael</p> <p>1988-01-01</p> <p>The shear strain rates in the Diablo Range of California have been calculated, and the slip rate along the Calaveras and Paicines faults in Central California have been estimated, on the basis of triangulation and trilateration data from two geodetic networks located between the western edge of the Great Valley and the San <span class="hlt">Andreas</span> Fault. The orientation of the principal compressive strain predicted from the azimuth of the major structures in the region is N 25 deg E, leading to an average shear strain value that corresponds to a relative shortening rate of 4.5 + or - 2.4 mm/yr. It is inferred that the measured strain is due to compression across the fold of this area. The hypothesized uniform, fault-normal compression within the Coast Ranges is not supported by these results.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45.3967J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45.3967J"><span>An Anisotropic Contrast in the Lithosphere Across the Central San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jiang, Chengxin; Schmandt, Brandon; Clayton, Robert W.</p> <p>2018-05-01</p> <p>Seismic anisotropy of the lithosphere and asthenosphere was investigated with a dense broadband seismic transect nearly orthogonal to the central San <span class="hlt">Andreas</span> fault (SAF). A contrast in SK(K)S splitting was found across the SAF, with a clockwise rotation of the fast orientation 26° closer to the strike of the SAF and greater delay times for stations located within 35 km to the east. Dense seismograph spacing requires heterogeneous anisotropy east of the SAF in the uppermost mantle or crust. Based on existing station coverage, such a contrast in splitting orientations across the SAF may be unusual along strike and its location coincides with the high-velocity Isabella anomaly in the upper mantle. If the Isabella anomaly is a fossil slab fragment translating with the Pacific plate, the anomalous splitting east of the SAF could indicate a zone of margin-parallel shear beneath the western edge of North America.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/2198714','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/2198714"><span>[Christian <span class="hlt">Andreas</span> Cothenius (1708-1789). A pro-memoria on the occasion of the 200th anniversary of his death].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Völker, A</p> <p>1990-04-01</p> <p>The 200th anniversary of the death of Christian <span class="hlt">Andreas</span> Cothenius gave occasion to appreciate life and work of this personage of a physician. Cothenius maintained manifold connections to Halle, of which the golden doctorate and the heritage of the pharmaceutic enterprises of his teacher Friedrich Hoffmann were treated in this place. The picture of the local relations was supplemented by the history of the Cothenius medal which is today awarded by the Leopoldina of Halle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29662122','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29662122"><span>Shallow deformation of the San <span class="hlt">Andreas</span> fault 5 years following the 2004 Parkfield earthquake (Mw6) combining ERS2 and Envisat InSAR.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bacques, Guillaume; de Michele, Marcello; Raucoules, Daniel; Aochi, Hideo; Rolandone, Frédérique</p> <p>2018-04-16</p> <p>This study focuses on the shallow deformation that occurred during the 5 years following the Parkfield earthquake (28/09/2004, Mw 6, San <span class="hlt">Andreas</span> Fault, California). We use Synthetic Aperture Radar interferometry (InSAR) to provide precise measurements of transient deformations after the Parkfield earthquake between 2005 and 2010. We propose a method to combine both ERS2 and ENVISAT interferograms to increase the temporal data sampling. Firstly, we combine 5 years of available Synthetic Aperture Radar (SAR) acquisitions including both ERS-2 and Envisat. Secondly, we stack selected interferograms (both from ERS2 and Envisat) for measuring the temporal evolution of the ground velocities at given time intervals. Thanks to its high spatial resolution, InSAR could provide new insights on the surface fault motion behavior over the 5 years following the Parkfield earthquake. As a complement to previous studies in this area, our results suggest that shallow transient deformations affected the Creeping-Parkfield-Cholame sections of the San <span class="hlt">Andreas</span> Fault after the 2004 Mw6 Parkfield earthquake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70030111','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70030111"><span>Structure of the California Coast Ranges and San <span class="hlt">Andreas</span> Fault at SAFOD from seismic waveform inversion and reflection imaging</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Bleibinhaus, F.; Hole, J.A.; Ryberg, T.; Fuis, G.S.</p> <p>2007-01-01</p> <p>A seismic reflection and refraction survey across the San <span class="hlt">Andreas</span> Fault (SAF) near Parkfield provides a detailed characterization of crustal structure across the location of the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD). Steep-dip prestack migration and frequency domain acoustic waveform tomography were applied to obtain highly resolved images of the upper 5 km of the crust for 15 km on either side of the SAF. The resulting velocity model constrains the top of the Salinian granite with great detail. Steep-dip reflection seismic images show several strong-amplitude vertical reflectors in the uppermost crust near SAFOD that define an ???2-km-wide zone comprising the main SAF and two or more local faults. Another prominent subvertical reflector at 2-4 km depth ???9 km to the northeast of the SAF marks the boundary between the Franciscan terrane and the Great Valley Sequence. A deep seismic section of low resolution shows several reflectors in the Salinian crust west of the SAF. Two horizontal reflectors around 10 km depth correlate with strains of seismicity observed along-strike of the SAF. They represent midcrustal shear zones partially decoupling the ductile lower crust from the brittle upper crust. The deepest reflections from ???25 km depth are interpreted as crust-mantle boundary. Copyright 2007 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030106079&hterms=InSAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DInSAR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030106079&hterms=InSAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DInSAR"><span>Fault Creep along the Southern San <span class="hlt">Andreas</span> from Interferometric Synthetic Aperture Radar, Permanent Scatterers, and Stacking</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lyons, Suzanne; Sandwell, David</p> <p>2003-01-01</p> <p>Interferometric synthetic aperture radar (InSAR) provides a practical means of mapping creep along major strike-slip faults. The small amplitude of the creep signal (less than 10 mm/yr), combined with its short wavelength, makes it difficult to extract from long time span interferograms, especially in agricultural or heavily vegetated areas. We utilize two approaches to extract the fault creep signal from 37 ERS SAR images along the southem San <span class="hlt">Andreas</span> Fault. First, amplitude stacking is utilized to identify permanent scatterers, which are then used to weight the interferogram prior to spatial filtering. This weighting improves correlation and also provides a mask for poorly correlated areas. Second, the unwrapped phase is stacked to reduce tropospheric and other short-wavelength noise. This combined processing enables us to recover the near-field (approximately 200 m) slip signal across the fault due to shallow creep. Displacement maps fiom 60 interferograms reveal a diffuse secular strain buildup, punctuated by localized interseismic creep of 4-6 mm/yr line of sight (LOS, 12-18 mm/yr horizontal). With the exception of Durmid Hill, this entire segment of the southern San <span class="hlt">Andreas</span> experienced right-lateral triggered slip of up to 10 cm during the 3.5-year period spanning the 1992 Landers earthquake. The deformation change following the 1999 Hector Mine earthquake was much smaller (4 cm) and broader than for the Landers event. Profiles across the fault during the interseismic phase show peak-to-trough amplitude ranging from 15 to 25 mm/yr (horizontal component) and the minimum misfit models show a range of creeping/locking depth values that fit the data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70189603','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70189603"><span>A large mantle water source for the northern San <span class="hlt">Andreas</span> Fault System: A ghost of subduction past</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kirby, Stephen H.; Wang, Kelin; Brocher, Thomas M.</p> <p>2014-01-01</p> <p>Recent research indicates that the shallow mantle of the Cascadia subduction margin under near-coastal Pacific Northwest U.S. is cold and partially serpentinized, storing large quantities of water in this wedge-shaped region. Such a wedge probably formed to the south in California during an earlier period of subduction. We show by numerical modeling that after subduction ceased with the creation of the San <span class="hlt">Andreas</span> Fault System (SAFS), the mantle wedge warmed, slowly releasing its water over a period of more than 25 Ma by serpentine dehydration into the crust above. This deep, long-term water source could facilitate fault slip in San <span class="hlt">Andreas</span> System at low shear stresses by raising pore pressures in a broad region above the wedge. Moreover, the location and breadth of the water release from this model gives insights into the position and breadth of the SAFS. Such a mantle source of water also likely plays a role in the occurrence of Non-Volcanic Tremor (NVT) that has been reported along the SAFS in central California. This process of water release from mantle depths could also mobilize mantle serpentinite from the wedge above the dehydration front, permitting upward emplacement of serpentinite bodies by faulting or by diapiric ascent. Specimens of serpentinite collected from tectonically emplaced serpentinite blocks along the SAFS show mineralogical and structural evidence of high fluid pressures during ascent from depth. Serpentinite dehydration may also lead to tectonic mobility along other plate boundaries that succeed subduction, such as other continental transforms, collision zones, or along present-day subduction zones where spreading centers are subducting.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2003/0469/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2003/0469/"><span>Photomosaics and logs of trenches on the San <span class="hlt">Andreas</span> Fault at Mill Canyon near Watsonville, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fumal, Thomas E.; Dawson, Timothy E.; Flowers, Rebecca; Hamilton, John C.; Heingartner, Gordon F.; Kessler, James; Samrad, Laura</p> <p>2004-01-01</p> <p>We present photomosaics and logs of the walls of trenches excavated for a paleoseismic study at Mill Canyon, one of two sites along the San <span class="hlt">Andreas</span> fault in the Santa Cruz Mtns. on the Kelley-Thompson Ranch. This site was a part of Rancho Salsipuedes begining in 1834. It was purchased by the present owner’s family in 1851. Remnants of a cabin/mill operations still exist up the canyon dating from 1908 when the area was logged. At this location, faulting has moved a shutter ridge across the mouth of Mill Canyon ponding Holocene sediment. Recent faulting is confined to a narrow zone near the break in slope.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1341854-tidal-triggering-earthquakes-suggests-poroelastic-behavior-san-andreas-fault','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1341854-tidal-triggering-earthquakes-suggests-poroelastic-behavior-san-andreas-fault"><span>Tidal triggering of earthquakes suggests poroelastic behavior on the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Delorey, Andrew A.; van der Elst, Nicholas J.; Johnson, Paul Allan</p> <p>2016-12-28</p> <p>Tidal triggering of earthquakes is hypothesized to provide quantitative information regarding the fault's stress state, poroelastic properties, and may be significant for our understanding of seismic hazard. To date, studies of regional or global earthquake catalogs have had only modest successes in identifying tidal triggering. We posit that the smallest events that may provide additional evidence of triggering go unidentified and thus we developed a technique to improve the identification of very small magnitude events. We identify events applying a method known as inter-station seismic coherence where we prioritize detection and discrimination over characterization. Here we show tidal triggering ofmore » earthquakes on the San <span class="hlt">Andreas</span> Fault. We find the complex interaction of semi-diurnal and fortnightly tidal periods exposes both stress threshold and critical state behavior. Lastly, our findings reveal earthquake nucleation processes and pore pressure conditions – properties of faults that are difficult to measure, yet extremely important for characterizing earthquake physics and seismic hazards.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1341854','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1341854"><span>Tidal triggering of earthquakes suggests poroelastic behavior on the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Delorey, Andrew A.; van der Elst, Nicholas J.; Johnson, Paul Allan</p> <p></p> <p>Tidal triggering of earthquakes is hypothesized to provide quantitative information regarding the fault's stress state, poroelastic properties, and may be significant for our understanding of seismic hazard. To date, studies of regional or global earthquake catalogs have had only modest successes in identifying tidal triggering. We posit that the smallest events that may provide additional evidence of triggering go unidentified and thus we developed a technique to improve the identification of very small magnitude events. We identify events applying a method known as inter-station seismic coherence where we prioritize detection and discrimination over characterization. Here we show tidal triggering ofmore » earthquakes on the San <span class="hlt">Andreas</span> Fault. We find the complex interaction of semi-diurnal and fortnightly tidal periods exposes both stress threshold and critical state behavior. Lastly, our findings reveal earthquake nucleation processes and pore pressure conditions – properties of faults that are difficult to measure, yet extremely important for characterizing earthquake physics and seismic hazards.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70192213','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70192213"><span>Tidal triggering of earthquakes suggests poroelastic behavior on the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Delorey, Andrew; Van Der Elst, Nicholas; Johnson, Paul</p> <p>2017-01-01</p> <p>Tidal triggering of earthquakes is hypothesized to provide quantitative information regarding the fault's stress state, poroelastic properties, and may be significant for our understanding of seismic hazard. To date, studies of regional or global earthquake catalogs have had only modest successes in identifying tidal triggering. We posit that the smallest events that may provide additional evidence of triggering go unidentified and thus we developed a technique to improve the identification of very small magnitude events. We identify events applying a method known as inter-station seismic coherence where we prioritize detection and discrimination over characterization. Here we show tidal triggering of earthquakes on the San <span class="hlt">Andreas</span> Fault. We find the complex interaction of semi-diurnal and fortnightly tidal periods exposes both stress threshold and critical state behavior. Our findings reveal earthquake nucleation processes and pore pressure conditions – properties of faults that are difficult to measure, yet extremely important for characterizing earthquake physics and seismic hazards.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://mo.water.usgs.gov/Reports/98-4211-Southard/index.htm','USGSPUBS'); return false;" href="http://mo.water.usgs.gov/Reports/98-4211-Southard/index.htm"><span>Base (100-year) flood elevations for selected sites in <span class="hlt">Marion</span> County, Missouri</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Southard, Rodney E.; Wilson, Gary L.</p> <p>1998-01-01</p> <p>The primary requirement for community participation in the National Flood Insurance Program is the adoption and enforcement of floodplain management requirements that minimize the potential for flood damages to new construction and avoid aggravating existing flooding conditions. This report provides base flood elevations (BFE) for a 100-year recurrence flood for use in the management and regulation of 14 flood-hazard areas designated by the Federal Emergency Management Agency as approximate Zone A areas in <span class="hlt">Marion</span> County, Missouri. The one-dimensional surface-water flow model, HEC-RAS, was used to compute the base (100-year) flood elevations for the 14 Zone A sites. The 14 sites were located at U.S., State, or County road crossings and the base flood elevation was determined at the upstream side of each crossing. The base (100-year) flood elevations for BFE 1, 2, and 3 on the South Fork North River near Monroe City, Missouri, are 627.7, 579.2, and 545.9 feet above sea level. The base (100-year) flood elevations for BFE 4, 5, 6, and 7 on the main stem of the North River near or at Philadelphia and Palmyra, Missouri, are 560.5, 539.7, 504.2, and 494.4 feet above sea level. BFE 8 is located on Big Branch near Philadelphia, a tributary to the North River, and the base (100-year) flood elevation at this site is 530.5 feet above sea level. One site (BFE 9) is located on the South River near Monroe City, Missouri. The base (100-year) flood elevation at this site is 619.1 feet above sea level. Site BFE 10 is located on Bear Creek near Hannibal, Missouri, and the base (100-year) elevation is 565.5 feet above sea level. The four remaining sites (BFE 11, 12, 13, and 14) are located on the South Fabius River near Philadelphia and Palmyra, Missouri. The base (100-year) flood elevations for BFE 11, 12, 13, and 14 are 591.2, 578.4, 538.7, and 506.9 feet above sea level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22039973','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22039973"><span>Carbonate organo-mineral micro- and ultrastructures in sub-fossil stromatolites: <span class="hlt">Marion</span> lake, South Australia.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Perri, E; Tucker, M E; Spadafora, A</p> <p>2012-03-01</p> <p>Sub-fossil stromatolites (5000-3000 years old) occur on the marginal flat surrounding <span class="hlt">Marion</span> Lake (South Australia). A micrite/microsparite crystal fabric characterises these fine-grained, well-laminated stromatolites, which lack trapped grains. The internal lamination is characterised by a sub-millimetric alternation of porous and dense laminae. The microfabric of the laminae is ubiquitously composed of a fine (10-20 μm) peloidal texture, with many thinner aphanitic layers. Aggregates of very fine, low-Mg calcite and aragonite constitute both peloidal and aphanitic micrite, which is coated, respectively, by spherulitic and fringing acicular microspar. Micrite, with a high organic matter content, is formed of coalescing nanospheres grading into small polyhedrons, probably composed mainly of aragonite, with less calcite enriched in Mg, Sr, Na and S. Bacteria-like microfossils and relics of extracellular polymeric substance (EPS) occur abundantly within this micritic framework. The former consist of empty moulds and mineralised bodies of coccoid forms, whereas EPS relics consist of sheet-like or filamentous structures that appear both mineralised and more often still preserved as a C-enriched dehydrated substance that represents the main organic matter component of the deposit. Acicular crystals, which show a prismatic elongate shape, are composed of Mg-depleted aragonite that lacks fossils or organic relicts. Degrading EPS and micro-organisms appear gradually to be replaced and entombed by the nanospherical precipitates, implying the existence of processes of organo-mineralisation within an original syn-sedimentary microbial community. Succeeding micron-scale crystals merge to form isolated or connected micritic aggregates (the peloids), followed by the gradual formation of the acicular crystals as purely inorganic precipitates. © 2011 Blackwell Publishing Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26137670','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26137670"><span>[<span class="hlt">Andreas</span> Vesalius: his rich imagination and colorful detail account in his book: 'Research of the anatomical observations of Gabriel Falloppius'].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gilias, Guy</p> <p>2015-03-01</p> <p>In a long letter, <span class="hlt">Andreas</span> Vesalius reacts to the comments made by Gabriel Falloppius to his work 'De Humani Corporis Fabrica'. In this letter, he proves Falloppius wrong in a number of assertions and corrects him on more than one occasion. In doing so, Vesalius as a renaissance humanist uses a classic Latin language with long elegant sentences in the style of the old Roman orator Cicero. Remarkably interesting is the fact that this whole argumentation is spiced with comparisons and examples from daily life. To make it clear to the reader what a certain part of the skeleton looks like, he compares this part with an object everybody knows. All parts of the human body are depicted in such an almost graphic way that even an interested reader without any medical or anatomic education can picture them. And Vesalius is very creative in doing so, an artist as it were with a very rich imagination. Moreover, it's remarkable how the famous anatomist manages to put himself on the level of any ordinary person, using comparative images on that level. This last work of Vesalius, which he himself considers to be a supplement to his De Humani Corporis Fabrica, deserves special attention, not only because it illustrates the scientific evolution of the anatomist Vesalius, but also because it offers an insight in the psychology of that fascinating scientist <span class="hlt">Andreas</span> Vesalius.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29474277','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29474277"><span>J. <span class="hlt">Marion</span> Sims and the Vesicovaginal Fistula: Historical Understanding, Medical Ethics, and Modern Political Sensibilities.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wall, L Lewis</p> <p></p> <p>To review the historical background surrounding the early work of Dr. J. <span class="hlt">Marion</span> Sims, who developed the first consistently successful surgical technique for the repair of obstetric vesicovaginal fistulas by operating on a group of young, enslaved, African American women who had this condition between 1846 and 1849. Review of primary source documents on Sims and his operations, early 19th century clinical literature on the treatment of vesicovaginal fistula, the introduction of ether and chloroform anesthesia into surgical practice, and the literature on the early 19th century medical ethics pertaining to surgical innovation. The goals are to understand Sims's operations within the clinical context of the 1840s and to avoid the problems of "presentism," in which beliefs, attitudes, and practices of the 21st century are anachronistically projected backward into the early 19th century. The object is to judge Sims within the context of his time, not to hold him accountable to standards of practice which were not developed until a century after his death. A narrative of what Sims did is presented within the context of the therapeutic options available to those with fistula in the early 19th century. Review of the available material demonstrates that Sims' first fistula operations were legal, that they were carried out with express therapeutic intent for the purpose of repairing these women's injuries, that they conformed to the ethical requirements of his time, and that they were performed with the patients' knowledge, cooperation, assent, and assistance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.T41C0631Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.T41C0631Y"><span>Crustal Density Variation Along the San <span class="hlt">Andreas</span> Fault Controls Its Secondary Faults Distribution and Dip Direction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, H.; Moresi, L. N.</p> <p>2017-12-01</p> <p>The San <span class="hlt">Andreas</span> fault forms a dominant component of the transform boundary between the Pacific and the North American plate. The density and strength of the complex accretionary margin is very heterogeneous. Based on the density structure of the lithosphere in the SW United States, we utilize the 3D finite element thermomechanical, viscoplastic model (Underworld2) to simulate deformation in the San <span class="hlt">Andreas</span> Fault system. The purpose of the model is to examine the role of a big bend in the existing geometry. In particular, the big bend of the fault is an initial condition of in our model. We first test the strength of the fault by comparing the surface principle stresses from our numerical model with the in situ tectonic stress. The best fit model indicates the model with extremely weak fault (friction coefficient < 0.1) is requisite. To the first order, there is significant density difference between the Great Valley and the adjacent Mojave block. The Great Valley block is much colder and of larger density (>200 kg/m3) than surrounding blocks. In contrast, the Mojave block is detected to find that it has lost its mafic lower crust by other geophysical surveys. Our model indicates strong strain localization at the jointer boundary between two blocks, which is an analogue for the Garlock fault. High density lower crust material of the Great Valley tends to under-thrust beneath the Transverse Range near the big bend. This motion is likely to rotate the fault plane from the initial vertical direction to dip to the southwest. For the straight section, north to the big bend, the fault is nearly vertical. The geometry of the fault plane is consistent with field observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70169888','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70169888"><span>Holocene geologic slip rate for the Banning strand of the southern San <span class="hlt">Andreas</span> Fault, southern California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Gold, Peter O.; Behr, Whitney M.; Rood, Dylan; Sharp, Warren D.; Rockwell, Thomas; Kendrick, Katherine J.; Salin, Aaron</p> <p>2015-01-01</p> <p>Northwest directed slip from the southern San <span class="hlt">Andreas</span> Fault is transferred to the Mission Creek, Banning, and Garnet Hill fault strands in the northwestern Coachella Valley. How slip is partitioned between these three faults is critical to southern California seismic hazard estimates but is poorly understood. In this paper, we report the first slip rate measured for the Banning fault strand. We constrain the depositional age of an alluvial fan offset 25 ± 5 m from its source by the Banning strand to between 5.1 ± 0.4 ka (95% confidence interval (CI)) and 6.4 + 3.7/−2.1 ka (95% CI) using U-series dating of pedogenic carbonate clast coatings and 10Be cosmogenic nuclide exposure dating of surface clasts. We calculate a Holocene geologic slip rate for the Banning strand of 3.9 + 2.3/−1.6 mm/yr (median, 95% CI) to 4.9 + 1.0/−0.9 mm/yr (median, 95% CI). This rate represents only 25–35% of the total slip accommodated by this section of the southern San <span class="hlt">Andreas</span> Fault, suggesting a model in which slip is less concentrated on the Banning strand than previously thought. In rejecting the possibility that the Banning strand is the dominant structure, our results highlight an even greater need for slip rate and paleoseismic measurements along faults in the northwestern Coachella Valley in order to test the validity of current earthquake hazard models. In addition, our comparison of ages measured with U-series and 10Be exposure dating demonstrates the importance of using multiple geochronometers when estimating the depositional age of alluvial landforms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70137269','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70137269"><span>Subsurface geometry of the San <span class="hlt">Andreas</span>-Calaveras fault junction: influence of serpentinite and the Coast Range Ophiolite</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Watt, Janet Tilden; Ponce, David A.; Graymer, Russell W.; Jachens, Robert C.; Simpson, Robert W.</p> <p>2014-01-01</p> <p>While an enormous amount of research has been focused on trying to understand the geologic history and neotectonics of the San <span class="hlt">Andreas</span>-Calaveras fault (SAF-CF) junction, fundamental questions concerning fault geometry and mechanisms for slip transfer through the junction remain. We use potential-field, geologic, geodetic, and seismicity data to investigate the 3-D geologic framework of the SAF-CF junction and identify potential slip-transferring structures within the junction. Geophysical evidence suggests that the San <span class="hlt">Andreas</span> and Calaveras fault zones dip away from each other within the northern portion of the junction, bounding a triangular-shaped wedge of crust in cross section. This wedge changes shape to the south as fault geometries change and fault activity shifts between fault strands, particularly along the Calaveras fault zone (CFZ). Potential-field modeling and relocated seismicity suggest that the Paicines and San Benito strands of the CFZ dip 65° to 70° NE and form the southwest boundary of a folded 1 to 3 km thick tabular body of Coast Range Ophiolite (CRO) within the Vallecitos syncline. We identify and characterize two steeply dipping, seismically active cross structures within the junction that are associated with serpentinite in the subsurface. The architecture of the SAF-CF junction presented in this study may help explain fault-normal motions currently observed in geodetic data and help constrain the seismic hazard. The abundance of serpentinite and related CRO in the subsurface is a significant discovery that not only helps constrain the geometry of structures but may also help explain fault behavior and the tectonic evolution of the SAF-CF junction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024566','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024566"><span>Post-1906 stress recovery of the San <span class="hlt">Andreas</span> fault system calculated from three-dimensional finite element analysis</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Parsons, T.</p> <p>2002-01-01</p> <p>The M = 7.8 1906 San Francisco earthquake cast a stress shadow across the San <span class="hlt">Andreas</span> fault system, inhibiting other large earthquakes for at least 75 years. The duration of the stress shadow is a key question in San Francisco Bay area seismic hazard assessment. This study presents a three-dimensional (3-D) finite element simulation of post-1906 stress recovery. The model reproduces observed geologic slip rates on major strike-slip faults and produces surface velocity vectors comparable to geodetic measurements. Fault stressing rates calculated with the finite element model are evaluated against numbers calculated using deep dislocation slip. In the finite element model, tectonic stressing is distributed throughout the crust and upper mantle, whereas tectonic stressing calculated with dislocations is focused mostly on faults. In addition, the finite element model incorporates postseismic effects such as deep afterslip and viscoelastic relaxation in the upper mantle. More distributed stressing and postseismic effects in the finite element model lead to lower calculated tectonic stressing rates and longer stress shadow durations (17-74 years compared with 7-54 years). All models considered indicate that the 1906 stress shadow was completely erased by tectonic loading no later than 1980. However, the stress shadow still affects present-day earthquake probability. Use of stressing rate parameters calculated with the finite element model yields a 7-12% reduction in 30-year probability caused by the 1906 stress shadow as compared with calculations not incorporating interactions. The aggregate interaction-based probability on selected segments (not including the ruptured San <span class="hlt">Andreas</span> fault) is 53-70% versus the noninteraction range of 65-77%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhDT.......173A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhDT.......173A"><span>Gravity constraints on the geometry of the Big Bend of the San <span class="hlt">Andreas</span> Fault in the southern Carrizo Plains and Pine Mountain egion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Altintas, Ali Can</p> <p></p> <p>The goal of this project is to combine gravity measurements with geologic observations to better understand the "Big Bend" of the San <span class="hlt">Andreas</span> Fault (SAF) and its role in producing hydrocarbon-bearing structures in the southern Central Valley of California. The SAF is the main plate boundary structure between the Pacific and North American plates and accommodates ?35 mm/yr of dextral motion. The SAF can be divided into three main parts: the northern, central and southern segments. The boundary between the central and southern segments is the "Big Bend", which is characterized by an ≈30°, eastward bend. This fault curvature led to the creation of a series of roughly east-west thrust faults and the transverse mountain ranges. Four high-resolution gravity transects were conducted across locations on either side of the bend. A total of 166 new gravity measurements were collected. Previous studies suggest significantly inclined dip angle for the San <span class="hlt">Andreas</span> Fault in the Big Bend area. Yet, our models indicate that the San <span class="hlt">Andreas</span> Fault is near vertical in the Big Bend area. Also gravity cross-section models suggest that flower structures occur on either side of the bend. These structures are dominated by sedimentary rocks in the north and igneous rocks in the south. The two northern transects in the Carrizo plains have an ≈-70 mgal Bouguer anomaly. The SAF has a strike of ≈315° near these transects. The northern transects are characterized by multiple fault strands which cut marine and terrestrial Miocene sedimentary rocks as well as Quaternary alluvial valley deposits. These fault strands are characterized by ?6 mgal short wavelength variations in the Bouguer gravity anomaly, which correspond to low density fault gouge and fault splays that juxtapose rocks of varying densities. The southern transects cross part of the SAF with a strike of 285°, have a Bouguer anomaly of ≈-50 mgal and are characterized by a broad 15 mgal high. At this location the rocks on</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4671350','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4671350"><span>Balloon Angioplasty – The Legacy of <span class="hlt">Andreas</span> Grüntzig, M.D. (1939–1985)</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Barton, Matthias; Grüntzig, Johannes; Husmann, Marc; Rösch, Josef</p> <p>2014-01-01</p> <p>In 1974, at the Medical Policlinic of the University of Zürich, German-born physician-scientist <span class="hlt">Andreas</span> Grüntzig (1939–1985) for the first time applied a balloon-tipped catheter to re-open a severely stenosed femoral artery, a procedure, which he initially called “percutaneous transluminal dilatation”. Balloon angioplasty as a therapy of atherosclerotic vascular disease, for which Grüntzig and Charles T. Dotter (1920–1985) received a nomination for the Nobel Prize in Physiology or Medicine in 1978, became one of the most successful examples of translational medicine in the twentieth century. Known today as percutaneous transluminal angioplasty (PTA) in peripheral arteries or percutaneous transluminal coronary angioplasty (PTCA) or percutaneous coronary intervention (PCI) in coronary arteries, balloon angioplasty has become the method of choice to treat patients with acute myocardial infarction or occluded leg arteries. On the occasion of the 40th anniversary of balloon angioplasty, we summarize Grüntzig’s life and career in Germany, Switzerland, and the United States and also review the developments in vascular medicine from the 1890s to the 1980s, including Dotter’s first accidental angioplasty in 1963. The work of pioneers of catheterization, including Pedro L. Fariñas in Cuba, André F. Cournand in France, Werner Forssmann, Werner Porstmann and Eberhard Zeitler in Germany, António Egas Moniz and Reynaldo dos Santos in Portugal, Sven-Ivar Seldinger in Sweden, and Barney Brooks, Thomas J. Fogarty, Melvin P. Judkins, Richard K. Myler, Dickinson W. Richards, and F. Mason Sones in the United States, is discussed. We also present quotes by Grüntzig and excerpts from his unfinished autobiography, statements of Grüntzig’s former colleagues and contemporary witnesses, and have included hitherto unpublished historic photographs and links to archive recordings and historic materials. This year, on June 25, 2014, <span class="hlt">Andreas</span> Grüntzig would have celebrated</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.T51C0676K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.T51C0676K"><span>Core Across the San <span class="hlt">Andreas</span> Fault at SAFOD - Photographs, Physical Properties Data, and Core-Handling Procedures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kirschner, D. L.; Carpenter, B.; Keenan, T.; Sandusky, E.; Sone, H.; Ellsworth, B.; Hickman, S.; Weiland, C.; Zoback, M.</p> <p>2007-12-01</p> <p>Core samples were obtained that cross three faults of the San <span class="hlt">Andreas</span> Fault Zone north of Parkfield, California, during the summer of 2007. The cored intervals were obtained by sidetracking off the SAFOD Main Hole that was rotary drilled across the San <span class="hlt">Andreas</span> in 2005. The first cored interval targeted the pronounced lithologic boundary between the Salinian terrane and the Great Valley and Franciscan formations. Eleven meters of pebbly conglomerate (with minor amounts of fine sands and shale) were obtained from 3141 to 3152 m (measured depth, MD). The two conglomerate units are heavily fractured with many fractures having accommodated displacement. Within this cored interval, there is a ~1m zone with highly sheared, fine-grained material, possibly ultracataclasite in part. The second cored interval crosses a creeping segment of a fault that has been deforming the cemented casing of the adjacent Main Hole. This cored interval sampled the fault 100 m above a seismogenic patch of M2 repeating earthquakes. Thirteen meters of core were obtained across this fault from 3186 to 3199 m (MD). This fault, which is hosted primarily in siltstones and shales, contains a serpentinite body embedded in a highly sheared shale and serpentinite-bearing fault gouge unit. The third cored interval crosses a second creeping fault that has also been deforming the cemented casing of the Main Hole. This fault, which is the most rapidly shearing fault in the San <span class="hlt">Andreas</span> fault zone based on casing deformation, contains multiple fine- grained clay-rich fault strands embedded in highly sheared shales and lesser deformed sandstones. Initial processing of the cores was carried out at the drill site. Each core came to the surface in 9 meter-long aluminum core barrels. These were cut into more manageable three-foot sections. The quarter-inch-thick aluminum liner of each section was cut and then split apart to reveal the 10 cm diameter cores. Depending on the fragility and porosity of the rock, the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/ds/413/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/ds/413/"><span>Data Files for Ground-Motion Simulations of the 1906 San Francisco Earthquake and Scenario Earthquakes on the Northern San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Aagaard, Brad T.; Barall, Michael; Brocher, Thomas M.; Dolenc, David; Dreger, Douglas; Graves, Robert W.; Harmsen, Stephen; Hartzell, Stephen; Larsen, Shawn; McCandless, Kathleen; Nilsson, Stefan; Petersson, N. Anders; Rodgers, Arthur; Sjogreen, Bjorn; Zoback, Mary Lou</p> <p>2009-01-01</p> <p>This data set contains results from ground-motion simulations of the 1906 San Francisco earthquake, seven hypothetical earthquakes on the northern San <span class="hlt">Andreas</span> Fault, and the 1989 Loma Prieta earthquake. The bulk of the data consists of synthetic velocity time-histories. Peak ground velocity on a 1/60th degree grid and geodetic displacements from the simulations are also included. Details of the ground-motion simulations and analysis of the results are discussed in Aagaard and others (2008a,b).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFMPP32B..02D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFMPP32B..02D"><span>The Initiation of the Southern Central Great Barrier Reef: Multiproxy Study on Periplatform Sediments From ODP Site 1195 on the <span class="hlt">Marion</span> Plateau (NE Australia)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dubois, N.; Kindler, P.; Spezzaferri, S.; Coric, S.</p> <p>2007-12-01</p> <p>The sediments deposited at ODP Site 1195 (<span class="hlt">Marion</span> Plateau, NE Australia) record synchronous shifts in their chemistry, mineralogy, grain size and colour at 6 meters below sea floor. These significant changes are interpreted to reflect the onset of the southern province of the Great Barrier Reef (GBR). An increased deposition of carbonate-rich sediments of neritic origin, coincident with a decline in both sedimentation rate and terrigenous input, is attributed to inshore trapping of materials by the reefs. Based on an age model combining magnetostratigraphic and biostratigraphic data, we propose that the southern province of the GBR initiated between 560 and 670 kyr B.P. Our best estimate concurs with previous studies reporting an age between 500 and 930 kyr B.P., albeit constraining more tightly these earlier age estimates. However, it does not support research placing the birth of the GBR in Marine Isotope Stage 11 (about 400 kyr B.P.), nor the recent theory of a worldwide modern reef development at that time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.G33A1082M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.G33A1082M"><span>Shallow and deep creep events observed and quantified with strainmeters along the San <span class="hlt">Andreas</span> Fault near Parkfield</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mencin, D.; Hodgkinson, K. M.; Mattioli, G. S.; Johnson, W.; Gottlieb, M. H.; Meertens, C. M.</p> <p>2016-12-01</p> <p>Three-component strainmeter data from numerous borehole strainmeters (BSM) along the San <span class="hlt">Andreas</span> Fault (SAF), including those that were installed and maintained as part of the EarthScope Plate Boundary Observatory (PBO), demonstrate that the characteristics of creep propagation events with sub-cm slip amplitudes can be quantified for slip events at 10 km source-to-sensor distances. The strainmeters are installed at depths of approximately 100 - 250 m and record data at a rate of 100 samples per second. Noise levels at periods of less than a few minutes are 10-11 strain, and for periods in the bandwidth hours to weeks, the periods of interest in the search for slow slip events, are of the order of 10-8 to 10-10 strain. Strainmeters, creepmeters, and tiltmeters have been operated along the San <span class="hlt">Andreas</span> Fault, observing creep events for decades. BSM data proximal to the SAF cover a significant temporal portion of the inferred earthquake cycle along this portion of the fault. A single instrument is capable of providing broad scale constraints of creep event asperity size, location, and depth and moreover can capture slow slip, coseismic rupture as well as afterslip. The synthesis of these BSM data presents a unique opportunity to constrain the partitioning between aseismic and seismic slip on the central SAF. We show that the creepmeters confirm that creep events that are imaged by the strainmeters, previously catalogued by the authors, are indeed occurring on the SAF, and are simultaneously being recorded on local creepmeters. We further show that simple models allow us to loosely constrain the location and depth of the creep event on the fault, even with a single instrument, and to image the accumulation and behavior of surface as well as crustal creep with time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024102','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024102"><span>Evidence for two surface ruptures in the past 500 years on the San <span class="hlt">Andreas</span> fault at Frazier Mountain, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lindvall, S.C.; Rockwell, T.K.; Dawson, T.E.; Helms, J.G.; Bowman, K.W.</p> <p>2002-01-01</p> <p>We conducted paleoseismic studies in a closed depression along the San <span class="hlt">Andreas</span> fault on the north flank of Frazier Mountain near Frazier Park, California. We recognized two earthquake ruptures in our trench exposure and interpreted the most recent rupture, event 1, to represent the historical 1857 earthquake. We also exposed evidence of an earlier surface rupture, event 2, along an older group of faults that did not rerupture during event 1. Radiocarbon dating of the stratigraphy above and below the earlier event constrains its probable age to between A.D. 1460 and 1600. Because we documented continuous, unfaulted stratigraphy between the earlier event horizon and the youngest event horizon in the portion of the fault zone exposed, we infer event 2 to be the penultimate event. We observed no direct evidence of an 1812 earthquake in our exposures. However, we cannot preclude the presence of this event at our site due to limited age control in the upper part of the section and the possibility of other fault strands beyond the limits of our exposures. Based on overlapping age ranges, event 2 at Frazier Mountain may correlate with event B at the Bidart fan site in the Carrizo Plain to the northwest and events V and W4 at Pallett Creek and Wrightwood, respectively, to the southeast. If the events recognized at these multiple sites resulted from the same surface rupture, then it appears that the San <span class="hlt">Andreas</span> fault has repeatedly failed in large ruptures similar in extent to 1857.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3743612','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3743612"><span>Christian <span class="hlt">Andreas</span> Doppler: A legendary man inspired by the dazzling light of the stars</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Katsi, V; Felekos, I; Kallikazaros, I</p> <p>2013-01-01</p> <p>Christian <span class="hlt">Andreas</span> Doppler is renowned primarily for his revolutionary theory of the Doppler effect, which has deeply influenced many areas of modern science and technology, including medicine. His work has laid the foundations for modern ultrasonography and his ideas are still inspiring discoveries more than a hundred years after his death. Doppler may well earn the title of Homo Universalis for his broad knowledge of physics, mathematics and astronomy and most of all for his indefatigable investigations for new ideas and his ingenious mind. According to Bolzano: “It is hard to believe how fruitful a genius Austria has in this man”. His legacy of scientific achievement have seen Doppler honoured in the later years on coinage and money, names of streets, educational institutions, rock groups, even of a lunar crater; while the ultimate tribute to his work is the countless references to the homonymous medical eponym. PMID:24376313</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/1970/0044/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/1970/0044/report.pdf"><span>Map showing recently active breaks along the San <span class="hlt">Andreas</span> Fault between Pt. Delgada and Bolinas Bay, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Brown, Robert D.; Wolfe, Edward W.</p> <p>1970-01-01</p> <p>This strip map is one of a series of maps showing recently active fault breaks along the San <span class="hlt">Andreas</span> and other active faults in California. It is designed to inform persons who are concerned with land use near the fault of the location of those fault breaks that have moved recently. The lines on the map are lines of rupture and creep that can be identified by field evidence and that clearly affect the present surface of the land. Map users should keep in mind that these lines are intended primarily as guides to help locate the fault; the mapped lines are not necessarily shown with the precision demanded by some engineering or land utilization needs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70042315','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70042315"><span>Low-altitude aerial color digital photographic survey of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lynch, David K.; Hudnut, Kenneth W.; Dearborn, David S.P.</p> <p>2010-01-01</p> <p>Ever since 1858, when Gaspard-Félix Tournachon (pen name Félix Nadar) took the first aerial photograph (Professional Aerial Photographers Association 2009), the scientific value and popular appeal of such pictures have been widely recognized. Indeed, Nadar patented the idea of using aerial photographs in mapmaking and surveying. Since then, aerial imagery has flourished, eventually making the leap to space and to wavelengths outside the visible range. Yet until recently, the availability of such surveys has been limited to technical organizations with significant resources. Geolocation required extensive time and equipment, and distribution was costly and slow. While these situations still plague older surveys, modern digital photography and lidar systems acquire well-calibrated and easily shared imagery, although expensive, platform-specific software is sometimes still needed to manage and analyze the data. With current consumer-level electronics (cameras and computers) and broadband internet access, acquisition and distribution of large imaging data sets are now possible for virtually anyone. In this paper we demonstrate a simple, low-cost means of obtaining useful aerial imagery by reporting two new, high-resolution, low-cost, color digital photographic surveys of selected portions of the San <span class="hlt">Andreas</span> fault in California. All pictures are in standard jpeg format. The first set of imagery covers a 92-km-long section of the fault in Kern and San Luis Obispo counties and includes the entire Carrizo Plain. The second covers the region from Lake of the Woods to Cajon Pass in Kern, Los Angeles, and San Bernardino counties (151 km) and includes Lone Pine Canyon soon after the ground was largely denuded by the Sheep Fire of October 2009. The first survey produced a total of 1,454 oblique digital photographs (4,288 x 2,848 pixels, average 6 Mb each) and the second produced 3,762 nadir images from an elevation of approximately 150 m above ground level (AGL) on the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2014/1002/pdf/ofr2014-1002_pamphlet.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2014/1002/pdf/ofr2014-1002_pamphlet.pdf"><span>Photomosaics and event evidence from the Frazier Mountain paleoseismic site, trench 1, cuts 1–4, San <span class="hlt">Andreas</span> Fault Zone, southern California (2007–2009)</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Scharer, Katherine M.; Fumal, Tom E.; Weldon, Ray J.; Streig, Ashley R.</p> <p>2014-01-01</p> <p>The Frazier Mountain paleoseismic site is located at the northwest end of the Mojave section of the San <span class="hlt">Andreas</span> Fault, in a small, closed depression at the base of Frazier Mountain near Tejon Pass, California (lat 34.8122° N., long 118.9034° W.). The site was known to contain a good record of earthquakes due to previous excavations by Lindvall and others (2002). This report provides data resulting from four nested excavations, or cuts, along trench 1 (T1) in 2007 and 2009 at the Frazier Mountain site. The four cuts were excavated progressively deeper and wider in an orientation perpendicular to the San <span class="hlt">Andreas</span> Fault, exposing distal fan and marsh sediments deposited since ca. A.D. 1200. The results of the trenching show that earthquakes that ruptured the site have repeatedly produced a small depression or sag on the surface, which is subsequently infilled with sand and silt deposits. This report provides high-resolution photomosaics and logs for the T1 cuts, a detailed stratigraphic column for the deposits, and a table summarizing all of the evidence for ground rupturing paleoearthquakes logged in the trenches.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26804787','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26804787"><span><span class="hlt">Andreas</span> Vesalius' 500th Anniversary: Initial Integral Understanding of Voice Production.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brinkman, Romy J; Hage, J Joris</p> <p>2017-01-01</p> <p>Voice production relies on the integrated functioning of a three-part system: respiration, phonation and resonance, and articulation. To commemorate the 500th anniversary of the great anatomist <span class="hlt">Andreas</span> Vesalius (1515-1564), we report on his understanding of this integral system. The text of Vesalius' masterpiece De Humani Corporis Fabrica Libri Septum and an eyewitness report of the public dissection of three corpses by Vesalius in Bologna, Italy, in 1540, were searched for references to the voice-producing anatomical structures and their function. We clustered the traced, separate parts for the first time. We found that Vesalius recognized the importance for voice production of many details of the respiratory system, the voice box, and various structures of resonance and articulation. He stressed that voice production was a cerebral function and extensively recorded the innervation of the voice-producing organs by the cranial nerves. Vesalius was the first to publicly record the concept of voice production as an integrated and cerebrally directed function of respiration, phonation and resonance, and articulation. In doing so nearly 500 years ago, he laid a firm basis for the understanding of the physiology of voice production and speech and its management as we know it today. Copyright © 2017 The Voice Foundation. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70191859','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70191859"><span>Ground-rupturing earthquakes on the northern Big Bend of the San <span class="hlt">Andreas</span> Fault, California, 800 A.D. to Present</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Scharer, Katherine M.; Weldon, Ray; Biasi, Glenn; Streig, Ashley; Fumal, Thomas E.</p> <p>2017-01-01</p> <p>Paleoseismic data on the timing of ground-rupturing earthquakes constrain the recurrence behavior of active faults and can provide insight on the rupture history of a fault if earthquakes dated at neighboring sites overlap in age and are considered correlative. This study presents the evidence and ages for 11 earthquakes that occurred along the Big Bend section of the southern San <span class="hlt">Andreas</span> Fault at the Frazier Mountain paleoseismic site. The most recent earthquake to rupture the site was the Mw7.7–7.9 Fort Tejon earthquake of 1857. We use over 30 trench excavations to document the structural and sedimentological evolution of a small pull-apart basin that has been repeatedly faulted and folded by ground-rupturing earthquakes. A sedimentation rate of 0.4 cm/yr and abundant organic material for radiocarbon dating contribute to a record that is considered complete since 800 A.D. and includes 10 paleoearthquakes. Earthquakes have ruptured this location on average every ~100 years over the last 1200 years, but individual intervals range from ~22 to 186 years. The coefficient of variation of the length of time between earthquakes (0.7) indicates quasiperiodic behavior, similar to other sites along the southern San <span class="hlt">Andreas</span> Fault. Comparison with the earthquake chronology at neighboring sites along the fault indicates that only one other 1857-size earthquake could have occurred since 1350 A.D., and since 800 A.D., the Big Bend and Mojave sections have ruptured together at most 50% of the time in Mw ≥ 7.3 earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1987BVol...49..399V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1987BVol...49..399V"><span>Contrasting types of surtseyan tuff cones on <span class="hlt">Marion</span> and Prince Edward islands, southwest Indian Ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Verwoerd, W. J.; Chevallier, L.</p> <p>1987-02-01</p> <p>Ten surtseyan tuff cones on <span class="hlt">Marion</span> island (46° 54' S, 37° 46' E) and seven on Prince Edward island (46° 38' S, 37° 57' E) were erupted on shallow submerged coastal plains related to normal faulting. They range from Pleistocene to Holocene in age and exhibit a variable degree of erosion by the sea. Fundamental differences, irrespective of age, exist between two types: Type I cones have diameters of 1 1.5 km, rim heights of about 200 m and steep (27°) outer slopes. Deposits are plastered against nearby cliffs. Beds are thin, including layers of accretionary lapilli and less than 10 % lithic clasts. Numerous bomb sags, soft sediment deformation structures and gravity slides occur. On one of these cones mudflows formed small tunnels which resemble lava tubes, associated with channels sometimes having oversteepened walls. These cones reflect comparatively low energy conditions and probably resulted from extremely wet surges interspersed with mudflows and ballistic falls. Type II cones have smaller diameters (˜0.5 km) but widespread fallout/surge aprons. Rim heights are about 100 m and average slope angles are 18°. Bedding is massive with variable lapilli/matrix ratio and more than 10 % lithic clasts without bomb sags. These cones formed under drier, perhaps hotter and more violently explosive conditions than Type I, though not as energetic as the phreatomagmatic eruptions of terrestrial tuff rings. The two types of surtseyan eruptions are explained by invoking not only different water/magma ratios in the conduit but also different mechanisms of water/magma interaction. The slurry model of Kokelaar is favoured for Type I and a fuel-coolant model for Type II. The decisive factor is considered to be rate of effusion, with rim closure and exclusion of sea water playing a secondary role.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11875901','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11875901"><span>[Lou <span class="hlt">Andreas</span>-Salome (1861-1937)--psychoanalytical and feministic contribution to understanding her biography].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bramness, J G</p> <p>2001-06-30</p> <p>Lou (Louise) <span class="hlt">Andreas</span>-Salomé's life and work has preoccupied many biographers. The interest may have be sparked by her liaisons with many of the greatest men of her time. She had an intimate relationship with Friedrich Nietzsche in a period of great change for him. She was Rainer Marie Rilke's mistress for several years. And she pursued a close friendship and working relationship with Sigmund Freud in the latter part of her life. But her significance goes beyond these associations. She was a celebrated novelist and essayist in her own right, with ten novels and more than 50 essays, also on psychoanalytical subjects. She has been viewed as femme fatale, opportunist, feminist, radical, liberal, but also as a significant contributor to psychoanalytical thought. There have been two biographical approaches: a psychoanalytical approach focusing on her loss of father-figures and later difficult relationships with famous men, and a feministic approach accusing psychoanalysts of not contributing to insight, but belittling Salomé's legitimate position. A fuller understanding may be obtained by integrating these two views.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.earthquakegeology.com/index.php?page=publications&s=6','USGSPUBS'); return false;" href="http://www.earthquakegeology.com/index.php?page=publications&s=6"><span>Testing geomorphology-derived rupture histories against the paleoseismic record of the southern San <span class="hlt">Andreas</span> fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Scharer, Katherine M.; Weldon, Ray; Bemis, Sean</p> <p>2016-01-01</p> <p>Evidence for the 340-km-long Fort Tejon earthquake of 1857 is found at each of the high-resolution paleoseismic sites on the southern San <span class="hlt">Andreas</span> Fault. Using trenching data from these sites, we find that the assemblage of dated paleoearthquakes recurs quasi-periodically (coefficient of variation, COV, of 0.6, Biasi, 2013) and requires ~80% of ruptures were shorter than the 1857 rupture with an average of Mw7.5. In contrast, paleorupture lengths reconstructed from preserved geomorphic offsets extracted from lidar are longer and have repeating displacements that are quite regular (COV=0.2; Zielke et al., 2015). Direct comparison shows that paleoruptures determined from geomorphic offset populations cannot be reconciled with dated paleoearthquakes. Our study concludes that the 1857 rupture was larger than average, average displacements must be < 5 m, and suggests that fault geometry may play a role in fault behavior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://woodshole.er.usgs.gov/pubs/of2004-1358/index.htm','USGSPUBS'); return false;" href="http://woodshole.er.usgs.gov/pubs/of2004-1358/index.htm"><span>Initial report of the IMAGES VIII/PAGE 127 gas hydrate and paleoclimate cruise on the RV <span class="hlt">Marion</span> Dufresne in the Gulf of Mexico, 2-18 July 2002</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Winters, William J.; Lorenson, T.D.; Paull, Charles K.</p> <p>2007-01-01</p> <p>The northern Gulf of Mexico contains many documented gas hydrate deposits near the sea floor. Although gas hydrate often is present in shallow subbottom sediment, the extent of hydrate occurrence deeper than 10 meters below sea floor in basins away from vents and other surface expressions is unknown. We obtained giant piston cores, box cores, and gravity cores and performed heat-flow analyses to study these shallow gas hydrate deposits aboard the RV <span class="hlt">Marion</span> Dufresne in July 2002. This report presents measurements and interpretations from that cruise. Our results confirm the presence of gas hydrate in vent-related sediments near the sea bed. The presence of gas hydrate near the vents is governed by the complex interaction of regional and local factors, including heat flow, fluid flow, faults, pore-water salinity, gas concentrations, and sediment properties. However, conditions appropriate for extensive gas hydrate formation were not found away from the vents.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006PhDT.......103B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006PhDT.......103B"><span>Dame Bug and her students: The science and environmental teaching of Edith <span class="hlt">Marion</span> Patch</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bird, Mary Dickinson</p> <p></p> <p>What can we learn about science teaching from a dead entomologist? And what will this learning mean in the context of our own teaching, half a century after her death? This project is an exploration of the educational ideas and practices of Edith <span class="hlt">Marion</span> Patch (1876-1954), a pioneering entomologist, educator, and environmentalist whose research and teaching shaped science learning among farmers, children, parents, teachers, and fellow scientists from the start of her career in 1903 until her retirement from professional endeavors in 1945. Patch was one of America's first university-trained female entomologists, the first woman president of the Entomological Society of America and one of the most respected scientists of her time. She was a beloved, bestselling author of more than one hundred natural history stories and books for young learners and dozens of nature and gardening articles for adults. She was also one of the earliest environmentalists to sound an alarm about the hazards of chemical pesticides, delivering her first address on the topic in 1906, one year before Rachel Carson was born. Because of the respect she earned in every milieu in which she worked, Edith Patch was uniquely positioned to influence a wide range of learners, from fellow scientists to home gardeners, to little children. Despite her accomplishments and her renown during her lifetime, Edith Patch is little remembered today. Yet her educational work may prove to be of particular relevance for our time, in which scientific literacy and environmental stewardship take on increasing urgency. Indeed, in the current climate of educational reform, which in some aspects reproduces the inquiry-based, experiential approach advocated by Patch and other education reformers a century ago, a close examination of Patch's teaching may prove to be of particular value. This study therefore seeks to fill the gap in our pedagogical memory regarding this eminent woman, considering the copious evidence of ways</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRB..122.3739S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRB..122.3739S"><span>A 15 year catalog of more than 1 million low-frequency earthquakes: Tracking tremor and slip along the deep San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shelly, David R.</p> <p>2017-05-01</p> <p>Low-frequency earthquakes (LFEs) are small, rapidly recurring slip events that occur on the deep extensions of some major faults. Their collective activation is often observed as a semicontinuous signal known as tectonic (or nonvolcanic) tremor. This manuscript presents a catalog of more than 1 million LFEs detected along the central San <span class="hlt">Andreas</span> Fault from 2001 to 2016. These events have been detected via a multichannel matched-filter search, cross-correlating waveform templates representing 88 different LFE families with continuous seismic data. Together, these source locations span nearly 150 km along the central San <span class="hlt">Andreas</span> Fault, ranging in depth from 16 to 30 km. This accumulating catalog has been the source for numerous studies examining the behavior of these LFE sources and the inferred slip behavior of the deep fault. The relatively high temporal and spatial resolutions of the catalog have provided new insights into properties such as tremor migration, recurrence, and triggering by static and dynamic stress perturbations. Collectively, these characteristics are inferred to reflect a very weak fault likely under near-lithostatic fluid pressure, yet the physical processes controlling the stuttering rupture observed as tremor and LFE signals remain poorly understood. This paper aims to document the LFE catalog assembly process and associated caveats, while also updating earlier observations and inferred physical constraints. The catalog itself accompanies this manuscript as part of the electronic supplement, with the goal of providing a useful resource for continued future investigations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S21B0704S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S21B0704S"><span>Break of slope in earthquake size distribution and creep rate along the San <span class="hlt">Andreas</span> Fault system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shebalin, P.; Narteau, C.; Vorobieva, I.</p> <p>2017-12-01</p> <p>Crustal faults accommodate slip either by a succession of earthquakes or continuous slip, andin most instances, both these seismic and aseismic processes coexist. Recorded seismicity and geodeticmeasurements are therefore two complementary data sets that together document ongoing deformationalong active tectonic structures. Here we study the influence of stable sliding on earthquake statistics.We show that creep along the San <span class="hlt">Andreas</span> Fault is responsible for a break of slope in the earthquake sizedistribution. This slope increases with an increasing creep rate for larger magnitude ranges, whereas itshows no systematic dependence on creep rate for smaller magnitude ranges. This is interpreted as a deficitof large events under conditions of faster creep where seismic ruptures are less likely to propagate. Theseresults suggest that the earthquake size distribution does not only depend on the level of stress but also onthe type of deformation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA02746&hterms=time+perspective&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dtime%2Bperspective','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA02746&hterms=time+perspective&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dtime%2Bperspective"><span>Perspective View, San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2000-01-01</p> <p><p/> The prominent linear feature straight down the center of this perspective view is the San <span class="hlt">Andreas</span> Fault in an image created with data from NASA's shuttle Radar Topography Mission (SRTM), which will be used by geologists studying fault dynamics and landforms resulting from active tectonics. This segment of the fault lies west of the city of Palmdale, California, about 100 kilometers (about 60 miles) northwest of Los Angeles. The fault is the active tectonic boundary between the North American plate on the right, and the Pacific plate on the left. Relative to each other, the Pacific plate is moving away from the viewer and the North American plate is moving toward the viewer along what geologists call a right lateral strike-slip fault. This area is at the junction of two large mountain ranges, the San Gabriel Mountains on the left and the Tehachapi Mountains on the right. Quail Lake Reservoir sits in the topographic depression created by past movement along the fault. Interstate 5 is the prominent linear feature starting at the left edge of the image and continuing into the fault zone, passing eventually over Tejon Pass into the Central Valley, visible at the upper left.<p/>This type of display adds the important dimension of elevation to the study of land use and environmental processes as observed in satellite images. The perspective view was created by draping a Landsat satellite image over an SRTM elevation model. Topography is exaggerated 1.5 times vertically. The Landsat image was provided by the United States Geological Survey's Earth Resources Observations Systems (EROS) Data Center, Sioux Falls, South Dakota.<p/>Elevation data used in this image was acquired by the Shuttle Radar Topography Mission (SRTM) aboard the Space Shuttle Endeavour, launched on February 11,2000. SRTM used the same radar instrument that comprised the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) that flew twice on the Space Shuttle Endeavour in 1994</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020682','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020682"><span>Ductile shear zones beneath strike-slip faults: Implications for the thermomechanics of the San <span class="hlt">Andreas</span> fault zone</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thatcher, W.; England, P.C.</p> <p>1998-01-01</p> <p>We have carried out two-dimensional (2-D) numerical experiments on the bulk flow of a layer of fluid that is driven in a strike-slip sense by constant velocities applied at its boundaries. The fluid has the (linearized) conventional rheology assumed to apply to lower crust/upper mantle rocks. The temperature dependence of the effective viscosity of the fluid and the shear heating that accompanies deformation have been incorporated into the calculations, as has thermal conduction in an overlying crustal layer. Two end-member boundary conditions have been considered, corresponding to a strong upper crust driving a weaker ductile substrate and a strong ductile layer driving a passive, weak crust. In many cases of practical interest, shear heating is concentrated close to the axial plane of the shear zone for either boundary condition. For these cases, the resulting steady state temperature field is well approximated by a cylindrical heat source embedded in a conductive half-space at a depth corresponding to the top of the fluid layer. This approximation, along with the application of a theoretical result for one-dimensional shear zones, permits us to obtain simple analytical approximations to the thermal effects of 2-D ductile shear zones for a range of assumed rheologies and crustal geotherms, making complex numerical calculations unnecessary. Results are compared with observable effects on heat flux near the San <span class="hlt">Andreas</span> fault using constraints on the slip distribution across the entire fault system. Ductile shearing in the lower crust or upper mantle can explain the observed increase in surface heat flux southeast of the Mendocino triple junction and match the amplitude of the regional heat flux anomaly in the California Coast Ranges. Because ductile dissipation depends only weakly on slip rate, faults moving only a few millimeters per year can be important heat sources, and the superposition of effects of localized ductile shearing on both currently active and now</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.T43C3011M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.T43C3011M"><span>Synthesis of Creep Measurements from Strainmeters and Creepmeters along the San <span class="hlt">Andreas</span> Fault: Implications for Seismic vs. Aseismic Partitioning</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mencin, D.; Gottlieb, M. H.; Hodgkinson, K. M.; Bilham, R. G.; Mattioli, G. S.; Johnson, W.; Van Boskirk, E.; Meertens, C. M.</p> <p>2015-12-01</p> <p>Strainmeters and creepmeters have been operated along the San <span class="hlt">Andreas</span> Fault, observing creep events for decades. In particular, the EarthScope Plate Boundary Observatory (PBO) has added a significant number of borehole strainmeters along the San <span class="hlt">Andreas</span> Fault (SAF) over the last decade. The geodetic data cover a significant temporal portion of the inferred earthquake cycle along this portion of the SAF. Creepmeters measure the surface displacement over time (creep) with short apertures and have the ability to capture slow slip, coseismic rupture, and afterslip. Modern creepmeters deployed by the authors have a resolution of 5 µm over a range of 10 mm and a dynamic sensor with a resolution 25 µm over a range 2.2 m. Borehole strainmeters measure local deformation some distance from the fault with a broader aperture. Borehole tensor strainmeters principally deployed as part of the PBO, measure the horizontal strain tensor at a depth of 100-200 m with a resolution of 10-11 strain and are located 4 - 10 km from the fault with the ability to image a 1 mm creep event acting on an area of ~500 m2 from over 4 km away (fault perpendicular). A single borehole tensor strainmeter is capable of providing broad constraints on the creep event asperity size, location, direction and depth of a single creep event. The synthesis of these data from all the available geodetic instruments proximal to the SAF presents a unique opportunity to constrain the partitioning between aseismic and seismic slip on the central SAF. We show that simple elastic half-space models allow us to loosely constrain the location and depth of any individual creep event on the fault, even with a single instrument, and to image the accumulation of creep with time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70042290','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70042290"><span>Precise tremor source locations and amplitude variations along the lower-crustal central San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Shelly, David R.; Hardebeck, Jeanne L.</p> <p>2010-01-01</p> <p>We precisely locate 88 tremor families along the central San <span class="hlt">Andreas</span> Fault using a 3D velocity model and numerous P and S wave arrival times estimated from seismogram stacks of up to 400 events per tremor family. Maximum tremor amplitudes vary along the fault by at least a factor of 7, with by far the strongest sources along a 25 km section of the fault southeast of Parkfield. We also identify many weaker tremor families, which have largely escaped prior detection. Together, these sources extend 150 km along the fault, beneath creeping, transitional, and locked sections of the upper crustal fault. Depths are mostly between 18 and 28 km, in the lower crust. Epicenters are concentrated within 3 km of the surface trace, implying a nearly vertical fault. A prominent gap in detectible activity is located directly beneath the region of maximum slip in the 2004 magnitude 6.0 Parkfield earthquake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940014955','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940014955"><span>Analysis of regional deformation and strain accumulation data adjacent to the San <span class="hlt">Andreas</span> fault</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Turcotte, Donald L.</p> <p>1991-01-01</p> <p>A new approach to the understanding of crustal deformation was developed under this grant. This approach combined aspects of fractals, chaos, and self-organized criticality to provide a comprehensive theory for deformation on distributed faults. It is hypothesized that crustal deformation is an example of comminution: Deformation takes place on a fractal distribution of faults resulting in a fractal distribution of seismicity. Our primary effort under this grant was devoted to developing an understanding of distributed deformation in the continental crust. An initial effort was carried out on the fractal clustering of earthquakes in time. It was shown that earthquakes do not obey random Poisson statistics, but can be approximated in many cases by coupled, scale-invariant fractal statistics. We applied our approach to the statistics of earthquakes in the New Hebrides region of the southwest Pacific because of the very high level of seismicity there. This work was written up and published in the Bulletin of the Seismological Society of America. This approach was also applied to the statistics of the seismicity on the San <span class="hlt">Andreas</span> fault system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0543.photos.220226p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0543.photos.220226p/"><span>Substructure detail view of the castinplace concrete bents and steel, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Substructure detail view of the cast-in-place concrete bents and steel, longitudinal "I" beams. - <span class="hlt">Marion</span> Creek Bridge, Spanning <span class="hlt">Marion</span> Creek at Milepoint 66.42 on North Santiam Highway (OR-22), <span class="hlt">Marion</span> Forks, Linn County, OR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24585578','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24585578"><span><span class="hlt">Andrea</span> Pasta (1706-1782), eclectic scholar of anatomy and clinical medicine, communication and the history of art.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Clerici, Carlo Alfredo; Veneroni, Laura; Patriarca, Carlo</p> <p>2014-11-01</p> <p><span class="hlt">Andrea</span> Pasta was an eclectic visionary light years ahead of his time. He made numerous contributions to the field of medicine, some recognized by his contemporaries and others so visionary that they are being applied only in modern times. His contributions spanned the disciplines of psychology, gynaecology, haematology, infectious diseases and the doctor-patient relationship. Well known among his contemporaries, he combined a passion for clinical medicine and a keen interest in history and art with a strict research methodology and an approach to caring for patients as human beings. By studying his life and works, we can better understand the magnitude and significance of his innovative method and its applicability in modern times and also the significance of his many contributions. © The Author(s) 2013 Reprints and permissions: sagepub.co.uk/journalsPermissions.nav.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70031927','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70031927"><span>A deep crustal fluid channel into the San <span class="hlt">Andreas</span> Fault system near Parkfield, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Becken, M.; Ritter, O.; Park, S.K.; Bedrosian, P.A.; Weckmann, U.; Weber, M.</p> <p>2008-01-01</p> <p>Magnetotelluric (MT) data from 66 sites along a 45-km-long profile across the San <span class="hlt">Andreas</span> Fault (SAF) were inverted to obtain the 2-D electrical resistivity structure of the crust near the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD). The most intriguing feature of the resistivity model is a steeply dipping upper crustal high-conductivity zone flanking the seismically defined SAF to the NE, that widens into the lower crust and appears to be connected to a broad conductivity anomaly in the upper mantle. Hypothesis tests of the inversion model suggest that upper and lower crustal and upper-mantle anomalies may be interconnected. We speculate that the high conductivities are caused by fluids and may represent a deep-rooted channel for crustal and/or mantle fluid ascent. Based on the chemical analysis of well waters, it was previously suggested that fluids can enter the brittle regime of the SAF system from the lower crust and mantle. At high pressures, these fluids can contribute to fault-weakening at seismogenic depths. These geochemical studies predicted the existence of a deep fluid source and a permeable pathway through the crust. Our resistivity model images a conductive pathway, which penetrates the entire crust, in agreement with the geochemical interpretation. However, the resistivity model also shows that the upper crustal branch of the high-conductivity zone is located NE of the seismically defined SAF, suggesting that the SAF does not itself act as a major fluid pathway. This interpretation is supported by both, the location of the upper crustal high-conductivity zone and recent studies within the SAFOD main hole, which indicate that pore pressures within the core of the SAF zone are not anomalously high, that mantle-derived fluids are minor constituents to the fault-zone fluid composition and that both the volume of mantle fluids and the fluid pressure increase to the NE of the SAF. We further infer from the MT model that the resistive Salinian block</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70014341','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70014341"><span>Retardations in fault creep rates before local moderate earthquakes along the San <span class="hlt">Andreas</span> fault system, central California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Burford, R.O.</p> <p>1988-01-01</p> <p>Records of shallow aseismic slip (fault creep) obtained along parts of the San <span class="hlt">Andreas</span> and Calaveras faults in central California demonstrate that significant changes in creep rates often have been associated with local moderate earthquakes. An immediate postearthquake increase followed by gradual, long-term decay back to a previous background rate is generally the most obvious earthquake effect on fault creep. This phenomenon, identified as aseismic afterslip, usually is characterized by above-average creep rates for several months to a few years. In several cases, minor step-like movements, called coseismic slip events, have occurred at or near the times of mainshocks. One extreme case of coseismic slip, recorded at Cienega Winery on the San <span class="hlt">Andreas</span> fault 17.5 km southeast of San Juan Bautista, consisted of 11 mm of sudden displacement coincident with earthquakes of ML=5.3 and ML=5.2 that occurred 2.5 minutes apart on 9 April 1961. At least one of these shocks originated on the main fault beneath the winery. Creep activity subsequently stopped at the winery for 19 months, then gradually returned to a nearly steady rate slightly below the previous long-term average. The phenomena mentioned above can be explained in terms of simple models consisting of relatively weak material along shallow reaches of the fault responding to changes in load imposed by sudden slip within the underlying seismogenic zone. In addition to coseismic slip and afterslip phenomena, however, pre-earthquake retardations in creep rates also have been observed. Onsets of significant, persistent decreases in creep rates have occurred at several sites 12 months or more before the times of moderate earthquakes. A 44-month retardation before the 1979 ML=5.9 Coyote Lake earthquake on the Calaveras fault was recorded at the Shore Road creepmeter site 10 km northwest of Hollister. Creep retardation on the San <span class="hlt">Andreas</span> fault near San Juan Bautista has been evident in records from one creepmeter site for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988PApGe.126..499B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988PApGe.126..499B"><span>Retardations in fault creep rates before local moderate earthquakes along the San <span class="hlt">Andreas</span> fault system, central California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burford, Robert O.</p> <p>1988-06-01</p> <p>Records of shallow aseismic slip (fault creep) obtained along parts of the San <span class="hlt">Andreas</span> and Calaveras faults in central California demonstrate that significant changes in creep rates often have been associated with local moderate earthquakes. An immediate postearthquake increase followed by gradual, long-term decay back to a previous background rate is generally the most obvious earthquake effect on fault creep. This phenomenon, identified as aseismic afterslip, usually is characterized by above-average creep rates for several months to a few years. In several cases, minor step-like movements, called coseismic slip events, have occurred at or near the times of mainshocks. One extreme case of coseismic slip, recorded at Cienega Winery on the San <span class="hlt">Andreas</span> fault 17.5 km southeast of San Juan Bautista, consisted of 11 mm of sudden displacement coincident with earthquakes of M L =5.3 and M L =5.2 that occurred 2.5 minutes apart on 9 April 1961. At least one of these shocks originated on the main fault beneath the winery. Creep activity subsequently stopped at the winery for 19 months, then gradually returned to a nearly steady rate slightly below the previous long-term average. The phenomena mentioned above can be explained in terms of simple models consisting of relatively weak material along shallow reaches of the fault responding to changes in load imposed by sudden slip within the underlying seismogenic zone. In addition to coseismic slip and afterslip phenomena, however, pre-earthquake retardations in creep rates also have been observed. Onsets of significant, persistent decreases in creep rates have occurred at several sites 12 months or more before the times of moderate earthquakes. A 44-month retardation before the 1979 M L =5.9 Coyote Lake earthquake on the Calaveras fault was recorded at the Shore Road creepmeter site 10 km northwest of Hollister. Creep retardation on the San <span class="hlt">Andreas</span> fault near San Juan Bautista has been evident in records from one creepmeter</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70032596','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70032596"><span>Broadband simulations for Mw 7.8 southern san <span class="hlt">andreas</span> earthquakes: Ground motion sensitivity to rupture speed</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Graves, R.W.; Aagaard, Brad T.; Hudnut, K.W.; Star, L.M.; Stewart, J.P.; Jordan, T.H.</p> <p>2008-01-01</p> <p>Using the high-performance computing resources of the Southern California Earthquake Center, we simulate broadband (0-10 Hz) ground motions for three Mw 7.8 rupture scenarios of the southern San <span class="hlt">Andreas</span> fault. The scenarios incorporate a kinematic rupture description with the average rupture speed along the large slip portions of the fault set at 0.96, 0.89, and 0.84 times the local shear wave velocity. Consistent with previous simulations, a southern hypocenter efficiently channels energy into the Los Angeles region along the string of basins south of the San Gabriel Mountains. However, we find the basin ground motion levels are quite sensitive to the prescribed rupture speed, with peak ground velocities at some sites varying by over a factor of two for variations in average rupture speed of about 15%. These results have important implications for estimating seismic hazards in Southern California and emphasize the need for improved understanding of earthquake rupture processes. Copyright 2008 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70026946','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70026946"><span>Fine-scale structure of the San <span class="hlt">Andreas</span> fault zone and location of the SAFOD target earthquakes</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thurber, C.; Roecker, S.; Zhang, H.; Baher, S.; Ellsworth, W.</p> <p>2004-01-01</p> <p>We present results from the tomographic analysis of seismic data from the Parkfield area using three different inversion codes. The models provide a consistent view of the complex velocity structure in the vicinity of the San <span class="hlt">Andreas</span>, including a sharp velocity contrast across the fault. We use the inversion results to assess our confidence in the absolute location accuracy of a potential target earthquake. We derive two types of accuracy estimates, one based on a consideration of the location differences from the three inversion methods, and the other based on the absolute location accuracy of "virtual earthquakes." Location differences are on the order of 100-200 m horizontally and up to 500 m vertically. Bounds on the absolute location errors based on the "virtual earthquake" relocations are ??? 50 m horizontally and vertically. The average of our locations places the target event epicenter within about 100 m of the SAF surface trace. Copyright 2004 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70026734','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70026734"><span>Re-evaluation of heat flow data near Parkfield, CA: Evidence for a weak San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fulton, P.M.; Saffer, D.M.; Harris, Reid N.; Bekins, B.A.</p> <p>2004-01-01</p> <p>Improved interpretations of the strength of the San <span class="hlt">Andreas</span> Fault near Parkfield, CA based on thermal data require quantification of processes causing significant scatter and uncertainty in existing heat flow data. These effects include topographic refraction, heat advection by topographically-driven groundwater flow, and uncertainty in thermal conductivity. Here, we re-evaluate the heat flow data in this area by correcting for full 3-D terrain effects. We then investigate the potential role of groundwater flow in redistributing fault-generated heat, using numerical models of coupled heat and fluid flow for a wide range of hydrologic scenarios. We find that a large degree of the scatter in the data can be accounted for by 3-D terrain effects, and that for plausible groundwater flow scenarios frictional heat generated along a strong fault is unlikely to be redistributed by topographically-driven groundwater flow in a manner consistent with the 3-D corrected data. Copyright 2004 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003AGUFM.S41B..06W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003AGUFM.S41B..06W"><span>Heat Flow in the SAFOD Pilot Hole and Implications for the Strength of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Williams, C. F.; Grubb, F. V.; Galanis, S. P.</p> <p>2003-12-01</p> <p>As part of an investigation into the physical properties of the San <span class="hlt">Andreas</span> fault (SAF) and adjacent crust, detailed thermal measurements have been acquired in the 2.2-km-deep pilot hole for the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD), located 1.8 km west of the SAF near Parkfield, California. Precision temperature logs have been combined with thermal conductivity measurements on drill cuttings in a detailed vertical profile of heat flow. The temperature at the bottom of the borehole is 92 ° C, and heat flow from the basement section of the borehole (770 to 2160 m) is 91+/-2 mW m-2. Within the resolution of the measurements, heat flow is constant across the identified faults that intersect the borehole, suggesting that any active fluid flow along these faults is at rates too low to alter the background conductive thermal regime. Heat flow in the SAFOD pilot hole is significantly higher than the 74 mW m-2 average for the Parkfield area reported by Sass et al. (JGR, v. 102, 1997) based on measurements in shallow holes but consistent with five measurements ranging from 84 to 100 mW m-2 near the SAF in Pancho Rico Canyon 20 km to the northwest. Reanalysis of the regional heat flow pattern indicates that high heat flow at the SAFOD site reflects an abrupt increase in heat flow along the SAF and within the Coast Ranges northwest of Parkfield. This transition corresponds to a shallowing of the base of seismicity on the SAF and may be related to a change in the mechanical behavior of the fault near the northern terminus of the M6 1966 Parkfield earthquake rupture. The persistence of elevated heat flow at sites more than 40 km west of the SAFOD pilot hole appears to rule out frictional heating on the SAF as a major source of the high SAFOD value. However, the correlation of along-strike variations in heat flow with changes in rupture patterns and fault characteristics may indicate a previously overlooked connection between laterally heterogeneous frictional</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..14.4968C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..14.4968C"><span>Numerical analysis of the creeping behavior of the S. <span class="hlt">Andrea</span> di Perarolo secondary landslide (Italian Eastern Alps)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cioli, C.; Genevois, R.; Iafelice, M.; Zorzi, L.</p> <p>2012-04-01</p> <p>The S. <span class="hlt">Andrea</span> landslide is a complex secondary phenomenon characterized by continuous movements causing a very high hazard condition for the near Perarolo di Cadore village (Italian Eastern Alps). A significant amount of geological and geotechnical investigations has been carried out in the past allowing the detection of the basal sliding surface. In specific, the sliding surface coincides with the contact between the bedrock and the overlying mass of an old landslides, involving a volume of about 180.000 cubic meters. A numerical approach has been adopted to analyze the stability of slope. This method is able to simulate the formation and development of shear zones as areas of strain localization in the model. Indeed, the S. <span class="hlt">Andrea</span> landslide has been, then, investigated using FLAC, a two-dimensional explicit finite difference program, particularly useful in case of slopes with complex geometry. In order to build up a suitable model, variation of geological, hydrogeological and geotechnical parameters have been identified from the interpretation of all available data. In a preliminary stage, a Mohr-Coulomb plasticity model has been adopted except for the bedrock, which was characterized by an isotropic elastic model. Groundwater flow condition has been performed evaluating the change in pore pressure coupled to the mechanical deformation calculation. Numerical results show that this model cannot simulate real displacement behavior of the slope mainly due to both the complex material behavior and lithological heterogeneity, and due to geotechnical spatial complexity of different soils and mechanical parameters. It has been assumed that it was necessary to improve the model in the light of a time dependent behavior of existing soils. An elastic-viscoplastic model has been then used to reproduce the observed creeping behavior, and only in viscoplastic region time effects have been considered. Discussion of results points out on: i) the evolution of the ``mechanical</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70189779','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70189779"><span>Slip rates and spatially variable creep on faults of the northern San <span class="hlt">Andreas</span> system inferred through Bayesian inversion of Global Positioning System data</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Murray, Jessica R.; Minson, Sarah E.; Svarc, Jerry L.</p> <p>2014-01-01</p> <p>Fault creep, depending on its rate and spatial extent, is thought to reduce earthquake hazard by releasing tectonic strain aseismically. We use Bayesian inversion and a newly expanded GPS data set to infer the deep slip rates below assigned locking depths on the San <span class="hlt">Andreas</span>, Maacama, and Bartlett Springs Faults of Northern California and, for the latter two, the spatially variable interseismic creep rate above the locking depth. We estimate deep slip rates of 21.5 ± 0.5, 13.1 ± 0.8, and 7.5 ± 0.7 mm/yr below 16 km, 9 km, and 13 km on the San <span class="hlt">Andreas</span>, Maacama, and Bartlett Springs Faults, respectively. We infer that on average the Bartlett Springs fault creeps from the Earth's surface to 13 km depth, and below 5 km the creep rate approaches the deep slip rate. This implies that microseismicity may extend below the locking depth; however, we cannot rule out the presence of locked patches in the seismogenic zone that could generate moderate earthquakes. Our estimated Maacama creep rate, while comparable to the inferred deep slip rate at the Earth's surface, decreases with depth, implying a slip deficit exists. The Maacama deep slip rate estimate, 13.1 mm/yr, exceeds long-term geologic slip rate estimates, perhaps due to distributed off-fault strain or the presence of multiple active fault strands. While our creep rate estimates are relatively insensitive to choice of model locking depth, insufficient independent information regarding locking depths is a source of epistemic uncertainty that impacts deep slip rate estimates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024510','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024510"><span>Response of the San <span class="hlt">Andreas</span> fault to the 1983 Coalinga-Nuñez earthquakes: an application of interaction-based probabilities for Parkfield</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Toda, Shinji; Stein, Ross S.</p> <p>2002-01-01</p> <p>The Parkfield-Cholame section of the San <span class="hlt">Andreas</span> fault, site of an unfulfilled earthquake forecast in 1985, is the best monitored section of the world's most closely watched fault. In 1983, the M = 6.5 Coalinga and M = 6.0 Nuñez events struck 25 km northeast of Parkfield. Seismicity rates climbed for 18 months along the creeping section of the San <span class="hlt">Andreas</span> north of Parkfield and dropped for 6 years along the locked section to the south. Right-lateral creep also slowed or reversed from Parkfield south. Here we calculate that the Coalinga sequence increased the shear and Coulomb stress on the creeping section, causing the rate of small shocks to rise until the added stress was shed by additional slip. However, the 1983 events decreased the shear and Coulomb stress on the Parkfield segment, causing surface creep and seismicity rates to drop. We use these observations to cast the likelihood of a Parkfield earthquake into an interaction-based probability, which includes both the renewal of stress following the 1966 Parkfield earthquake and the stress transfer from the 1983 Coalinga events. We calculate that the 1983 shocks dropped the 10-year probability of a M ∼ 6 Parkfield earthquake by 22% (from 54 ± 22% to 42 ± 23%) and that the probability did not recover until about 1991, when seismicity and creep resumed. Our analysis may thus explain why the Parkfield earthquake did not strike in the 1980s, but not why it was absent in the 1990s. We calculate a 58 ± 17% probability of a M ∼ 6 Parkfield earthquake during 2001–2011.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2015/1147/ofr20151147_pamphlet.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2015/1147/ofr20151147_pamphlet.pdf"><span>Photomosaics and event evidence from the Frazier Mountain paleoseismic site, trench 1, cuts 5–24, San <span class="hlt">Andreas</span> Fault Zone, southern California (2010–2012)</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Scharer, Katherine M.; Fumal, Tom E.; Weldon, Ray J.; Streig, Ashley R.</p> <p>2015-08-24</p> <p>The Frazier Mountain paleoseismic site is located within the northern Big Bend of the southern San <span class="hlt">Andreas</span> Fault (lat 34.8122° N., lon 118.9034° W.), in a small structural basin formed by the fault (fig. 1). The site has been the focus of over a decade of paleoseismic study due to high stratigraphic resolution and abundant dateable material. Trench 1 (T1) was initially excavated as a 50-m long, fault-perpendicular trench crossing the northern half of the basin (Lindvall and others, 2002; Scharer and others, 2014a). Owing to the importance of a high-resolution trench site at this location on a 200-km length of the fault with no other long paleoseismic records, later work progressively lengthened and deepened T1 in a series of excavations, or cuts, that enlarged the original excavation. Scharer and others (2014a) provide the photomosaics and event evidence for the first four cuts, which largely show the upper section of the site, represented by alluvial deposits that date from about A.D. 1500 to present. Scharer and others (2014b) discuss the earthquake evidence and dating at the site within the context of prehistoric rupture lengths and magnitudes on the southern San <span class="hlt">Andreas</span> Fault. Here we present the photomosaics and event evidence for a series of cuts from the lower section, covering sediments that were deposited from about A.D. 500 to 1500 (fig. 2).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.T13C1395Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.T13C1395Z"><span>Noncharacteristic Slip on the Northern San <span class="hlt">Andreas</span> Fault at the Vedanta Marsh, Marin County, CA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, H.; Niemi, T. M.; Allison, A.; Fumal, T. E.</p> <p>2004-12-01</p> <p>Three-dimensional excavations along the 1906 trace of the northern San <span class="hlt">Andreas</span> fault at the Vedanta marsh paleoseismic site near Olema, CA have yielded new data on the timing and amount of slip during the penultimate earthquake on this fault section. The excavations exposed a 3-m-wide paleochannel that has been offset right-laterally 7.8-8.3 m by coseismic slip during the past two large earthquakes: 1906 and the penultimate earthquake. The paleochannel was eroded into a silty clay marsh deposit and was filled after AD 1400. Both the silty clay layer and the paleochannel deposit are directly overlain by an in situ burn/peat sequence. The penultimate earthquake occurred while the peat was at the ground surface whereas faulting from the 1906 earthquake terminates within an overlying gravel/fill sequence. Preliminary OxCal analyses of radiocarbon dates indicate that the penultimate earthquake occurred in the late 17th to early 18th century. In plan view, two main fault traces were mapped in the excavation. The northwestern portion of the paleochannel is offset across a single fault trace. Just southeast of this portion of the channel the fault splits into two traces. We believe that one of these traces likely slipped only during 1906 and the other trace slipped on during the penultimate earthquake. Unfortunately, the overlying stratigraphic section that could resolve the exact reconstruction of movement on these faults is missing due to the excavation of an artificial drainage ditch at this location in the 1940's. Matching the north margin of the paleochannel to the first exposure of gravel in the zone between the two fault traces gives an offset of 5 m. We have historic records that show the 1906 coseismic slip near the study site was about 5m from field notes of David Starr Jordan (Stanford University Archives) who describes two 16 ft (5m) offsets: one of a tree located about 150m SE of the offset channel and the other of a path to the Shafter barn located about 300m</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1986PApGe.124..127W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1986PApGe.124..127W"><span>Gravity anomaly and density structure of the San <span class="hlt">Andreas</span> fault zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Chi-Yuen; Rui, Feng; Zhengsheng, Yao; Xingjue, Shi</p> <p>1986-01-01</p> <p>A densely spaced gravity survey across the San <span class="hlt">andreas</span> fault zone was conducted near Bear Valley, about 180 km south of San Francisco, along a cross-section where a detailed seismic reflection profile was previously made by McEvilly (1981). With Feng and McEvilly's velocity structure (1983) of the fault zone at this cross-section as a constraint, the density structure of the fault zone is obtained through inversion of the gravity data by a method used by Parker (1973) and Oldenburg (1974). Although the resulting density picture cannot be unique, it is better constrained and contains more detailed information about the structure of the fault than was previously possible. The most striking feature of the resulting density structure is a deeply seated tongue of low-density material within the fault zone, probably representing a wedge of fault gouge between the two moving plates, which projects from the surface to the base of the seismogenic zone. From reasonable assumptions concerning the density of the solid grains and the state of saturation of the fault zone the average porosity of this low-density fault gouge is estimated as about 12%. Stress-induced cracks are not expected to create so much porosity under the pressures in the deep fault zone. Large-scaled removal of fault-zone material by hydrothermal alteration, dissolution, and subsequent fluid transport may have occurred to produce this pronounced density deficiency. In addition, a broad, funnel-shaped belt of low density appears about the upper part of the fault zone, which probably represents a belt of extensively shattered wall rocks.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70135097','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70135097"><span>Quaternary landscape development, alluvial fan chronology and erosion of the Mecca Hills at the southern end of the San <span class="hlt">Andreas</span> Fault zone</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Gray, Harrison J.; Owen, Lewis A.; Dietsch, Craig; Beck, Richard A.; Caffee, Marc A.; Finkelman, Robert B.; Mahan, Shannon</p> <p>2014-01-01</p> <p>Quantitative geomorphic analysis combined with cosmogenic nuclide 10Be-based geochronology and denudation rates have been used to further the understanding of the Quaternary landscape development of the Mecca Hills, a zone of transpressional uplift along the southern end of the San <span class="hlt">Andreas</span> Fault, in southern California. The similar timing of convergent uplifts along the San <span class="hlt">Andreas</span> Fault with the initiation of the sub-parallel San Jacinto Fault suggest a possible link between the two tectonic events. The ages of alluvial fans and the rates of catchment-wide denudation have been integrated to assess the relative influence of climate and tectonic uplift on the development of catchments within the Mecca Hills. Ages for major geomorphic surfaces based on 10Be surface exposure dating of boulders and 10Be depth profiles define the timing of surface stabilization to 2.6 +5.6/–1.3 ka (Qyf1 surface), 67.2 ± 5.3 ka (Qvof2 surface), and 280 ± 24 ka (Qvof1 surface). Comparison of 10Be measurements from active channel deposits (Qac) and fluvial terraces (Qt) illustrate a complex history of erosion, sediment storage, and sediment transport in this environment. Beryllium-10 catchment-wide denudation rates range from 19.9 ± 3.2 to 149 ± 22.5 m/Ma and demonstrate strong correlations with mean catchment slope and with total active fault length normalized by catchment area. The lack of strong correlation with other geomorphic variables suggests that tectonic uplift and rock weakening have the greatest control. The currently measured topography and denudation rates across the Mecca Hills may be most consistent with a model of radial topographic growth in contrast to a model based on the rapid uplift and advection of crust.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.G31A..06H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.G31A..06H"><span>Characterization of a Strain Rate Transient Along the San <span class="hlt">Andreas</span> and San Jacinto Faults Following the October 1999 Hector Mine Earthquake.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hernandez, D.; Holt, W. E.; Bennett, R. A.; Dimitrova, L.; Haines, A. J.</p> <p>2006-12-01</p> <p>We are continuing work on developing and refining a tool for recognizing strain rate transients as well as for quantifying the magnitude and style of their temporal and spatial variations. We determined time-averaged velocity values in 0.05 year epochs using time-varying velocity estimates for continuous GPS station data from the Southern California Integrated GPS Network (SCIGN) for the time period between October 1999 and February 2004 [Li et al., 2005]. A self-consistent model velocity gradient tensor field solution is determined for each epoch by fitting bi-cubic Bessel interpolation to the GPS velocity vectors and we determine model dilatation strain rates, shear strain rates, and the rotation rates. Departures of the time dependent model strain rate and velocity fields from a master solution, obtained from a time-averaged solution for the period 1999-2004, with imposed plate motion constraints and Quaternary fault data, are evaluated in order to best characterize the time dependent strain rate field. A particular problem in determining the transient strain rate fields is the level of smoothing or damping that is applied. Our current approach is to choose a damping that both maximizes the departure of the transient strain rate field from the long-term master solution and achieves a reduced chi-squared value between model and observed GPS velocities of around 1.0 for all time epochs. We observe several noteworthy time-dependent changes. First, in the Eastern California Shear Zone (ECSZ) region, immediately following the October 1999 Hector Mine earthquake, there occurs a significant spatial increase of relatively high shear strain rate, which encompasses a significant portion of the ECSZ. Second, also following the Hector Mine event, there is a strain rate corridor that extends through the Pinto Mt. fault connecting the ECSZ to the San <span class="hlt">Andreas</span> fault segment in the Salton Trough region. As this signal slowly decays, shear strain rates on segments of the San</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024275','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024275"><span>Evidence for large earthquakes on the San <span class="hlt">Andreas</span> fault at the Wrightwood, California paleoseismic site: A.D. 500 to present</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fumal, T.E.; Weldon, R.J.; Biasi, G.P.; Dawson, T.E.; Seitz, G.G.; Frost, W.T.; Schwartz, D.P.</p> <p>2002-01-01</p> <p>We present structural and stratigraphic evidence from a paleoseismic site near Wrightwood, California, for 14 large earthquakes that occurred on the southern San <span class="hlt">Andreas</span> fault during the past 1500 years. In a network of 38 trenches and creek-bank exposures, we have exposed a composite section of interbedded debris flow deposits and thin peat layers more than 24 m thick; fluvial deposits occur along the northern margin of the site. The site is a 150-m-wide zone of deformation bounded on the surface by a main fault zone along the northwest margin and a secondary fault zone to the southwest. Evidence for most of the 14 earthquakes occurs along structures within both zones. We identify paleoearthquake horizons using infilled fissures, scarps, multiple rupture terminations, and widespread folding and tilting of beds. Ages of stratigraphic units and earthquakes are constrained by historic data and 72 14C ages, mostly from samples of peat and some from plant fibers, wood, pine cones, and charcoal. Comparison of the long, well-resolved paleoseimic record at Wrightwood with records at other sites along the fault indicates that rupture lengths of past earthquakes were at least 100 km long. Paleoseismic records at sites in the Coachella Valley suggest that each of the past five large earthquakes recorded there ruptured the fault at least as far northwest as Wrightwood. Comparisons with event chronologies at Pallett Creek and sites to the northwest suggests that approximately the same part of the fault that ruptured in 1857 may also have failed in the early to mid-sixteenth century and several other times during the past 1200 years. Records at Pallett Creek and Pitman Canyon suggest that, in addition to the 14 earthquakes we document, one and possibly two other large earthquakes ruptured the part of the fault including Wrightwood since about A.D. 500. These observations and elapsed times that are significantly longer than mean recurrence intervals at Wrightwood and sites to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70037694','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70037694"><span>Locating non-volcanic tremor along the San <span class="hlt">Andreas</span> Fault using a multiple array source imaging technique</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ryberg, T.; Haberland, C.H.; Fuis, G.S.; Ellsworth, W.L.; Shelly, D.R.</p> <p>2010-01-01</p> <p>Non-volcanic tremor (NVT) has been observed at several subduction zones and at the San <span class="hlt">Andreas</span> Fault (SAF). Tremor locations are commonly derived by cross-correlating envelope-transformed seismic traces in combination with source-scanning techniques. Recently, they have also been located by using relative relocations with master events, that is low-frequency earthquakes that are part of the tremor; locations are derived by conventional traveltime-based methods. Here we present a method to locate the sources of NVT using an imaging approach for multiple array data. The performance of the method is checked with synthetic tests and the relocation of earthquakes. We also applied the method to tremor occurring near Cholame, California. A set of small-aperture arrays (i.e. an array consisting of arrays) installed around Cholame provided the data set for this study. We observed several tremor episodes and located tremor sources in the vicinity of SAF. During individual tremor episodes, we observed a systematic change of source location, indicating rapid migration of the tremor source along SAF. ?? 2010 The Authors Geophysical Journal International ?? 2010 RAS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23439700','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23439700"><span>Juan Valverde de Hamusco's unauthorized reproduction of a brain dissection by <span class="hlt">Andreas</span> Vesalius.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lanska, Douglas J; Lanska, John R</p> <p>2013-02-26</p> <p>The objective of the present work is to examine images of the brain dissection by Flemish-born anatomist <span class="hlt">Andreas</span> Vesalius (1514-1564) as originally represented in the Fabrica (1543), and later copied without Vesalius' permission by Spanish anatomist Juan Valverde de Hamusco (c1525-c1587) in Historia de la composicion del cuerpo humano (1556). Illustrations of the brain dissection in the Fabrica were obtained in digital form, resized, and arranged in a comparable montage to that presented by Valverde. Computer manipulations were used to assess image correspondence. The Valverde illustrations are approximately half the size and are mirror images of those in the Fabrica, but otherwise show the same dissection stages, and identical transverse brain levels and structures. The Valverde illustrations lack shadowing and show minor variations in perspective and fine details (e.g., branching pattern of the middle meningeal artery) from those in the Fabrica. Craftsmen under the direction of Valverde copied the woodcut prints in the Fabrica in close but approximate form by freehand engraving onto copper plates. Differences in the sizes of the images, and in perspective and fine detail, preclude direct tracing of images as the means of copying. Because engravings are in effect "flipped over" to make further prints, subsequent prints made from Valverde's copperplate engravings are mirror images of the prints in Vesalius' Fabrica.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFM.T43A1099T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFM.T43A1099T"><span>Paleomagnetic and Seismologic Evidence for Oblique-Slip Partitioning to the Coalinga Anticline From the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tetreault, J. L.; Jones, C. H.</p> <p>2007-12-01</p> <p>The Coalinga Anticline is a one of a series of fault-related folds in the central Coast Ranges, California, oriented subparallel to the San <span class="hlt">Andreas</span> Fault (SAF). The development of the Central Coast Range anticlines can be related to the relative strength of the SAF. If positing a weak SAF, fault-normal slip is partitioned to these subparallel compressional folds. If the SAF is strong, these folds rotated to their current orientation during wrenching. Another possibility is that the Coast Range anticlines are accommodating oblique-slip partitioned from the SAF. The 1983 Coalinga earthquake does not have a purely thrusting focal mechanism (rake =100°), reflecting the likelihood that oblique slip is being partitioned to this anticline, even though surface expression of fold-axis-parallel slip has not been identified. Paleomagnetic vertical-axis rotations and focal mechanism strain inversions were used to quantify oblique-slip deformation within the Coalinga Anticline. Clockwise rotations of 10° to 16° are inferred from paleomagnetic sites located in late Miocene to Pliocene beds on the steeply dipping forelimb and backlimb of the fold. Significant vertical-axis rotations are not identified in the paleomagnetic sites within the nose of the anticline. The varying vertical axis rotations conflict with wrench tectonics (strong SAF) as the mechanism of fold development. We use focal mechanisms inversions of earthquakes that occurred between 1983 to 2006 to constrain the seismogenic strain within the fold that presumably help to build it over time. In the upper 7 km, the principal shortening axis is oriented N37E to N40E, statistically indistinguishable from normal to the fold (N45E). The right-lateral shear in the folded strata above the fault tip, evident from the paleomagnetically determined clockwise vertical-axis rotations, is being accommodated aseismically or interseismically. In the region between 7 and 11 km, where the mainshock occurred, the shortening</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18807610','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18807610"><span>Leonardo da Vinci and <span class="hlt">Andreas</span> Vesalius; the shoulder girdle and the spine, a comparison.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ganseman, Y; Broos, P</p> <p>2008-01-01</p> <p>Leonardo Da Vinci and <span class="hlt">Andreas</span> Vesalius were two important renaissance persons; Vesalius was a surgeon-anatomist who delivered innovative work on the study of the human body, Leonardo da Vinci was an artist who delivered strikingly accurate and beautiful drawings on the human body. Below we compare both masters with regard to their knowledge of the working of the muscles, their method and system of dissection and their system and presentation of the drawings. The investigation consisted of a comparison between both anatomists, in particular concerning their study on the shoulder girdle and spine, by reviewing their original work as well as already existing literature on this subject. The investigation led to the conclusion that the drawings mentioned meant a change in history, and were of high quality, centuries ahead of their time. Both were anatomists, both were revolutionary, only one changed history at the moment itself, while the other changed history centuries later. Leonardo has made beautiful drawings that are at a match with the drawings of today or are even better. Vesalius set the start for medicine as a science as it is until this day. Their lives differed as strongly as their impact. In the light of their time, the achievement they made was extraordinary.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70127834','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70127834"><span>Paleoearthquakes at Frazier Mountain, California delimit extent and frequency of past San <span class="hlt">Andreas</span> Fault ruptures along 1857 trace</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Scharer, Katherine M.; Weldon, Ray; Streig, Ashley; Fumal, Thomas</p> <p>2014-01-01</p> <p>Large earthquakes are infrequent along a single fault, and therefore historic, well-characterized earthquakes exert a strong influence on fault behavior models. This is true of the 1857 Fort Tejon earthquake (estimated M7.7–7.9) on the southern San <span class="hlt">Andreas</span> Fault (SSAF), but an outstanding question is whether the 330 km long rupture was typical. New paleoseismic data for six to seven ground-rupturing earthquakes on the Big Bend of the SSAF restrict the pattern of possible ruptures on the 1857 stretch of the fault. In conjunction with existing sites, we show that over the last ~650 years, at least 75% of the surface ruptures are shorter than the 1857 earthquake, with estimated rupture lengths of 100 to <300 km. These results suggest that the 1857 rupture was unusual, perhaps leading to the long open interval, and that a return to pre-1857 behavior would increase the rate of M7.3–M7.7 earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/0698/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/0698/report.pdf"><span>Petrographic and chemical reconnaissance study of some granitic and gneissic rocks near the San <span class="hlt">Andreas</span> fault from Bodega Head to Cajon Pass, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ross, Donald C.</p> <p>1972-01-01</p> <p>This petrographic and chemical study is based on reconnaissance sampling of granitic and related gneissic rock in the California Coast and Transverse Ranges. In the Coast Ranges, granitic rocks are restricted to an elongate belt, the Salinian block, between the San <span class="hlt">Andreas</span> and Sur-Nacimiento fault zones. These rocks have a considerable compositional range, but are dominantly quartz monzonite and granodiorite. Moist of the Salinian block seems to be a structurally coherent basement block of chemically related granitic rocks. However, on both the east and the west sides of the block, gneiss crops out in abundance; these rocks may be structurally separate from the main part of the Salinian block. In the Transverse Ranges, the granitic and related rocks are dominantly of granodiorite composition, and in many areas granitic and gneissic rocks are intimately intermixed.Chemically the rocks of the California Coast and Transverse Ranges are somewhat intermediate in character between those of the east-central part of the Sierra Nevada batholith and those of the western part of the Sierra Nevada batholith and the southern California batholith. Probably the closest similarity is to the east-central Sierra Nevada rocks, but the rocks of the Coast and Transverse Ranges are somewhat higher in Al2O3 and lower in K2O than Sierran rocks of the comparable SiO2 content.Granitic basement rocks of the Salinian block are now anomalously sandwiched between Franciscan terranes. The petrographic and chemical data are compatible with the concept that the Salinian rocks were originally part of the great batholithic belt along the west coast, which is exemplified by the Sierra Nevada hatholith. It also seems most likely that the Salinian block was transported from somewhere south of the Sierra Nevada batholith by large-scale right-lateral movement along the San <span class="hlt">Andreas</span> fault zone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770006642','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770006642"><span>Monitoring of crustal movements in the San <span class="hlt">Andreas</span> fault zone by a satellite-borne ranging system. Ph.D. Thesis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kumar, M.</p> <p>1976-01-01</p> <p>The Close Grid Geodynamic Measurement System is conceived as an orbiting ranging device with a ground base grid of reflectors or transponders (spacing 1.0 to 30 km), which are projected to be of low cost (maintenance free and unattended), and which will permit the saturation of a local area to obtain data useful to monitor crustal movements in the San <span class="hlt">Andreas</span> fault zone. The system includes a station network of 75 stations covering an area between 36 deg N and 38 deg N latitudes, and 237 deg E and 239 deg E longitudes, with roughly half of the stations on either side of the faults. In addition, the simulation of crustal movements through the introduction of changes in the relative positions between grid stations, weather effect for intervisibility between satellite and station and loss of observations thereof, and comparative evaluation of various observational scheme-patterns have been critically studied.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70028956','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70028956"><span>Seismic evidence for rock damage and healing on the San <span class="hlt">Andreas</span> fault associated with the 2004 M 6.0 Parkfield earthquake</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Li, Y.-G.; Chen, P.; Cochran, E.S.; Vidale, J.E.; Burdette, T.</p> <p>2006-01-01</p> <p>We deployed a dense linear array of 45 seismometers across and along the San <span class="hlt">Andreas</span> fault near Parkfield a week after the M 6.0 Parkfield earthquake on 28 September 2004 to record fault-zone seismic waves generated by aftershocks and explosions. Seismic stations and explosions were co-sited with our previous experiment conducted in 2002. The data from repeated shots detonated in the fall of 2002 and 3 months after the 2004 M 6.0 mainshock show ???1.0%-1.5% decreases in seismic-wave velocity within an ???200-m-wide zone along the fault strike and smaller changes (0.2%-0.5%) beyond this zone, most likely due to the coseismic damage of rocks during dynamic rupture in the 2004 M 6.0 earthquake. The width of the damage zone characterized by larger velocity changes is consistent with the low-velocity waveguide model on the San <span class="hlt">Andreas</span> fault, near Parkfield, that we derived from fault-zone trapped waves (Li et al., 2004). The damage zone is not symmetric but extends farther on the southwest side of the main fault trace. Waveform cross-correlations for repeated aftershocks in 21 clusters, with a total of ???130 events, located at different depths and distances from the array site show ???0.7%-1.1% increases in S-wave velocity within the fault zone in 3 months starting a week after the earthquake. The velocity recovery indicates that the damaged rock has been healing and regaining the strength through rigidity recovery with time, most likely . due to the closure of cracks opened during the mainshock. We estimate that the net decrease in seismic velocities within the fault zone was at least ???2.5%, caused by the 2004 M 6.0 Parkfield earthquake. The healing rate was largest in the earlier stage of the postmainshock healing process. The magnitude of fault healing varies along the rupture zone, being slightly larger for the healing beneath Middle Mountain, correlating well with an area of large mapped slip. The fault healing is most prominent at depths above ???7 km.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.S42B..08N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.S42B..08N"><span>Nonvolcanic Deep Tremors in the Transform Plate Bounding San <span class="hlt">Andreas</span> Fault Zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nadeau, R. M.; Dolenc, D.</p> <p>2004-12-01</p> <p>Recently, deep ( ˜ 20 to 40 km) nonvolcanic tremor activity has been observed on convergent plate boundaries in Japan and in the Cascadia region of North America (Obara, 2002; Rodgers and Dragert, 2003; Szeliga et al., 2004). Because of the abundance of available fluids from subduction processes in these convergent zones, fluids are believed to play an important role in the generation of the tremor activity. The transient rates of tremor activity in these regions are also observed to correlate either with the occurrence of larger earthquakes (Obara, 2002) or with geodetically determined transient creep events that release large amounts of strain energy deep beneath the locked Cascadia megathrust (M.M. Miller et al., 2002; Rodgers and Dragert, 2003; Szeliga et al., 2004). These associations suggest that nonvolcanic tremor activity may participate in a fundamental mode of deep moment release and in the triggering of large subduction zone events (Rodgers and Dragert, 2003). We report the discovery of deep ( ˜ 20 to 45 km) nonvolcanic tremor activity on the transform plate bounding San <span class="hlt">Andreas</span> Fault (SAF) in central California where, in contrast to subduction zones, long-term deformation directions are horizontal and fluid availability from subduction zone processes is absent. The source region of the SAF tremors lies beneath the epicentral region of the great 1857 magnitude (M) ˜ 8, Fort Tejon earthquake whose rupture zone is currently locked (Sieh, 1978). Activity rate transients of the tremors occurring since early 2001 also correlate with seismicity rate variations above the tremor source region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25375329','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25375329"><span>Satellite tagging and biopsy sampling of killer whales at subantarctic <span class="hlt">Marion</span> Island: effectiveness, immediate reactions and long-term responses.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Reisinger, Ryan R; Oosthuizen, W Chris; Péron, Guillaume; Cory Toussaint, Dawn; Andrews, Russel D; de Bruyn, P J Nico</p> <p>2014-01-01</p> <p>Remote tissue biopsy sampling and satellite tagging are becoming widely used in large marine vertebrate studies because they allow the collection of a diverse suite of otherwise difficult-to-obtain data which are critical in understanding the ecology of these species and to their conservation and management. Researchers must carefully consider their methods not only from an animal welfare perspective, but also to ensure the scientific rigour and validity of their results. We report methods for shore-based, remote biopsy sampling and satellite tagging of killer whales Orcinus orca at Subantarctic <span class="hlt">Marion</span> Island. The performance of these methods is critically assessed using 1) the attachment duration of low-impact minimally percutaneous satellite tags; 2) the immediate behavioural reactions of animals to biopsy sampling and satellite tagging; 3) the effect of researcher experience on biopsy sampling and satellite tagging; and 4) the mid- (1 month) and long- (24 month) term behavioural consequences. To study mid- and long-term behavioural changes we used multievent capture-recapture models that accommodate imperfect detection and individual heterogeneity. We made 72 biopsy sampling attempts (resulting in 32 tissue samples) and 37 satellite tagging attempts (deploying 19 tags). Biopsy sampling success rates were low (43%), but tagging rates were high with improved tag designs (86%). The improved tags remained attached for 26±14 days (mean ± SD). Individuals most often showed no reaction when attempts missed (66%) and a slight reaction-defined as a slight flinch, slight shake, short acceleration, or immediate dive-when hit (54%). Severe immediate reactions were never observed. Hit or miss and age-sex class were important predictors of the reaction, but the method (tag or biopsy) was unimportant. Multievent trap-dependence modelling revealed considerable variation in individual sighting patterns; however, there were no significant mid- or long-term changes following</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4222950','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4222950"><span>Satellite Tagging and Biopsy Sampling of Killer Whales at Subantarctic <span class="hlt">Marion</span> Island: Effectiveness, Immediate Reactions and Long-Term Responses</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Reisinger, Ryan R.; Oosthuizen, W. Chris; Péron, Guillaume; Cory Toussaint, Dawn; Andrews, Russel D.; de Bruyn, P. J. Nico</p> <p>2014-01-01</p> <p>Remote tissue biopsy sampling and satellite tagging are becoming widely used in large marine vertebrate studies because they allow the collection of a diverse suite of otherwise difficult-to-obtain data which are critical in understanding the ecology of these species and to their conservation and management. Researchers must carefully consider their methods not only from an animal welfare perspective, but also to ensure the scientific rigour and validity of their results. We report methods for shore-based, remote biopsy sampling and satellite tagging of killer whales Orcinus orca at Subantarctic <span class="hlt">Marion</span> Island. The performance of these methods is critically assessed using 1) the attachment duration of low-impact minimally percutaneous satellite tags; 2) the immediate behavioural reactions of animals to biopsy sampling and satellite tagging; 3) the effect of researcher experience on biopsy sampling and satellite tagging; and 4) the mid- (1 month) and long- (24 month) term behavioural consequences. To study mid- and long-term behavioural changes we used multievent capture-recapture models that accommodate imperfect detection and individual heterogeneity. We made 72 biopsy sampling attempts (resulting in 32 tissue samples) and 37 satellite tagging attempts (deploying 19 tags). Biopsy sampling success rates were low (43%), but tagging rates were high with improved tag designs (86%). The improved tags remained attached for 26±14 days (mean ± SD). Individuals most often showed no reaction when attempts missed (66%) and a slight reaction–defined as a slight flinch, slight shake, short acceleration, or immediate dive–when hit (54%). Severe immediate reactions were never observed. Hit or miss and age-sex class were important predictors of the reaction, but the method (tag or biopsy) was unimportant. Multievent trap-dependence modelling revealed considerable variation in individual sighting patterns; however, there were no significant mid- or long-term changes following</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.S12A..05P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.S12A..05P"><span>Tremor evidence for dynamically triggered creep events on the deep San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Peng, Z.; Shelly, D. R.; Hill, D. P.; Aiken, C.</p> <p>2010-12-01</p> <p>Deep tectonic tremor has been observed along major subduction zones and the San <span class="hlt">Andreas</span> fault (SAF) in central and southern California. It appears to reflect deep fault slip, and it is often seen to be triggered by small stresses, including passing seismic waves from large regional and teleseismic earthquakes. Here we examine tremor activity along the Parkfield-Cholame section of the SAF from mid-2001 to early 2010, scrutinizing its relationship with regional and teleseismic earthquakes. Based on similarities in the shape and timing of seismic waveforms, we conclude that triggered and ambient tremor share common sources and a common physical mechanism. Utilizing this similarity in waveforms, we detect tremor triggered by numerous large events, including previously unreported triggering from the recent 2009 Mw7.3 Honduras, 2009 Mw8.1 Samoa, and 2010 Mw8.8 Chile earthquakes at teleseismic distances, and the relatively small 2007 Mw5.4 Alum Rock and 2008 Mw5.4 Chino Hills earthquakes at regional distances. We also find multiple examples of systematic migration in triggered tremor, similar to ambient tremor migration episodes observed at other times. Because these episodes propagate much more slowly than the triggering waves, the migration likely reflects a small, triggered creep event. As with ambient tremor bursts, triggered tremor at times persists for multiple days, probably indicating a somewhat larger creep event. This activity provides a clear example of delayed dynamic triggering, with a mechanism perhaps also relevant for triggering of regular earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70128987','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70128987"><span>Using surface creep rate to infer fraction locked for sections of the San <span class="hlt">Andreas</span> fault system in northern California from alignment array and GPS data</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lienkaemper, James J.; McFarland, Forrest S.; Simpson, Robert W.; Caskey, S. John</p> <p>2014-01-01</p> <p>Surface creep rate, observed along five branches of the dextral San <span class="hlt">Andreas</span> fault system in northern California, varies considerably from one section to the next, indicating that so too may the depth at which the faults are locked. We model locking on 29 fault sections using each section’s mean long‐term creep rate and the consensus values of fault width and geologic slip rate. Surface creep rate observations from 111 short‐range alignment and trilateration arrays and 48 near‐fault, Global Positioning System station pairs are used to estimate depth of creep, assuming an elastic half‐space model and adjusting depth of creep iteratively by trial and error to match the creep observations along fault sections. Fault sections are delineated either by geometric discontinuities between them or by distinctly different creeping behaviors. We remove transient rate changes associated with five large (M≥5.5) regional earthquakes. Estimates of fraction locked, the ratio of moment accumulation rate to loading rate, on each section of the fault system provide a uniform means to inform source parameters relevant to seismic‐hazard assessment. From its mean creep rates, we infer the main branch (the San <span class="hlt">Andreas</span> fault) ranges from only 20%±10% locked on its central creeping section to 99%–100% on the north coast. From mean accumulation rates, we infer that four urban faults appear to have accumulated enough seismic moment to produce major earthquakes: the northern Calaveras (M 6.8), Hayward (M 6.8), Rodgers Creek (M 7.1), and Green Valley (M 7.1). The latter three faults are nearing or past their mean recurrence interval.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018410','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018410"><span>Crustal velocity field near the big bend of California's San <span class="hlt">Andreas</span> fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Snay, R.A.; Cline, M.W.; Philipp, C.R.; Jackson, D.D.; Feng, Y.; Shen, Z.-K.; Lisowski, M.</p> <p>1996-01-01</p> <p>We use geodetic data spanning the 1920-1992 interval to estimate the horizontal velocity field near the big bend segment of California's San <span class="hlt">Andreas</span> fault (SAF). More specifically, we estimate a horizontal velocity vector for each node of a two-dimensional grid that has a 15-min-by-15-min mesh and that extends between latitudes 34.0??N and 36.0??N and longitudes 117.5??W and 120.5??W. For this estimation process, we apply bilinear interpolation to transfer crustal deformation information from geodetic sites to the grid nodes. The data include over a half century of triangulation measurements, over two decades of repeated electronic distance measurements, a decade of repeated very long baseline interferometry measurements, and several years of Global Positioning System measurements. Magnitudes for our estimated velocity vectors have formal standard errors ranging from 0.7 to 6.8 mm/yr. Our derived velocity field shows that (1) relative motion associated with the SAF exceeds 30 mm/yr and is distributed on the Earth's surface across a band (> 100 km wide) that is roughly centered on this fault; (2) when velocities are expressed relative to a fixed North America plate, the motion within our primary study region has a mean orientation of N44??W ?? 2?? and the surface trace of the SAF is congruent in shape to nearby contours of constant speed yet this trace is oriented between 5?? and 10?? counterclockwise relative to these contours; and (3) large strain rates (shear rates > 150 nrad/yr and/or areal dilatation rates < -150 nstr/yr) exist near the Garlock fault, near the White Wolf fault, and in the Ventura basin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JSG....42..246H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JSG....42..246H"><span>A microstructural study of fault rocks from the SAFOD: Implications for the deformation mechanisms and strength of the creeping segment of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hadizadeh, Jafar; Mittempergher, Silvia; Gratier, Jean-Pierre; Renard, Francois; Di Toro, Giulio; Richard, Julie; Babaie, Hassan A.</p> <p>2012-09-01</p> <p>The San <span class="hlt">Andreas</span> Fault zone in central California accommodates tectonic strain by stable slip and microseismic activity. We study microstructural controls of strength and deformation in the fault using core samples provided by the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) including gouge corresponding to presently active shearing intervals in the main borehole. The methods of study include high-resolution optical and electron microscopy, X-ray fluorescence mapping, X-ray powder diffraction, energy dispersive X-ray spectroscopy, white light interferometry, and image processing. The fault zone at the SAFOD site consists of a strongly deformed and foliated core zone that includes 2-3 m thick active shear zones, surrounded by less deformed rocks. Results suggest deformation and foliation of the core zone outside the active shear zones by alternating cataclasis and pressure solution mechanisms. The active shear zones, considered zones of large-scale shear localization, appear to be associated with an abundance of weak phases including smectite clays, serpentinite alteration products, and amorphous material. We suggest that deformation along the active shear zones is by a granular-type flow mechanism that involves frictional sliding of microlithons along phyllosilicate-rich Riedel shear surfaces as well as stress-driven diffusive mass transfer. The microstructural data may be interpreted to suggest that deformation in the active shear zones is strongly displacement-weakening. The fault creeps because the velocity strengthening weak gouge in the active shear zones is being sheared without strong restrengthening mechanisms such as cementation or fracture sealing. Possible mechanisms for the observed microseismicity in the creeping segment of the SAF include local high fluid pressure build-ups, hard asperity development by fracture-and-seal cycles, and stress build-up due to slip zone undulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.T41A2869K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.T41A2869K"><span>Understanding strain transfer and basin evolution complexities in the Salton pull-apart basin near the Southern San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kell, A. M.; Sahakian, V. J.; Kent, G. M.; Driscoll, N. W.; Harding, A. J.; Baskin, R. L.; Barth, M.; Hole, J. A.; Stock, J. M.; Fuis, G. S.</p> <p>2015-12-01</p> <p>Active source seismic data in the Salton Sea provide insight into the complexity of the pull-apart system development. Seismic reflection data combined with tomographic cross sections give constraints on the timing of basin development and strain partitioning between the two dominant dextral faults in the region; the Imperial fault to the southwest and the Southern San <span class="hlt">Andreas</span> fault (SSAF) to the northeast. Deformation associated with this step-over appears young, having formed in the last 20-40 k.a. The complexity seen in the Salton Sea is similar to that seen in pull-apart basins worldwide. In the southern basin of the Salton Sea, a zone of transpression is noted near the southern termination of the San <span class="hlt">Andreas</span> fault, though this stress regime quickly transitions to a region of transtension in the northern reaches of the sea. The evolution seen in the basin architecture is likely related to a transition of the SSAF dying to the north, and giving way to youthful segments of the Brawley seismic zone and Imperial fault. Stratigraphic signatures seen in seismic cross-sections also reveal a long-term component of slip to the southwest on a fault 1-2 km west of the northeastern Salton Sea shoreline. Numerous lines of evidence, including seismic reflection data, high-resolution bathymetry within the Salton Sea, and folding patterns in the Borrego Formation to the east of the sea support an assertion of a previously unmapped fault, the Salton Trough fault (STF), parallel to the SAF and just offshore within the Salton Sea. Seismic observations are seen consistently within two datasets of varying vertical resolutions, up to depths of 4-5 km, suggesting that this fault strand is much longer-lived than the evolution seen in the southern sub-basin. The existence of the STF unifies discrepancies between the onshore seismic studies and data collected within the sea. The STF likely serves as the current bounding fault to the active pull-apart system, as it aligns with the "rung</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025730','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025730"><span>Shallow soil CO2 flow along the San <span class="hlt">Andreas</span> and Calaveras Faults, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lewicki, J.L.; Evans, William C.; Hilley, G.E.; Sorey, M.L.; Rogie, J.D.; Brantley, S.L.</p> <p>2003-01-01</p> <p>We evaluate a comprehensive soil CO2 survey along the San <span class="hlt">Andreas</span> fault (SAF) in Parkfield, and the Calaveras fault (CF) in Hollister, California, in the context of spatial and temporal variability, origin, and transport of CO2 in fractured terrain. CO2 efflux was measured within grids with portable instrumentation and continously with meteorological parameters at a fixed station, in both faulted and unfaulted areas. Spatial and temporal variability of surface CO2 effluxes was observed to be higher at faulted SAF and CF sites, relative to comparable background areas. However, ??13C (-23.3 to - 16.4???) and ??14C (75.5 to 94.4???) values of soil CO2 in both faulted and unfaulted areas are indicative of biogenic CO2, even though CO2 effluxes in faulted areas reached values as high as 428 g m-2 d-1. Profiles of soil CO2 concentration as a function of depth were measured at multiple sites within SAF and CF grids and repeatedly at two locations at the SAF grid. Many of these profiles suggest a surprisingly high component of advective CO2 flow. Spectral and correlation analysis of SAF CO2 efflux and meteorological parameter time series indicates that effects of wind speed variations on atmospheric air flow though fractures modulate surface efflux of biogenic CO2. The resulting areal patterns in CO2 effluxes could be erroneously attributed to a deep gas source in the absence of isotopic data, a problem that must be addressed in fault zone soil gas studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRB..123..457T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRB..123..457T"><span>Using Low-Frequency Earthquake Families on the San <span class="hlt">Andreas</span> Fault as Deep Creepmeters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thomas, A. M.; Beeler, N. M.; Bletery, Q.; Burgmann, R.; Shelly, D. R.</p> <p>2018-01-01</p> <p>The central section of the San <span class="hlt">Andreas</span> Fault hosts tectonic tremor and low-frequency earthquakes (LFEs) similar to subduction zone environments. LFEs are often interpreted as persistent regions that repeatedly fail during the aseismic shear of the surrounding fault allowing them to be used as creepmeters. We test this idea by using the recurrence intervals of individual LFEs within LFE families to estimate the timing, duration, recurrence interval, slip, and slip rate associated with inferred slow slip events. We formalize the definition of a creepmeter and determine whether this definition is consistent with our observations. We find that episodic families reflect surrounding creep over the interevent time, while the continuous families and the short time scale bursts that occur as part of the episodic families do not. However, when these families are evaluated on time scales longer than the interevent time these events can also be used to meter slip. A straightforward interpretation of episodic families is that they define sections of the fault where slip is distinctly episodic in well-defined slow slip events that slip 16 times the long-term rate. In contrast, the frequent short-term bursts of the continuous and short time scale episodic families likely do not represent individual creep events but rather are persistent asperities that are driven to failure by quasi-continuous creep on the surrounding fault. Finally, we find that the moment-duration scaling of our inferred creep events are inconsistent with the proposed linear moment-duration scaling. However, caution must be exercised when attempting to determine scaling with incomplete knowledge of scale.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G43B0943V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G43B0943V"><span>Looking for Off-Fault Deformation and Measuring Strain Accumulation During the Past 70 years on a Portion of the Locked San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vadman, M.; Bemis, S. P.</p> <p>2017-12-01</p> <p>Even at high tectonic rates, detection of possible off-fault plastic/aseismic deformation and variability in far-field strain accumulation requires high spatial resolution data and likely decades of measurements. Due to the influence that variability in interseismic deformation could have on the timing, size, and location of future earthquakes and the calculation of modern geodetic estimates of strain, we attempt to use historical aerial photographs to constrain deformation through time across a locked fault. Modern photo-based 3D reconstruction techniques facilitate the creation of dense point clouds from historical aerial photograph collections. We use these tools to generate a time series of high-resolution point clouds that span 10-20 km across the Carrizo Plain segment of the San <span class="hlt">Andreas</span> fault. We chose this location due to the high tectonic rates along the San <span class="hlt">Andreas</span> fault and lack of vegetation, which may obscure tectonic signals. We use ground control points collected with differential GPS to establish scale and georeference the aerial photograph-derived point clouds. With a locked fault assumption, point clouds can be co-registered (to one another and/or the 1.7 km wide B4 airborne lidar dataset) along the fault trace to calculate relative displacements away from the fault. We use CloudCompare to compute 3D surface displacements, which reflect the interseismic strain accumulation that occurred in the time interval between photo collections. As expected, we do not observe clear surface displacements along the primary fault trace in our comparisons of the B4 lidar data against the aerial photograph-derived point clouds. However, there may be small scale variations within the lidar swath area that represent near-fault plastic deformation. With large-scale historical photographs available for the Carrizo Plain extending back to at least the 1940s, we can potentially sample nearly half the interseismic period since the last major earthquake on this portion of</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70154742','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70154742"><span>Southern San <span class="hlt">Andreas</span> Fault seismicity is consistent with the Gutenberg-Richter magnitude-frequency distribution</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Page, Morgan T.; Felzer, Karen</p> <p>2015-01-01</p> <p>The magnitudes of any collection of earthquakes nucleating in a region are generally observed to follow the Gutenberg-Richter (G-R) distribution. On some major faults, however, paleoseismic rates are higher than a G-R extrapolation from the modern rate of small earthquakes would predict. This, along with other observations, led to formulation of the characteristic earthquake hypothesis, which holds that the rate of small to moderate earthquakes is permanently low on large faults relative to the large-earthquake rate (Wesnousky et al., 1983; Schwartz and Coppersmith, 1984). We examine the rate difference between recent small to moderate earthquakes on the southern San <span class="hlt">Andreas</span> fault (SSAF) and the paleoseismic record, hypothesizing that the discrepancy can be explained as a rate change in time rather than a deviation from G-R statistics. We find that with reasonable assumptions, the rate changes necessary to bring the small and large earthquake rates into alignment agree with the size of rate changes seen in epidemic-type aftershock sequence (ETAS) modeling, where aftershock triggering of large earthquakes drives strong fluctuations in the seismicity rates for earthquakes of all magnitudes. The necessary rate changes are also comparable to rate changes observed for other faults worldwide. These results are consistent with paleoseismic observations of temporally clustered bursts of large earthquakes on the SSAF and the absence of M greater than or equal to 7 earthquakes on the SSAF since 1857.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018229','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018229"><span>Three-dimensional simulations of ground motions in the San Bernardino Valley, California, for hypothetical earthquakes on the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Frankel, A.</p> <p>1993-01-01</p> <p>Three-dimensional finite difference simulations of elastic waves in the San Bernardino Valley were performed for two hypothetical earthquakes on the San <span class="hlt">Andreas</span> fault: a point source with moment magnitude M5 and an extended rupture with M6.5. A method is presented for incorporating a source with arbitrary focal mechanism in the grid. Synthetics from the 3-D simulations are compared with those derived from 2-D (vertical cross section) and 1-D (flat-layered) models. The synthetic seismograms from the 3-D and 2-D simulations exhibit large surface waves produced by conversion of incident S waves at the edge of the basin. Seismograms from the flat-layered model do not contain these converted surface waves and underestimate the duration of shaking. Maps of maximum ground velocities occur in localized portions of the basin. The location of the largest velocities changes with the rupture propagation direction. Contours of maximum shaking are also dependent on asperity positions and radiation pattern. -from Author</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1984PApGe.122..376S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1984PApGe.122..376S"><span>Anomalous hydrogen emissions from the San <span class="hlt">Andreas</span> fault observed at the Cienega Winery, central California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sato, Motoaki; Sutton, A. J.; McGee, K. A.</p> <p>1984-03-01</p> <p>We began continuous monitoring of H2 concentration in soil along the San <span class="hlt">Andreas</span> and Calaveras faults in central California in December 1980, using small H2/O2 fuel-cell sensors. Ten monitoring stations deployed to date have shown that anomalous H2 emissions take place occasionally in addition to diurnal changes. Among the ten sites, the Cienega Winery site has produced data that are characterized by very small diurnal changes, a stable baseline, and remarkably distinct spike-like H2 anomalies since its installation in July 1982. A major peak appeared on 1 10 November 1982, and another on 3 April 1983, and a medium peak on 1 November 1983. The occurrences of these peaks coincided with periods of very low seismicity within a radius of 50 km from the site. In order to methodically assess how these peaks are related to earthquakes, three H2 degassing models were examined. A plausible correlational pattern was obtained by using a model that (1) adopts a hemicircular spreading pattern of H2 along an incipient fracture plane from the hypocenter of an earthquake, (2) relies on the FeO-H2O reaction for H2 generation, and (3) relates the accumulated amount of H2 to the mass of serpentinization of underlying ophiolitic rocks; the mass was tentatively assumed to be proportional to the seismic energy of the earthquake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1992/4159/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1992/4159/report.pdf"><span>Hydrogeology, water quality, and potential for contamination of the Upper Floridan aquifer in the Silver Springs ground-water basin, central <span class="hlt">Marion</span> County, Florida</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Phelps, G.G.</p> <p>1994-01-01</p> <p>The Upper Floridan aquifer, composed of a thick sequence of very porous limestone and dolomite, is the principal source of water supply in the Silver Springs ground-water basin of central <span class="hlt">Marion</span> County, Florida. The karstic nature of the local geology makes the aquifer susceptible to contaminants from the land surface. Contaminants can enter the aquifer by seepage through surficial deposits and through sinkholes and drainage wells. Potential contaminants include agricultural chemicals, landfill leachates and petroleum products from leaking storage tanks and accidental spills. More than 560 sites of potential contamination sources were identified in the basin in 1990. Detailed investigation of four sites were used to define hydrologic conditions at representative sites. Ground-water flow velocities determined from dye trace studies ranged from about 1 foot per hour under natural flow conditions to about 10 feet per hour under pumping conditions, which is considerably higher than velocities estimated using Darcy's equation for steady-state flow in a porous medium. Water entering the aquifer through drainage wells contained bacteria, elevated concentrations of nutrients, manganese and zinc, and in places, low concentrations of organic compounds. On the basis of results from the sampling of 34 wells in 1989 and 1990, and from the sampling of water entering the Upper Floridan aquifer through drainage wells, there has been no widespread degradation of water quality in the study area. In an area of karst, particularly one in which fracture flow is significant, evaluating the effects from contaminants is difficult and special care is required when interpolating hydrogeologic data from regional studies to a specific. (USGS)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mo0509.photos.097273p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mo0509.photos.097273p/"><span>7. GENERAL VIEW. VIEW FROM SOUTH SIDE OF BROADWAY LOOKING ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>7. GENERAL VIEW. VIEW FROM SOUTH SIDE OF BROADWAY LOOKING NORTHEAST SHOWS PRESENT SENIOR CITIZENS HIGHRISE ON THE WEST AND <span class="hlt">MARION</span> COUNTY COURTHOUSE ON THE EAST - Mitchell-Anderson House, 1008 Broadway, Hannibal, <span class="hlt">Marion</span> County, MO</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AGUFM.G52A..11S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AGUFM.G52A..11S"><span>Interseismic Deformation on the San <span class="hlt">Andreas</span> Fault System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Segall, P.</p> <p>2001-12-01</p> <p> ˙ {s}{HRC } ~ 13 mm/yr, ˙ {s}CGV ~ 9 mm/yr. Adding the constraint that the coseismic slip in major Hayward and San <span class="hlt">Andreas</span> events not exceed 3.0 m and 7.0 m, respectively yields an optimal model with: H ~ 18 km, tR ~ 36 years, TSAF = 280 years, ˙ {s}SAF = 25 mm/yr, tHRC = 225 years, T{ HRC} = 276 years, and ˙ {s}{HRC } ~ 11 mm/yr, ˙ {s}CGV ~ 9 mm/yr. These estimates are in reasonable accord with independent paleoseismic results. The conclusion of this pilot study is that by combining the present day deformation field, post-1906 strain data, and geologic bounds on slip-rate and maximum earthquake slip, we can estimate parameters of considerable geophysical interest, including time since past quakes and average recurrence interval.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0464.photos.581236p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0464.photos.581236p/"><span>ENTRANCE ROAD FROM MAIN ENTRANCE GATE, WITH ADMINISTRATION BUILDING AT ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>ENTRANCE ROAD FROM MAIN ENTRANCE GATE, WITH ADMINISTRATION BUILDING AT LEFT, LOWER BARN AT CENTER AND UPPER BARN AT RIGHT BACKGROUND. VIEW TO SOUTH. - <span class="hlt">Marion</span> National Cemetery, 1700 East Thirty-eighth Street, <span class="hlt">Marion</span>, Grant County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/al0675.photos.006218p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/al0675.photos.006218p/"><span>1. Historic American Buildings Survey Copied by Alex Bush, Photographer, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>1. Historic American Buildings Survey Copied by Alex Bush, Photographer, March 15, 1935 OLD COLLEGE PAMPHLET. NOT COPYRIGHTED. FRONT AND SIDE VIEW S.E. (BEFORE ALTERATION). - <span class="hlt">Marion</span> Female Seminary, Monroe & Centreville Streets, <span class="hlt">Marion</span>, Perry County, AL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.S53C4544M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.S53C4544M"><span>Geometry and Pore Pressure Shape the Pattern of the Tectonic Tremors Activity on the Deep San <span class="hlt">Andreas</span> Fault with Periodic, Period-Multiplying Recurrence Intervals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mele Veedu, D.; Barbot, S.</p> <p>2014-12-01</p> <p>A never before recorded pattern of periodic, chaotic, and doubled, earthquake recurrence intervals was detected in the sequence of deep tectonic tremors of the Parkfield segment of the San <span class="hlt">Andreas</span> Fault (Shelly, 2010). These observations may be the most puzzling seismological observations of the last decade: The pattern was regularly oscillating with a period doubling of 3 and 6 days from mid-2003 until it was disrupted by the 2004 Mw 6.0 Parkfield earthquake. But by the end of 2007, the previous pattern resumed. Here, we assume that the complex dynamics of the tremors is caused by slip on a single asperity on the San <span class="hlt">Andreas</span> Fault with homogeneous friction properties. We developed a three-dimensional model based on the rate-and-state friction law with a single patch and simulated fault slip during all stages of the earthquake cycle using the boundary integral method of Lapusta & Liu (2009). We find that homogeneous penny-shaped asperities cannot induce the observed period doubling, and that the geometry itself of the velocity-weakening asperity is critical in enabling the characteristic behavior of the Parkfield tremors. We also find that the system is sensitive to perturbations in pore pressure, such that the ones induced by the 2004 Parkfield earthquake are sufficient to dramatically alter the dynamics of the tremors for two years, as observed by Shelly (2010). An important finding is that tremor magnitude is amplified more by macroscopic slip duration on the source asperity than by slip amplitude, indicative of a time-dependent process for the breakage of micro-asperities that leads to seismic emissions. Our simulated event duration is in the range of 25 to 150 seconds, closely comparable to the event duration of a typical Parkfield tectonic tremor. Our simulations reproduce the unique observations of the Parkfield tremor activity. This study vividly illustrates the critical role of geometry in shaping the dynamics of fault slip evolution on a seismogenic fault.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70035969','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70035969"><span>Precise location of San <span class="hlt">Andreas</span> Fault tremors near Cholame, California using seismometer clusters: Slip on the deep extension of the fault?</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Shelly, D.R.; Ellsworth, W.L.; Ryberg, T.; Haberland, C.; Fuis, G.S.; Murphy, J.; Nadeau, R.M.; Burgmann, R.</p> <p>2009-01-01</p> <p>We examine a 24-hour period of active San <span class="hlt">Andreas</span> Fault (SAF) tremor and show that this tremor is largely composed of repeated similar events. Utilizing this similarity, we locate the subset of the tremor with waveforms similar to an identified low frequency earthquake (LFE) "master template," located using P and S wave arrivals to be ???26 km deep. To compensate for low signal-to-noise, we estimate event-pair differential times at "clusters" of nearby stations rather than at single stations. We find that the locations form a near-linear structure in map view, striking parallel to the SAF and near the surface trace. Therefore, we suggest that at least a portion of the tremor occurs on the deep extension of the fault, likely reflecting shear slip, similar to subduction zone tremor. If so, the SAF may extend to the base of the crust, ???10 km below the deepest regular earthquakes on the fault. ?? 2009 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70193655','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70193655"><span>Constraints on the source parameters of low-frequency earthquakes on the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thomas, Amanda M.; Beroza, Gregory C.; Shelly, David R.</p> <p>2016-01-01</p> <p>Low-frequency earthquakes (LFEs) are small repeating earthquakes that occur in conjunction with deep slow slip. Like typical earthquakes, LFEs are thought to represent shear slip on crustal faults, but when compared to earthquakes of the same magnitude, LFEs are depleted in high-frequency content and have lower corner frequencies, implying longer duration. Here we exploit this difference to estimate the duration of LFEs on the deep San <span class="hlt">Andreas</span> Fault (SAF). We find that the M ~ 1 LFEs have typical durations of ~0.2 s. Using the annual slip rate of the deep SAF and the average number of LFEs per year, we estimate average LFE slip rates of ~0.24 mm/s. When combined with the LFE magnitude, this number implies a stress drop of ~104 Pa, 2 to 3 orders of magnitude lower than ordinary earthquakes, and a rupture velocity of 0.7 km/s, 20% of the shear wave speed. Typical earthquakes are thought to have rupture velocities of ~80–90% of the shear wave speed. Together, the slow rupture velocity, low stress drops, and slow slip velocity explain why LFEs are depleted in high-frequency content relative to ordinary earthquakes and suggest that LFE sources represent areas capable of relatively higher slip speed in deep fault zones. Additionally, changes in rheology may not be required to explain both LFEs and slow slip; the same process that governs the slip speed during slow earthquakes may also limit the rupture velocity of LFEs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUSM.S24A..08D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUSM.S24A..08D"><span>Ground Motion Prediction for M7+ scenarios on the San <span class="hlt">Andreas</span> Fault using the Virtual Earthquake Approach</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Denolle, M.; Dunham, E. M.; Prieto, G.; Beroza, G. C.</p> <p>2013-05-01</p> <p>There is no clearer example of the increase in hazard due to prolonged and amplified shaking in sedimentary, than the case of Mexico City in the 1985 Michoacan earthquake. It is critically important to identify what other cities might be susceptible to similar basin amplification effects. Physics-based simulations in 3D crustal structure can be used to model and anticipate those effects, but they rely on our knowledge of the complexity of the medium. We propose a parallel approach to validate ground motion simulations using the ambient seismic field. We compute the Earth's impulse response combining the ambient seismic field and coda-wave enforcing causality and symmetry constraints. We correct the surface impulse responses to account for the source depth, mechanism and duration using a 1D approximation of the local surface-wave excitation. We call the new responses virtual earthquakes. We validate the ground motion predicted from the virtual earthquakes against moderate earthquakes in southern California. We then combine temporary seismic stations on the southern San <span class="hlt">Andreas</span> Fault and extend the point source approximation of the Virtual Earthquake Approach to model finite kinematic ruptures. We confirm the coupling between source directivity and amplification in downtown Los Angeles seen in simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70014061','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70014061"><span>Anomalous hydrogen emissions from the San <span class="hlt">Andreas</span> fault observed at the Cienega Winery, central California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sato, M.; Sutton, A.J.; McGee, K.A.</p> <p>1985-01-01</p> <p>We began continuous monitoring of H2 concentration in soil along the San <span class="hlt">Andreas</span> and Calaveras faults in central California in December 1980, using small H2/O2 fuel-cell sensors. Ten monitoring stations deployed to date have shown that anomalous H2 emissions take place occasionally in addition to diurnal changes. Among the ten sites, the Cienega Winery site has produced data that are characterized by very small diurnal changes, a stable baseline, and remarkably distinct spike-like H2 anomalies since its installation in July 1982. A major peak appeared on 1-10 November 1982, and another on 3 April 1983, and a medium peak on 1 November 1983. The occurrences of these peaks coincided with periods of very low seismicity within a radius of 50 km from the site. In order to methodically assess how these peaks are related to earthquakes, three H2 degassing models were examined. A plausible correlational pattern was obtained by using a model that (1) adopts a hemicircular spreading pattern of H2 along an incipient fracture plane from the hypocenter of an earthquake, (2) relies on the FeO-H2O reaction for H2 generation, and (3) relates the accumulated amount of H2 to the mass of serpentinization of underlying ophiolitic rocks; the mass was tentatively assumed to be proportional to the seismic energy of the earthquake. ?? 1985 Birkha??user Verlag.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.T21A0543R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.T21A0543R"><span>Searching for geodetic transient slip signals along the Parkfield segment of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rousset, B.; Burgmann, R.</p> <p>2017-12-01</p> <p>The Parkfield section of the San <span class="hlt">Andreas</span> fault is at the transition between a segment locked since the 1857 Mw 7.9 Fort Tejon earthquake to its south and a creeping segment to the north. It is particularly well instrumented since it is the many previous studies have focused on studying the coseismic and postseismic phases of the two most recent earthquake cycles, the interseismic phase is exhibiting interesting dynamics at the down-dip edge of the seismogenic zone, characterized by a very large number of low frequency earthquakes (LFE) with different behaviors depending on location. Interseismic fault creep rates appear to vary over a wide range of spatial and temporal scales, from the Earth's surface to the base of crust. In this study, we take advantage of the dense Global Positioning System (GPS) network, with 77 continuous stations located within a circle of radius 80 km centered on Parkfield. We correct these time series for the co- and postseismic signals of the 2003 Mw 6.3 San Simeon and 2004 Mw 6.0 Parkfield earthquakes. We then cross-correlate the residual time series with synthetic slow-slip templates following the approach of Rousset et al. (2017). Synthetic tests with transient events contained in GPS time series with realistic noise show the limit of detection of the method. In the application with real GPS time series, the highest correlation amplitudes are compared with micro-seismicity rates, as well as tremor and LFE observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70034995','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70034995"><span>The ShakeOut scenario: A hypothetical Mw7.8 earthquake on the Southern San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Porter, K.; Jones, L.; Cox, D.; Goltz, J.; Hudnut, K.; Mileti, D.; Perry, S.; Ponti, D.; Reichle, M.; Rose, A.Z.; Scawthorn, C.R.; Seligson, H.A.; Shoaf, K.I.; Treiman, J.; Wein, A.</p> <p>2011-01-01</p> <p>In 2008, an earthquake-planning scenario document was released by the U.S. Geological Survey (USGS) and California Geological Survey that hypothesizes the occurrence and effects of a Mw7.8 earthquake on the southern San <span class="hlt">Andreas</span> Fault. It was created by more than 300 scientists and engineers. Fault offsets reach 13 m and up to 8 m at lifeline crossings. Physics-based modeling was used to generate maps of shaking intensity, with peak ground velocities of 3 m/sec near the fault and exceeding 0.5 m/sec over 10,000 km2. A custom HAZUS??MH analysis and 18 special studies were performed to characterize the effects of the earthquake on the built environment. The scenario posits 1,800 deaths and 53,000 injuries requiring emergency room care. Approximately 1,600 fires are ignited, resulting in the destruction of 200 million square feet of the building stock, the equivalent of 133,000 single-family homes. Fire contributes $87 billion in property and business interruption loss, out of the total $191 billion in economic loss, with most of the rest coming from shakerelated building and content damage ($46 billion) and business interruption loss from water outages ($24 billion). Emergency response activities are depicted in detail, in an innovative grid showing activities versus time, a new format introduced in this study. ?? 2011, Earthquake Engineering Research Institute.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.T21A2308W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.T21A2308W"><span>Evaluating the relationship between lateral slip and repeated fold deformation along a transtensive step-over on the San <span class="hlt">Andreas</span> fault at the Frazier Mountain site</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Weldon, R. J.; Streig, A. R.; Frazier Mountain SoSAFE Trenching Team</p> <p>2011-12-01</p> <p>Transtensive step-overs known as sags are among the most ubiquitous features of strike slip faults. These structures create closed depressions that collect sediment, are often wet and thus preserve organic material that can be used to date the thick and rapidly accumulating section. It is clear from historical ruptures that these depressions grow incrementally with each earthquake. We are developing methods to carefully document and separate individual folding events, and to relate the amount of sagging or folding to the amount of horizontal slip creating the sag, with the goal of generating slip per event chronologies. This will be useful as sags are often the best sites for preserving evidence of earthquake timing, and determining slip at these sites will eliminate the ambiguity inherent in tying earthquake age data from micro-stratigraphic sites to nearby undated sites with good micro-geomorphic slip evidence. We apply this approach to the Frazier Mountain site on the Southern San <span class="hlt">Andreas</span> fault where we integrate trenching, cone penetrometer testing (CPT), surveying, photomosaicing, B4 LiDAR data and GIS techniques to make a detailed 3D map of subsurface geology, fault traces and related folds across the site. These data are used to generate structure contour and isopach maps for key stratigraphic units in order to evaluate fold deformation of paleo-ground surfaces across a transtensional step-over on the San <span class="hlt">Andreas</span> fault. Approximately 20 trenches show the main active trace of the San <span class="hlt">Andreas</span> fault right stepping ~30 m over ~100 m along strike producing two small synclinal sags that dramatically thicken the stratigraphic section. The northwest sag is about 50 m long, 5 m wide, and the southwest sag measures 20 m long and about 8 m wide. The Frazier Mountain site has yielded good earthquake chronologies, and relationships between fold deformation and surface fault rupture for the last 6 earthquakes. We observe that the degree of sagging in the synclines varies</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..14..124D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..14..124D"><span>Mantle Sources Beneath the SW Indian Ridge - Remelting the African Superplume</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dick, H. J. B.; Zhou, H.</p> <p>2012-04-01</p> <p>The SW Indian Ridge runs some 7700 km from the Bouvet to the Rodgriguez Triple Junction, crossing over or near two postulated mantle plumes. The latter are associated with large oceanic rises where the ridge axis shoals dramatically in the vicinity of the mantle hotspot. The <span class="hlt">Marion</span> Rise, extends 3100 km from the Andrew Bain FZ to near the Rodriguez TJ, with an along axis rise of 5600-m to it crest north of <span class="hlt">Marion</span> Island. The rise has thin crust inferred on the basis of abundant exposures of mantle peridotites along its length. We suggest that this is the result of its sub-axial mantle source, which is a depleted residue originally emplaced by the African Superplume into the asthenosphere beneath southern Africa during the Karoo volcanic event ~185 Ma. Based on shallow mantle anisotropy, plate reconstructions, and hotspot traces, it now forms the mantle substrate for the SW Indian Ridge due to the breakup of Gondwanaland. The <span class="hlt">Marion</span> Rise is associated with <span class="hlt">Marion</span> Island, the present location of the <span class="hlt">Marion</span> Hotspot, some 256 km south of the modern ridge. This plume is a vestigial remnant of the African Superplume now imbedded in and centered on asthenospheric mantle derived from the Karoo event. Based on the numerous large offset fracture zones, which would dam sub-axial asthenospheric flow along the ridge, the low postulated flux of the <span class="hlt">Marion</span> plume, its off-axis position, and the thin crust along the ridge it is clear that the present day plume does not support the <span class="hlt">Marion</span> Rise. Instead, this must be supported isostatically by the underlying mantle residue of the Karoo event. The Bouvet Rise is much shorter than the <span class="hlt">Marion</span> Rise, extending ~664 km from the Conrad FZ on the American-Antarctic Ridge to the Shaka FZ on the SW Indian Ridge. It has ~3000-m of axial relief, peaking at Speiss Smt at Speiss Ridge: the last spreading segment of the SW Indian Ridge adjacent to the Bouvet TJ. Unlike the <span class="hlt">Marion</span> plume, Bouvet is ridge-centered, and much of its rise is likely</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMGP32A..05Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMGP32A..05Z"><span>Structural Mapping Along the Central San <span class="hlt">Andreas</span> Fault-zone Using Airborne Electromagnetics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zamudio, K. D.; Bedrosian, P.; Ball, L. B.</p> <p>2017-12-01</p> <p>Investigations of active fault zones typically focus on either surface expressions or the associated seismogenic zones. However, the largely aseismic upper kilometer can hold significant insight into fault-zone architecture, strain partitioning, and fault-zone permeability. Geophysical imaging of the first kilometer provides a link between surface fault mapping and seismically-defined fault zones and is particularly important in geologically complex regions with limited surface exposure. Additionally, near surface imaging can provide insight into the impact of faulting on the hydrogeology of the critical zone. Airborne electromagnetic (AEM) methods offer a unique opportunity to collect a spatially-large, detailed dataset in a matter of days, and are used to constrain subsurface resistivity to depths of 500 meters or more. We present initial results from an AEM survey flown over a 60 kilometer long segment of the central San <span class="hlt">Andreas</span> Fault (SAF). The survey is centered near Parkfield, California, the site of the SAFOD drillhole, which marks the transition between a creeping fault segment to the north and a locked zone to the south. Cross sections with a depth of investigation up to approximately 500 meters highlight the complex Tertiary and Mesozoic geology that is dismembered by the SAF system. Numerous fault-parallel structures are imaged across a more than 10 kilometer wide zone centered on the surface trace. Many of these features can be related to faults and folds within Plio-Miocene sedimentary rocks found on both sides of the fault. Northeast of the fault, rocks of the Mesozoic Franciscan and Great Valley complexes are extremely heterogeneous, with highly resistive volcanic rocks within a more conductive background. The upper 300 meters of a prominent fault-zone conductor, previously imaged to 1-3 kilometers depth by magnetotellurics, is restricted to a 20 kilometer long segment of the fault, but is up to 4 kilometers wide in places. Elevated fault</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.T42A..03S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.T42A..03S"><span>A New Correlation of Large Earthquakes Along the Southern San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Scharer, K. M.; Weldon, R. J.; Biasi, G. P.</p> <p>2010-12-01</p> <p>There are now three sites on the southern San <span class="hlt">Andreas</span> fault (SSAF) with records of 10 or more dated ground rupturing earthquakes (Frazier Mountain, Wrightwood and Pallett Creek) and at least seven other sites with 3-5 dated events. Numerous sites have related information including geomorphic offsets caused by 1 to a few earthquakes, a known amount of slip spanning a specific interval of time or number of earthquakes, or the number (but not necessarily the exact ages) of earthquakes in an interval of time. We use this information to construct a record of recent large earthquakes on the SSAF. Strongly overlapping C-14 age ranges, especially between closely spaced sites like Pallett Creek and Wrightwood on the Mojave segment and Thousand Palms, Indio, Coachella and Salt Creek on the southernmost 100 kms of the fault, and overlap between the more distant Frazier Mountain and Bidart Fan sites on the northernmost part of the fault suggest that the paleoseismic data are robust and can be explained by a relatively small number of events that span substantial portions of the fault. This is consistent with the extent of rupture of the two historic events (1857 was ~300 km long and 1812 was 100-200 km long); slip per event data that averages 3-5 m per event at most sites; and the long historical hiatus since 1857. While some sites have smaller offsets for individual events, correlation between sites suggests that many small offsets are near the end of long ruptures. While the long event series on the Mojave are quasi-periodic, individual intervals range about an order of magnitude, from a few decades up to ~200 years. This wide range of intervals and the apparent anti-slip predictable behavior of ruptures (small intervals are not followed by small events) suggest weak clustering or periods of time spanning multiple intervals when strain release is higher low lower than average. These properties defy the application of simple hazard analysis but need to be understood to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70189759','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70189759"><span>Inferring fault rheology from low-frequency earthquakes on the San <span class="hlt">Andreas</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Beeler, Nicholas M.; Thomas, Amanda; Bürgmann, Roland; Shelly, David R.</p> <p>2013-01-01</p> <p>Families of recurring low-frequency earthquakes (LFEs) within nonvolcanic tremor (NVT) on the San <span class="hlt">Andreas</span> fault in central California show strong sensitivity to shear stress induced by the daily tidal cycle. LFEs occur at all levels of the tidal shear stress and are in phase with the very small, ~400 Pa, stress amplitude. To quantitatively explain the correlation, we use a model from the existing literature that assumes the LFE sources are small, persistent regions that repeatedly fail during shear of a much larger scale, otherwise aseismically creeping fault zone. The LFE source patches see tectonic loading, creep of the surrounding fault which may be modulated by the tidal stress, and direct tidal loading. If the patches are small relative to the surrounding creeping fault then the stressing is dominated by fault creep, and if patch failure occurs at a threshold stress, then the resulting seismicity rate is proportional to the fault creep rate or fault zone strain rate. Using the seismicity rate as a proxy for strain rate and the tidal shear stress, we fit the data with possible fault rheologies that produce creep in laboratory experiments at temperatures of 400 to 600°C appropriate for the LFE source depth. The rheological properties of rock-forming minerals for dislocation creep and dislocation glide are not consistent with the observed fault creep because strong correlation between small stress perturbations and strain rate requires perturbation on the order of the ambient stress. The observed tidal modulation restricts ambient stress to be at most a few kilopascal, much lower than rock strength. A purely rate dependent friction is consistent with the observations only if the product of the friction rate dependence and effective normal stress is ~ 0.5 kPa. Extrapolating the friction rate strengthening dependence of phyllosilicates (talc) to depth would require the effective normal stress to be ~50 kPa, implying pore pressure is lithostatic. If the LFE</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060439p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060439p/"><span>5. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>5. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. PLOT PLAN; DRAWING 1R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0694.photos.060465p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0694.photos.060465p/"><span>3. Photograph of an historic photograph in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>3. Photograph of an historic photograph in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, Illinois Veterans Administration Medical Center. DINING HALL BUILDING #2, LOOKING NE; DATED JANUARY 31, 1942. - Veterans Administration Medical Center, Building No. 2, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060455p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060455p/"><span>16. Photograph of an historic photograph in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>16. Photograph of an historic photograph in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. HOSPITAL BUILDING #1, LOOKING NE; DATED NOVEMBER 30, 1941. - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70029429','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70029429"><span>Pliocene transpressional modification of depositional basins by convergent thrusting adjacent to the "Big Bend" of the San <span class="hlt">Andreas</span> fault: An example from Lockwood Valley, southern California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kellogg, K.S.; Minor, S.A.</p> <p>2005-01-01</p> <p>The "Big Bend" of the San <span class="hlt">Andreas</span> fault in the western Transverse Ranges of southern California is a left stepping flexure in the dextral fault system and has long been recognized as a zone of relatively high transpression compared to adjacent regions. The Lockwood Valley region, just south of the Big Bend, underwent a profound change in early Pliocene time (???5 Ma) from basin deposition to contraction, accompanied by widespread folding and thrusting. This change followed the recently determined initiation of opening of the northern Gulf of California and movement along the southern San <span class="hlt">Andreas</span> fault at about 6.1 Ma, with the concomitant formation of the Big Bend. Lockwood Valley occupies a 6-km-wide, fault-bounded structural basin in which converging blocks of Paleoproterozoic and Cretaceous crystalline basement and upper Oligocene and lower Miocene sedimentary rocks (Plush Ranch Formation) were thrust over Miocene and Pliocene basin-fill sedimentary rocks (in ascending order, Caliente Formation, Lockwood Clay, and Quatal Formation). All the pre-Quatal sedimentary rocks and most of the Pliocene Quatal Formation were deposited during a mid-Tertiary period of regional transtension in a crustal block that underwent little clockwise vertical-axis rotation as compared to crustal blocks to the south. Ensuing Pliocene and Quaternary transpression in the Big Bend region began during deposition of the poorly dated Quatal Formation and was marked by four converging thrust systems, which decreased the areal extent of the sedimentary basin and formed the present Lockwood Valley structural basin. None of the thrusts appears presently active. Estimated shortening across the center of the basin was about 30 percent. The fortnerly defined eastern Big Pine fault, now interpreted to be two separate, oppositely directed, contractional reverse or thrust faults, marks the northwestern structural boundary of Lockwood Valley. The complex geometry of the Lockwood Valley basin is similar</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMED21D3471R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMED21D3471R"><span>Geomorphic evidence of active tectonics in the San Gorgonio Pass region of the San <span class="hlt">Andreas</span> Fault system: an example of discovery-based research in undergraduate teaching</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reinen, L. A.; Yule, J. D.</p> <p>2014-12-01</p> <p>Student-conducted research in courses during the first two undergraduate years can increase learning and improve student self-confidence in scientific study, and is recommended for engaging and retaining students in STEM fields (PCAST, 2012). At Pomona College, incorporating student research throughout the geology curriculum tripled the number of students conducting research prior to their senior year that culminated in a professional conference presentation (Reinen et al., 2006). Here we present an example of discovery-based research in Neotectonics, a second-tier course predominantly enrolling first-and second-year students; describe the steps involved in the four week project; and discuss early outcomes of student confidence, engagement and retention. In the San Gorgonio Pass region (SGPR) in southern California, the San <span class="hlt">Andreas</span> fault undergoes a transition from predominantly strike-slip to a complex system of faults with significant dip-slip, resulting in diffuse deformation and raising the question of whether a large earthquake on the San <span class="hlt">Andreas</span> could propagate through the region (Yule, 2009). In spring 2014, seven students in the Neotectonics course conducted original research investigating quantifiable geomorphic evidence of tectonic activity in the SGPR. Students addressed questions of [1] unequal uplift in the San Bernardino Mountains, [2] fault activity indicated by stream knick points, [3] the role of fault style on mountain front sinuosity, and [4] characteristic earthquake slip determined via fault scarp degradation models. Students developed and revised individual projects, collaborated with each other on methods, and presented results in a public forum. A final class day was spent reviewing the projects and planning future research directions. Pre- and post-course surveys show increases in students' self-confidence in the design, implementation, and presentation of original scientific inquiries. 5 of 6 eligible students participated in research the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020131','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020131"><span>Fault-zone guided waves from explosions in the San <span class="hlt">Andreas</span> fault at Parkfield and Cienega Valley, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Li, Y.-G.; Ellsworth, W.L.; Thurber, C.H.; Malin, P.E.; Aki, K.</p> <p>1997-01-01</p> <p>Fault-zone guided waves were successfully excited by near-surface explosions in the San <span class="hlt">Andreas</span> fault zone both at Parkfield and Cienega Valley, central California. The guided waves were observed on linear, three-component seismic arrays deployed across the fault trace. These waves were not excited by explosions located outside the fault zone. The amplitude spectra of guided waves show a maximum peak at 2 Hz at Parkfield and 3 Hz at Cienega Valley. The guided wave amplitude decays sharply with observation distance from the fault trace. The explosion-excited fault-zone guided waves are similar to those generated by earthquakes at Parkfield but have lower frequencies and travel more slowly. These observations suggest that the fault-zone wave guide has lower seismic velocities as it approaches the surface at Parkfield. We have modeled the waveforms as S waves trapped in a low-velocity wave guide sandwiched between high-velocity wall rocks, resulting in Love-type fault-zone guided waves. While the results are nonunique, the Parkfield data are adequately fit by a shallow wave guide 170 m wide with an S velocity 0.85 km/sec and an apparent Q ??? 30 to 40. At Cienega Valley, the fault-zone wave guide appears to be about 120 m wide with an S velocity 0.7 km/sec and a Q ??? 30.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title40-vol18/pdf/CFR-2012-title40-vol18-sec81-318.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title40-vol18/pdf/CFR-2012-title40-vol18-sec81-318.pdf"><span>40 CFR 81.318 - Kentucky.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2012&page.go=Go">Code of Federal Regulations, 2012 CFR</a></p> <p></p> <p>2012-07-01</p> <p>... Madison Co. in Richmond X Rest of Madison Co X Magoffin County X <span class="hlt">Marion</span> County X Marshall County X Martin... X Magoffin County X <span class="hlt">Marion</span> County X Marshall County X Martin County X Mason County X Meade County X... Marshall County Unclassifiable/Attainment Martin County Unclassifiable/Attainment Mason County...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/wri/1985/4263/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/wri/1985/4263/report.pdf"><span>Transit losses and traveltimes for water-supply releases <span class="hlt">Marion</span> Lake during drought conditions, Cottonwood River, east-central Kansas</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Jordan, P.R.; Hart, R.J.</p> <p>1985-01-01</p> <p>A streamflow routing model was used to calculate the transit losses and traveltimes. Channel and aquifer characteristics, and the model control parameters, were estimated from available data and then verified to the extent possible by comparing model simulated streamflow to observed streamflow at streamflow gaging stations. Transit losses and traveltimes for varying reservoir release rates and durations then were simulated for two different antecedent streamflow (drought) conditions. For the severe-drought antecedent-streamflow condition, it was assumed that only the downstream water use requirement would be released from the reservoir. For a less severe drought (LSD) antecedent streamflow condition, it was assumed than any releases from <span class="hlt">Marion</span> Lake for water supply use downstream, would be in addition to a nominal dry weather release of 5 cu ft/sec. Water supply release rates of 10 and 25 cu ft/sec for the severe drought condition and 5, 10, and 25 cu ft/sec for the less severe drought condition were simulated for periods of 28 and 183 days commencing on July 1. Transit losses for the severe drought condition for all reservoir release rates and durations ranged from 12% to 78% of the maximum downstream flow rate and from 27% to 91% of the total volume of reservoir storage released. For the LSD condition, transit losses ranged from 7% to 29% of the maximum downstream flow rate and from 10% to 48% of the total volume of release. The 183-day releases had larger total transit losses, but losses on a percentage basis were less than the losses for the 28-day release period for both antecedent streamflow conditions. Traveltimes to full response (80% of the maximum downstream flow rate), however, showed considerable variation. For the release of 5 cu ft/sec during LSD conditions, base flow exceeded 80% of the maximum flow rate near the confluence; the traveltime to full response was undefined for those simulations. For the releases of 10 and 25 cu ft/sec during the same</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED509726.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED509726.pdf"><span>Perspectives of Young Children: How Do They Really Think?</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Costley, Kevin C.</p> <p>2010-01-01</p> <p>In his monumental research, although Piaget primarily relayed information about children's developmental stages of cognitive growth, Marian <span class="hlt">Marion</span> goes on to discuss not only the developmental stages, yet focuses on how children think. In her textbook, "Guidance of Young Children", <span class="hlt">Marion</span> conveys how teachers need to understand children and help…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0697.photos.060475p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0697.photos.060475p/"><span>1. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>1. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, Illinois Veterans Administration Medical Center. (PLANS & ELEVATIONS); DRAWING 14-1R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Building No. 14, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060457p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060457p/"><span>18. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>18. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. SIDE (EAST) ELEVATION & SIDE (WEST) ELEVATION; DRAWING 1-9R, DATED SEPTEMBER 4, 1940. - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060454p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060454p/"><span>15. Photograph of an historic photograph in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>15. Photograph of an historic photograph in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. HOSPITAL BUILDING #1; E & N (FRONT) ELEVATIONS, LOOKING SW; DATED MAY 18, 1942. - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060461p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060461p/"><span>22. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>22. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. MAIN LOBBY & ELEVATOR LOBBY DETAILS; DRAWING 1-20R, DATED SEPTEMBER 4, 1940. - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol17/pdf/CFR-2010-title40-vol17-sec81-325.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol17/pdf/CFR-2010-title40-vol17-sec81-325.pdf"><span>40 CFR 81.325 - Mississippi.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-07-01</p> <p>... Lee County X Leflore County X Lincoln County X Lowndes County X Madison County X <span class="hlt">Marion</span> County X... Lee County X Leflore County X Lincoln County X Lowndes County X Madison County X <span class="hlt">Marion</span> County X... County Kemper County Lafayette County Lamar County Lauderdale County Lawrence County Leake County Lee...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-341.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-341.pdf"><span>40 CFR 81.341 - South Carolina.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... County X Laurens County X Lee County X Lexington County X McCormick County X <span class="hlt">Marion</span> County X Marlboro... X Jasper County X Kershaw County X Lancaster County X Laurens County X Lee County X Lexington County... County Kershaw County Lancaster County Laurens County Lee County Lexington County <span class="hlt">Marion</span> County Marlboro...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol17/pdf/CFR-2010-title40-vol17-sec81-341.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol17/pdf/CFR-2010-title40-vol17-sec81-341.pdf"><span>40 CFR 81.341 - South Carolina.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-07-01</p> <p>... County X Laurens County X Lee County X Lexington County X McCormick County X <span class="hlt">Marion</span> County X Marlboro... X Jasper County X Kershaw County X Lancaster County X Laurens County X Lee County X Lexington County... County Kershaw County Lancaster County Laurens County Lee County Lexington County <span class="hlt">Marion</span> County Marlboro...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-325.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-325.pdf"><span>40 CFR 81.325 - Mississippi.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... Lee County X Leflore County X Lincoln County X Lowndes County X Madison County X <span class="hlt">Marion</span> County X... Lee County X Leflore County X Lincoln County X Lowndes County X Madison County X <span class="hlt">Marion</span> County X... County Kemper County Lafayette County Lamar County Lauderdale County Lawrence County Leake County Lee...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70016318','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70016318"><span>Crustal strain near the Big Bend of the San <span class="hlt">Andreas</span> Fault: analysis of the Los Padres-Tehachapi Trilateration Networks, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Eberhart-Phillips, D.; Lisowski, M.</p> <p>1990-01-01</p> <p>In the region of the Los Padres-Tehachapi geodetic network, the San <span class="hlt">Andreas</span> fault (SAF) changes its orientation by over 30?? from N40??W, close to that predicted by plate motion for a transform boundary, to N73??W. The strain orientation near the SAF is consistent with right-lateral shear along the fault, with maximum shear rate of 0.38??0.01??rad/yr at N63??W. In contrast, away from the SAF the strain orientations on both sides of the fault are consistent with the plate motion direction, with maximum shear rate of 0.19??0.01??rad/yr at N44??W. The best fitting Garlock fault model had computed left-lateral slip rate of 11??2mm/yr below 10km. Buried left-lateral slip of 15??6mm/yr on the Big Pine fault, within the Western Transverse Ranges, provides significant reduction in line length residuals; however, deformation there may be more complicated than a single vertical fault. A subhorizontal detachment on the southern side of the SAF cannot be well constrained by these data. -from Authors</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060456p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060456p/"><span>17. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>17. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. FRONT (NORTH) ELEVATION; DRAWING 1-7R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0695.photos.060470p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0695.photos.060470p/"><span>4. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, Illinois Veterans Administration Medical Center. (PLANS, ELEVATIONS & DETAILS); DRAWING 8-1R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Building No. 8, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060462p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060462p/"><span>23. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>23. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. TYPICAL LOBBY DETAILS; DRAWING 1-21R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0696.photos.060474p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0696.photos.060474p/"><span>4. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. Photograph of a line drawing in the possession of the Engineer's office of the <span class="hlt">Marion</span>, Illinois Veterans Administration Medical Center. (PLANS, ELEVATIONS & DETAILS); DRAWING 13-1R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Building No. 13, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0698.photos.060476p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0698.photos.060476p/"><span>1. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>1. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, Illinois Veterans Administration Medical Center. (PLAN, ELEVATIONS, & SECTIONS); DRAWING 15-1R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Building No. 15, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title20-vol2/pdf/CFR-2011-title20-vol2-sec404-435.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title20-vol2/pdf/CFR-2011-title20-vol2-sec404-435.pdf"><span>20 CFR 404.435 - Excess earnings; months to which excess earnings can or cannot be charged; grace year defined.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-04-01</p> <p>... reason of having a child in his or her care), or to mother's or father's benefits is entitled to a...: <span class="hlt">Marion</span> was entitled to mother's insurance benefits from 1998 because she had a child in her care. Because... 2000, and the child's benefits terminated in April 2000. <span class="hlt">Marion</span>'s entitlement to mother's benefits also...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-10-06/pdf/2011-25871.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-10-06/pdf/2011-25871.pdf"><span>76 FR 61954 - Suspension of Community Eligibility</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-10-06</p> <p>... available in SFHAs Region IV South Carolina: <span class="hlt">Marion</span>, City of, <span class="hlt">Marion</span> County.. 450142 March 4, 1974, Emerg...... 260083 December 26, 1974, Emerg; March 1, 1982, ......do Do. Reg; October 18, 2011, Susp. St. Louis, City..., Van Buren 190266 January 14, 1976, Emerg; July 2, 1987, ......do Do. County. Reg; October 18, 2011...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009CoMP..157..173S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009CoMP..157..173S"><span>On the origin of mixed-layered clay minerals from the San <span class="hlt">Andreas</span> Fault at 2.5-3 km vertical depth (SAFOD drillhole at Parkfield, California)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schleicher, A. M.; Warr, L. N.; van der Pluijm, B. A.</p> <p>2009-02-01</p> <p>A detailed mineralogical study is presented of the matrix of mudrocks sampled from spot coring at three key locations along the San <span class="hlt">Andreas</span> Fault Observatory at depth (SAFOD) drill hole. The characteristics of authigenic illite-smectite (I-S) and chlorite-smectite (C-S) mixed-layer mineral clays indicate a deep diagenetic origin. A randomly ordered I-S mineral with ca. 20-25% smectite layers is one of the dominant authigenic clay species across the San <span class="hlt">Andreas</span> Fault zone (sampled at 3,066 and 3,436 m measured depths/MD), whereas an authigenic illite with ca. 2-5% smectite layers is the dominant phase beneath the fault (sampled at 3,992 m MD). The most smectite-rich mixed-layered assemblage with the highest water content occurs in the actively deforming creep zone at ca. 3,300-3,353 m (true vertical depth of ca. 2.7 km), with I-S (70:30) and C-S (50:50). The matrix of all mudrock samples show extensive quartz and feldspar (both plagioclase and K-feldspar) dissolution associated with the crystallization of pore-filling clay minerals. However, the effect of rock deformation in the matrix appears only minor, with weak flattening fabrics defined largely by kinked and fractured mica grains. Adopting available kinetic models for the crystallization of I-S in burial sedimentary environments and the current borehole depths and thermal structure, the conditions and timing of I-S growth can be evaluated. Assuming a typical K+ concentration of 100-200 ppm for sedimentary brines, a present-day geothermal gradient of 35°C/km and a borehole temperature of ca. 112°C for the sampled depths, most of the I-S minerals can be predicted to have formed over the last 4-11 Ma and are probably still in equilibrium with circulating fluids. The exception to this simple burial pattern is the occurrence of the mixed layered phases with higher smectite content than predicted by the burial model. These minerals, which characterize the actively creeping section of the fault and local thin film</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002AGUFM.T11F..11D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002AGUFM.T11F..11D"><span>Structural features of the San <span class="hlt">Andreas</span> fault at Tejon Pass, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dewers, T. A.; Reches, Z.; Brune, J. N.</p> <p>2002-12-01</p> <p>We mapped a 2 km belt along the San <span class="hlt">Andreas</span> fault (SAF) in the Tejon Pass area where road cuts provide fresh exposures of the fault zone and surrounding rocks. Our 1:2,000 structural mapping is focused on analysis of faulting processes and is complementary to regional mapping at 1:12,000 scale by Ramirez (M.Sc., UC Santa Barbara, 1984). The dominant rock units are the Hungry Valley Formation of Pliocene age (clastic sediments) exposed south of the SAF, and the Tejon Lookout granite (Cretaceous) and Neenach Volcanic Formation exposed north of it. Ramirez (1983) deduced ~220 km of post-Miocene lateral slip. The local trend of the SAF is about N60W and it includes at least three main, subparallel segments that form a 200 m wide zone. The traces of the segments are quasi-linear, discontinuous, and they are stepped with respect to each other, forming at least five small pull-aparts and sag ponds in the mapping area. The three segments were not active semi-contemporaneously and the southern segment is apparently the oldest. The largest pull-apart, 60-70 m wide, displays young (Quaternary?) silt and shale layers. We found two rock bodies that are suspected as fault-rocks. One is a 1-2 m thick sheet-like body that separates the Tejon Lookout granite from young (Recent?) clastic rocks. In the field, it appears as a gouge zone composed of poorly cemented, dark clay size grains; however, the microstructure of this rock does not reveal clear shear features. The second body is the 80-120 m wide zone of Tejon Lookout granite that extends for less than 1 km along the SAF in the mapped area. It is characterized by three structural features: (1) pulverization into friable, granular material by multitude of grain-crossing fractures; (2) abundance of dip-slip small faults that are gently dipping toward and away from the SAF; and (3) striking lack of evidence for shear parallel to the SAF. The relationships between these features and the large right-lateral shear along the SAF are</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.T12A..02L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.T12A..02L"><span>Geometry of the southern San <span class="hlt">Andreas</span> fault and its implications for seismic hazard</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Langenheim, V. E.; Dorsey, R. J.; Fuis, G. S.; Cooke, M. L.; Fattaruso, L.; Barak, S.</p> <p>2015-12-01</p> <p>The southern San <span class="hlt">Andreas</span> fault (SSAF) provides rich opportunities for studying the geometry and connectivity of fault stepovers and intersections, including recently recognized NE tilting of the Salton block between the SSAF and San Jacinto fault (SJF) that likely results from slight obliquity of relative plate motion to the strike of the SSAF. Fault geometry and predictions of whether the SSAF will rupture through the restraining bend in San Gorgonio Pass (SGP) are controversial, with significant implications for seismic hazard. The evolution of faulting in SGP has led to various models of strain accommodation, including clockwise rotation of fault-bounded blocks east of the restraining bend, and generation of faults that siphon strike slip away from the restraining bend onto the SJF (also parallel to the SSAF). Complex deformation is not restricted to the upper crust but extends to mid- and lower-crustal depths according to magnetic data and ambient-noise surface-wave tomography. Initiation of the SJF ~1.2 Ma led to formation of the relatively intact Salton block, and end of extension on the West Salton detachment fault on the west side of Coachella Valley. Geologic and geomorphic data show asymmetry of the southern Santa Rosa Mountains, with a steep fault-bounded SW flank produced by active uplift, and gentler topographic gradients on the NE flank with tilted, inactive late Pleistocene fans that are incised by modern upper fan channels. Gravity data indicate the basin floor beneath Coachella Valley is also asymmetric, with a gently NE-dipping basin floor bound by a steep SSAF; seismic-reflection data suggest that NE tilting took place during Quaternary time. 3D numerical modeling predicts gentle NE dips in the Salton block that result from the slight clockwise orientation of relative motion across a NE-dipping SSAF. A NE dip of the SSAF, supported by various geophysical datasets, would reduce shaking in Coachella Valley compared to a vertical fault.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060460p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060460p/"><span>21. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>21. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. 3/4' SCALE DETAIL - MAIN ENTRANCE UNIT, HALF ELEVATION & SECTION; DRAWING 1-15R, DATED SEPTEMBER 4, 1940. - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060458p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060458p/"><span>19. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>19. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. DETAILS MAIN STEPS & PLATFORM (SIDE ELEVATION, HALF PLAN, SECTIONS); DRAWING 1-12R, DATED SEPTEMBER 4, 1940. - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Food+AND+products+AND+research&id=EJ907608','ERIC'); return false;" href="https://eric.ed.gov/?q=Food+AND+products+AND+research&id=EJ907608"><span>Big Food, Big Agra, and the Research University</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Warren, Cat</p> <p>2010-01-01</p> <p>This article presents an interview with <span class="hlt">Marion</span> Nestle, the New York University food scientist and author of "Food Politics." <span class="hlt">Marion</span> Nestle is Paulette Goddard Professor in the Department of Nutrition, Food Studies, and Public Health at New York University. From 1986 to 1988, she was senior nutrition policy adviser in the Department of Health and…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA02733&hterms=time+perspective&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dtime%2Bperspective','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA02733&hterms=time+perspective&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dtime%2Bperspective"><span>Perspective view, Landsat overlay San <span class="hlt">Andreas</span> Fault, Palmdale, California</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2000-01-01</p> <p><p/> The prominent linear feature straight down the center of this perspective view is the San <span class="hlt">Andreas</span> Fault. This segment of the fault lies near the city of Palmdale, California (the flat area in the right half of the image) about 60 kilometers (37 miles) north of Los Angeles. The fault is the active tectonic boundary between the North American plate on the right, and the Pacific plate on the left. Relative to each other, the Pacific plate is moving away from the viewer and the North American plate is moving toward the viewer along what geologists call a right lateral strike-slip fault. Two large mountain ranges are visible, the San Gabriel Mountains on the left and the Tehachapi Mountains in the upper right. The Lake Palmdale Reservoir, approximately 1.5 kilometers (0.9 miles) across, sits in the topographic depression created by past movement along the fault. Highway 14 is the prominent linear feature starting at the lower left edge of the image and continuing along the far side of the reservoir. The patterns of residential and agricultural development around Palmdale are seen in the Landsat imagery in the right half of the image. SRTM topographic data will be used by geologists studying fault dynamics and landforms resulting from active tectonics.<p/>This type of display adds the important dimension of elevation to the study of land use and environmental processes as observed in satellite images. The perspective view was created by draping a Landsat satellite image over an SRTM elevation model. Topography is exaggerated 1.5 times vertically. The Landsat image was provided by the United States Geological Survey's Earth Resources Observations Systems (EROS) Data Center, Sioux Falls, South Dakota.<p/>Elevation data used in this image was acquired by the Shuttle Radar Topography Mission (SRTM) aboard the Space Shuttle Endeavour, launched on February 11,2000. SRTM used the same radar instrument that comprised the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70175071','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70175071"><span>Hydrothermal frictional strengths of rock and mineral samples relevant to the creeping section of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Moore, Diane E.; Lockner, David A.; Hickman, Stephen H.</p> <p>2016-01-01</p> <p>We compare frictional strengths in the temperature range 25–250 °C of fault gouge from SAFOD (CDZ and SDZ) with quartzofeldspathic wall rocks typical of the central creeping section of the San <span class="hlt">Andreas</span> Fault (Great Valley sequence and Franciscan Complex). The Great Valley and Franciscan samples have coefficients of friction, μ > 0.35 at all experimental conditions. Strength is unchanged between 25° and 150 °C, but μ increases at higher temperatures, exceeding 0.50 at 250 °C. Both samples are velocity strengthening at room temperature but show velocity-weakening behavior beginning at 150 °C and stick-slip motion at 250 °C. These rocks, therefore, have the potential for unstable seismic slip at depth. The CDZ gouge, with a high saponite content, is weak (μ = 0.09–0.17) and velocity strengthening in all experiments, and μ decreases at temperatures above 150 °C. Behavior of the SDZ is intermediate between the CDZ and wall rocks: μ < 0.2 and does not vary with temperature. Although saponite is probably not stable at depths greater than ∼3 km, substitution of the frictionally similar minerals talc and Mg-rich chlorite for saponite at higher temperatures could potentially extend the range of low strength and stable slip down to the base of the seismogenic zone.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70010424','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70010424"><span>Quaternary crustal deformation along a major branch of the San <span class="hlt">Andreas</span> fault in central California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Weber, G.E.; Lajoie, K.R.; Wehmiller, J.F.</p> <p>1979-01-01</p> <p>Deformed marine terraces and alluvial deposits record Quaternary crustal deformation along segments of a major, seismically active branch of the San <span class="hlt">Andreas</span> fault which extends 190 km SSE roughly parallel to the California coastline from Bolinas Lagoon to the Point Sur area. Most of this complex fault zone lies offshore (mapped by others using acoustical techniques), but a 4-km segment (Seal Cove fault) near Half Moon Bay and a 26-km segment (San Gregorio fault) between San Gregorio and Point Ano Nuevo lie onshore. At Half Moon Bay, right-lateral slip and N-S horizontal compression are expressed by a broad, synclinal warp in the first (lowest: 125 ka?) and second marine terraces on the NE side of the Seal Cove fault. This structure plunges to the west at an oblique angle into the fault plane. Linear, joint0controlled stream courses draining the coastal uplands are deflected toward the topographic depression along the synclinal axis where they emerge from the hills to cross the lowest terrace. Streams crossing the downwarped part of this terrace adjacent to Half Moon Bay are depositing alluvial fans, whereas streams crossing the uplifted southern limb of the syncline southwest of the bay are deeply incised. Minimum crustal shortening across this syncline parallel to the fault is 0.7% over the past 125 ka, based on deformation of the shoreline angle of the first terrace. Between San Gregorio and Point Ano Nuevo the entire fault zone is 2.5-3.0 km wide and has three primary traces or zones of faulting consisting of numerous en-echelon and anastomozing secondary fault traces. Lateral discontinuities and variable deformation of well-preserved marine terrace sequences help define major structural blocks and document differential motions in this area and south to Santa Cruz. Vertical displacement occurs on all of the fault traces, but is small compared to horizontal displacement. Some blocks within the fault zone are intensely faulted and steeply tilted. One major block 0</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0694.photos.060466p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0694.photos.060466p/"><span>4. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, Illinois Veterans Administration Medical Center. WEST ELEVATION, SOUTH ELEVATION, EAST ELEVATION, NORTH ELEVATION; DRAWING 2-3R, DATED SEPTEMBER 4, 1940. (8 x 10 negative) - Veterans Administration Medical Center, Building No. 2, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.epa.gov/renewable-fuel-standard-program/poet-marion-approval','PESTICIDES'); return false;" href="https://www.epa.gov/renewable-fuel-standard-program/poet-marion-approval"><span>Poet <span class="hlt">Marion</span> Approval</span></a></p> <p><a target="_blank" href="http://www.epa.gov/pesticides/search.htm">EPA Pesticide Factsheets</a></p> <p></p> <p></p> <p>This update August 9, 2016 letter from EPA approves, with modifications, the petition from Poet Biorefining-North Manchester, LLC, regarding non-grandfathered ethanol produced through a dry mill process, qualifying under the Clean Air Act for renewable</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25521005','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25521005"><span>Zoogeography of the San <span class="hlt">Andreas</span> Fault system: Great Pacific Fracture Zones correspond with spatially concordant phylogeographic boundaries in western North America.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gottscho, Andrew D</p> <p>2016-02-01</p> <p>The purpose of this article is to provide an ultimate tectonic explanation for several well-studied zoogeographic boundaries along the west coast of North America, specifically, along the boundary of the North American and Pacific plates (the San <span class="hlt">Andreas</span> Fault system). By reviewing 177 references from the plate tectonics and zoogeography literature, I demonstrate that four Great Pacific Fracture Zones (GPFZs) in the Pacific plate correspond with distributional limits and spatially concordant phylogeographic breaks for a wide variety of marine and terrestrial animals, including invertebrates, fish, amphibians, reptiles, birds, and mammals. These boundaries are: (1) Cape Mendocino and the North Coast Divide, (2) Point Conception and the Transverse Ranges, (3) Punta Eugenia and the Vizcaíno Desert, and (4) Cabo Corrientes and the Sierra Transvolcanica. However, discussion of the GPFZs is mostly absent from the zoogeography and phylogeography literature likely due to a disconnect between biologists and geologists. I argue that the four zoogeographic boundaries reviewed here ultimately originated via the same geological process (triple junction evolution). Finally, I suggest how a comparative phylogeographic approach can be used to test the hypothesis presented here. © 2014 Cambridge Philosophical Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70036286','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70036286"><span>Tremor reveals stress shadowing, deep postseismic creep, and depth-dependent slip recurrence on the lower-crustal San <span class="hlt">Andreas</span> fault near Parkfield</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Shelly, David R.; Johnson, Kaj M.</p> <p>2011-01-01</p> <p>The 2003 magnitude 6.5 San Simeon and the 2004 magnitude 6.0 Parkfield earthquakes induced small, but significant, static stress changes in the lower crust on the central San <span class="hlt">Andreas</span> fault, where recently detected tectonic tremor sources provide new constraints on deep fault creep processes. We find that these earthquakes affect tremor rates very differently, consistent with their differing transferred static shear stresses. The San Simeon event appears to have cast a "stress shadow" north of Parkfield, where tremor activity was stifled for 3-6 weeks. In contrast, the 2004 Parkfield earthquake dramatically increased tremor activity rates both north and south of Parkfield, allowing us to track deep postseismic slip. Following this event, rates initially increased by up to two orders of magnitude for the relatively shallow tremor sources closest to the rupture, with activity in some sources persisting above background rates for more than a year. We also observe strong depth dependence in tremor recurrence patterns, with shallower sources generally exhibiting larger, less-frequent bursts, possibly signaling a transition toward steady creep with increasing temperature and depth. Copyright 2011 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70193652','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70193652"><span>Delayed dynamic triggering of deep tremor along the Parkfield-Cholame section of the San <span class="hlt">Andreas</span> Fault following the 2014 M6.0 South Napa earthquake</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Peng, Zhigang; Shelly, David R.; Ellsworth, William L.</p> <p>2015-01-01</p> <p>Large, distant earthquakes are known to trigger deep tectonic tremor along the San <span class="hlt">Andreas</span> Fault and in subduction zones. However, there are relatively few observations of triggering from regional distance earthquakes. Here we show that a small tremor episode about 12–18 km NW of Parkfield was triggered during and immediately following the passage of surface waves from the 2014 Mw 6.0 South Napa main shock. More notably, a major tremor episode followed, beginning about 12 h later, and centered SE of Parkfield near Cholame. This major episode is one of the largest seen over the past several years, containing intense activity for ~3 days and taking more than 3 weeks to return to background levels. This episode showed systematic along-strike migration at ~5 km/d, suggesting that it was driven by a slow-slip event. Our results suggest that moderate-size earthquakes are capable of triggering major tremor and deep slow slip at regional distances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMNS31A0004A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMNS31A0004A"><span>Magnetic profiling of the San <span class="hlt">Andreas</span> Fault using a dual magnetometer UAV aerial survey system.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Abbate, J. A.; Angelopoulos, V.; Masongsong, E. V.; Yang, J.; Medina, H. R.; Moon, S.; Davis, P. M.</p> <p>2017-12-01</p> <p>Aeromagnetic survey methods using planes are more time-effective than hand-held methods, but can be far more expensive per unit area unless large areas are covered. The availability of low cost UAVs and low cost, lightweight fluxgate magnetometers (FGMs) allows, with proper offset determination and stray fields correction, for low-cost magnetic surveys. Towards that end, we have developed a custom multicopter UAV for magnetic mapping using a dual 3-axis fluxgate magnetometer system: the GEOphysical Drone Enhanced Survey Instrument (GEODESI). A high precision sensor measures the UAV's position and attitude (roll, pitch, and yaw) and is recorded using a custom Arduino data processing system. The two FGMs (in-board and out-board) are placed on two ends of a vertical 1m boom attached to the base of the UAV. The in-board FGM is most sensitive to stray fields from the UAV and its signal is used, after scaling, to clean the signal of the out-board FGM from the vehicle noise. The FGMs record three orthogonal components of the magnetic field in the UAV body coordinates which are then transformed into a north-east-down coordinate system using a rotation matrix determined from the roll-pitch-yaw attitude data. This ensures knowledge of the direction of all three field components enabling us to perform inverse modeling of magnetic anomalies with greater accuracy than total or vertical field measurements used in the past. Field tests were performed at Dragon's Back Pressure Ridge in the Carrizo Plain of California, where there is a known crossing of the San <span class="hlt">Andreas</span> Fault. Our data and models were compared to previously acquired LiDAR and hand-held magnetometer measurements. Further tests will be carried out to solidify our results and streamline our processing for educational use in the classroom and student field training.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70192794','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70192794"><span>A new perspective on the geometry of the San <span class="hlt">Andreas</span> Fault in southern California and its relationship to lithospheric structure</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fuis, Gary S.; Scheirer, Daniel S.; Langenheim, Victoria; Kohler, Monica D.</p> <p>2012-01-01</p> <p>The widely held perception that the San <span class="hlt">Andreas</span> fault (SAF) is vertical or steeply dipping in most places in southern California may not be correct. From studies of potential‐field data, active‐source imaging, and seismicity, the dip of the SAF is significantly nonvertical in many locations. The direction of dip appears to change in a systematic way through the Transverse Ranges: moderately southwest (55°–75°) in the western bend of the SAF in the Transverse Ranges (Big Bend); vertical to steep in the Mojave Desert; and moderately northeast (37°–65°) in a region extending from San Bernardino to the Salton Sea, spanning the eastern bend of the SAF in the Transverse Ranges. The shape of the modeled SAF is crudely that of a propeller. If confirmed by further studies, the geometry of the modeled SAF would have important implications for tectonics and strong ground motions from SAF earthquakes. The SAF can be traced or projected through the crust to the north side of a well documented high‐velocity body (HVB) in the upper mantle beneath the Transverse Ranges. The north side of this HVB may be an extension of the plate boundary into the mantle, and the HVB would appear to be part of the Pacific plate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T21B2806F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T21B2806F"><span>Structure of the San <span class="hlt">Andreas</span> Fault Zone in the Salton Trough Region of Southern California: A Comparison with San <span class="hlt">Andreas</span> Fault Structure in the Loma Prieta Area of Central California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fuis, G. S.; Catchings, R.; Scheirer, D. S.; Goldman, M.; Zhang, E.; Bauer, K.</p> <p>2016-12-01</p> <p>The San <span class="hlt">Andreas</span> fault (SAF) in the northern Salton Trough, or Coachella Valley, in southern California, appears non-vertical and non-planar. In cross section, it consists of a steeply dipping segment (75 deg dip NE) from the surface to 6- to 9-km depth, and a moderately dipping segment below 6- to 9-km depth (50-55 deg dip NE). It also appears to branch upward into a flower-like structure beginning below about 10-km depth. Images of the SAF zone in the Coachella Valley have been obtained from analysis of steep reflections, earthquakes, modeling of potential-field data, and P-wave tomography. Review of seismological and geodetic research on the 1989 M 6.9 Loma Prieta earthquake, in central California (e.g., U.S. Geological Survey Professional Paper 1550), shows several features of SAF zone structure similar to those seen in the northern Salton Trough. Aftershocks in the Loma Prieta epicentral area form two chief clusters, a tabular zone extending from 18- to 9-km depth and a complex cluster above 5-km depth. The deeper cluster has been interpreted to surround the chief rupture plane, which dips 65-70 deg SW. When double-difference earthquake locations are plotted, the shallower cluster contains tabular subclusters that appear to connect the main rupture with the surface traces of the Sargent and Berrocal faults. In addition, a diffuse cluster may surround a steep to vertical fault connecting the main rupture to the surface trace of the SAF. These interpreted fault connections from the main rupture to surface fault traces appear to define a flower-like structure, not unlike that seen above the moderately dipping segment of the SAF in the Coachella Valley. But importantly, the SAF, interpreted here to include the main rupture plane, appears segmented, as in the Coachella Valley, with a moderately dipping segment below 9-km depth and a steep to vertical segment above that depth. We hope to clarify fault-zone structure in the Loma Prieta area by reanalyzing active</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.T41A2095R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.T41A2095R"><span>Permeability of the San <span class="hlt">Andreas</span> Fault Zone at Depth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rathbun, A. P.; Song, I.; Saffer, D.</p> <p>2010-12-01</p> <p>Quantifying fault rock permeability is important toward understanding both the regional hydrologic behavior of fault zones, and poro-elastic processes that affect fault mechanics by mediating effective stress. These include long-term fault strength as well as dynamic processes that may occur during earthquake slip, including thermal pressurization and dilatancy hardening. Despite its importance, measurements of fault zone permeability for relevant natural materials are scarce, owing to the difficulty of coring through active fault zones seismogenic depths. Most existing measurements of fault zone permeability are from altered surface samples or from thinner, lower displacement faults than the SAF. Here, we report on permeability measurements conducted on gouge from the actively creeping Central Deformation Zone (CDZ) of the San <span class="hlt">Andreas</span> Fault, sampled in the SAFOD borehole at a depth of ~2.7 km (Hole G, Run 4, sections 4,5). The matrix of the gouge in this interval is predominantly composed of particles <10 µm, with ~5 vol% clasts of serpentinite, very fine-grained sandstone, and siltstone. The 2.6 m-thick CDZ represents the main fault trace and hosts ~90% of the active slip on the SAF at this location, as documented by repeated casing deformation surveys. We measured permeability in two different configurations: (1) in a uniaxial pressure cell, in which a sample is placed into a rigid steel ring which imposes a zero lateral strain condition and subjected to axial load, and (2) in a standard triaxial system under isostatic stress conditions. In the uniaxial configuration, we obtained permeabilities at axial effective stresses up to 90 MPa, and in the triaxial system up to 10 MPa. All experiments were conducted on cylindrical subsamples of the SAFOD core 25 mm in diameter, with lengths ranging from 18mm to 40mm, oriented for flow approximately perpendicular to the fault. In uniaxial tests, permeability is determined by running constant rate of strain (CRS) tests up</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70029192','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70029192"><span>Orientation of three-component geophones in the San <span class="hlt">Andreas</span> Fault observatory at depth Pilot Hole, Parkfield, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Oye, V.; Ellsworth, W.L.</p> <p>2005-01-01</p> <p>To identify and constrain the target zone for the planned SAFOD Main Hole through the San <span class="hlt">Andreas</span> Fault (SAF) near Parkfield, California, a 32-level three-component (3C) geophone string was installed in the Pilot Hole (PH) to monitor and improve the locations of nearby earthquakes. The orientation of the 3C geophones is essential for this purpose, because ray directions from sources may be determined directly from the 3D particle motion for both P and S waves. Due to the complex local velocity structure, rays traced from explosions and earthquakes to the PH show strong ray bending. Observed azimuths are obtained from P-wave polarization analysis, and ray tracing provides theoretical estimates of the incoming wave field. The differences between the theoretical and the observed angles define the calibration azimuths. To investigate the process of orientation with respect to the assumed velocity model, we compare calibration azimuths derived from both a homogeneous and 3D velocity model. Uncertainties in the relative orientation between the geophone levels were also estimated for a cluster of 36 earthquakes that was not used in the orientation process. The comparison between the homogeneous and the 3D velocity model shows that there are only minor changes in these relative orientations. In contrast, the absolute orientations, with respect to global North, were significantly improved by application of the 3D model. The average data residual decreased from 13?? to 7??, supporting the importance of an accurate velocity model. We explain the remaining residuals by methodological uncertainties and noise and with errors in the velocity model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70048665','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70048665"><span>Deep rock damage in the San <span class="hlt">Andreas</span> Fault revealed by P- and S-type fault-zone-guided waves</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ellsworth, William L.; Malin, Peter E.</p> <p>2011-01-01</p> <p>Damage to fault-zone rocks during fault slip results in the formation of a channel of low seismic-wave velocities. Within such channels guided seismic waves, denoted by Fg, can propagate. Here we show with core samples, well logs and Fg-waves that such a channel is crossed by the SAFOD (San <span class="hlt">Andreas</span> Fault Observatory at Depth) borehole at a depth of 2.7 km near Parkfield, California, USA. This laterally extensive channel extends downwards to at least half way through the seismogenic crust, more than about 7 km. The channel supports not only the previously recognized Love-type- (FL) and Rayleigh-type- (FR) guided waves, but also a new fault-guided wave, which we name FF. As recorded 2.7 km underground, FF is normally dispersed, ends in an Airy phase, and arrives between the P- and S-waves. Modelling shows that FF travels as a leaky mode within the core of the fault zone. Combined with the drill core samples, well logs and the two other types of guided waves, FF at SAFOD reveals a zone of profound, deep, rock damage. Originating from damage accumulated over the recent history of fault movement, we suggest it is maintained either by fracturing near the slip surface of earthquakes, such as the 1857 Fort Tejon M 7.9, or is an unexplained part of the fault-creep process known to be active at this site.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060459p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060459p/"><span>20. Photograph of a line drawing in the possession of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>20. Photograph of a line drawing in the possession of the Engineer's Office of the <span class="hlt">Marion</span>, IL Veterans Administration Medical Center. 3/4' SCALE DETAIL - MAIN ENT. UNIT (PLAN SECTIONS, HALF ELEVATION, SECTIONS THROUGH WINDOW & DOOR. HALF PLAN & HALF REFLECTED PLAN OF MARQUISE); DRAWING 1-13R, DATED SEPTEMBER 4, 1940. - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S53F..08T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S53F..08T"><span>Using low-frequency earthquake families on the San <span class="hlt">Andreas</span> fault as deep creepmeters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thomas, A.; Beeler, N. M.; Bletery, Q.; Burgmann, R.; Shelly, D. R.</p> <p>2017-12-01</p> <p>The San <span class="hlt">Andreas</span> fault hosts tectonic tremor and low-frequency earthquakes (LFEs) similar to those in subduction zone environments. These LFEs are grouped into families based on waveform similarity and locate between 16 and 29 km depth along a 150-km-long section of the fault centered on Parkfield, CA. ­Within individual LFE families event occurrence is not steady. In some families, bursts of a few events recur on timescales of days while in other families there are nearly quiescent periods that often last for months followed by episodes where hundreds of events occur over the course of a few days. These two different styles of LFE occurrence are called continuous and episodic respectively. LFEs are often assumed to reflect persistent regions that periodically fail during the aseismic shear of the surrounding fault allowing them to be used as creepmeters. We test this idea by formalizing the definition of a creepmeter (the LFE occurrence rate is proportional to the local fault slip rate), determining whether this definition is consistent with the observations, and over what timescale. We use the recurrence intervals of LFEs within individual families to create a catalog of LFE bursts. For the episodic families, we consider both longer duration (multiday) inferred creep episodes (dubbed long-timescale episodic) as well as the frequent short-term bursts of events that occur many times during inferred creep episodes (dubbed short-timescale episodic). We then use the recurrence intervals of LFE bursts to estimate the timing, duration, recurrence interval, slip, and slip rate associated with inferred slow slip events. We find that continuous families and the short-timescale episodic families appear to be inconsistent with our definition of a creepmeter (defined on the recurrence interval timescale) because their estimated durations are not physically meaningful. A straight-forward interpretation of the frequent short-term bursts of the continuous and short</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JSeis..20..291Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JSeis..20..291Z"><span>Frictional strength and heat flow of southern San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhu, P. P.</p> <p>2016-01-01</p> <p>Frictional strength and heat flow of faults are two related subjects in geophysics and seismology. To date, the investigation on regional frictional strength and heat flow still stays at the stage of qualitative estimation. This paper is concentrated on the regional frictional strength and heat flow of the southern San <span class="hlt">Andreas</span> Fault (SAF). Based on the in situ borehole measured stress data, using the method of 3D dynamic faulting analysis, we quantitatively determine the regional normal stress, shear stress, and friction coefficient at various seismogenic depths. These new data indicate that the southern SAF is a weak fault within the depth of 15 km. As depth increases, all the regional normal and shear stresses and friction coefficient increase. The former two increase faster than the latter. Regional shear stress increment per kilometer equals 5.75 ± 0.05 MPa/km for depth ≤15 km; regional normal stress increment per kilometer is equal to 25.3 ± 0.1 MPa/km for depth ≤15 km. As depth increases, regional friction coefficient increment per kilometer decreases rapidly from 0.08 to 0.01/km at depths less than ~3 km. As depth increases from ~3 to ~5 km, it is 0.01/km and then from ~5 to 15 km, and it is 0.002/km. Previously, frictional strength could be qualitatively determined by heat flow measurements. It is difficult to obtain the quantitative heat flow data for the SAF because the measured heat flow data exhibit large scatter. However, our quantitative results of frictional strength can be employed to investigate the heat flow in the southern SAF. We use a physical quantity P f to describe heat flow. It represents the dissipative friction heat power per unit area generated by the relative motion of two tectonic plates accommodated by off-fault deformation. P f is called "fault friction heat." On the basis of our determined frictional strength data, utilizing the method of 3D dynamic faulting analysis, we quantitatively determine the regional long-term fault</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JGRB..11612111S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JGRB..11612111S"><span>A reevaluation of the Pallett Creek earthquake chronology based on new AMS radiocarbon dates, San <span class="hlt">Andreas</span> fault, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Scharer, Katherine M.; Biasi, Glenn P.; Weldon, Ray J., II</p> <p>2011-12-01</p> <p>The Pallett Creek paleoseismic record occupies a keystone position in most attempts to develop rupture histories for the southern San <span class="hlt">Andreas</span> fault. Previous estimates of earthquake ages at Pallett Creek were determined by decay counting radiocarbon methods. That method requires large samples which can lead to unaccounted sources of uncertainty in radiocarbon ages because of the heterogeneous composition of organic layers. In contrast, accelerator mass spectrometry (AMS) radiocarbon dates may be obtained from small samples that have known carbon sources and also allow for a more complete sampling of the section. We present 65 new AMS radiocarbon dates that span nine ground-rupturing earthquakes at Pallett Creek. Overall, the AMS dates are similar to and reveal no dramatic bias in the conventional dates. For many layers, however, individual charcoal samples were younger than the conventional dates, leading to earthquake ages that are overall slightly younger than previously reported. New earthquake ages are determined by Bayesian refinement of the layer ages based on stratigraphic ordering and sedimentological constraints. The new chronology is more regular than previously published records in large part due to new samples constraining the age of event R. The closed interval from event C to 1857 has a mean recurrence of 135 years (σ = 83.2 years) and a quasiperiodic coefficient of variation (COV) of 0.61. We show that the new dates and resultant earthquake chronology have a stronger effect on COV than the specific membership of this long series and dating precision improvements from sedimentation rates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011HydJ...19..237A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011HydJ...19..237A"><span>Helium measurements of pore fluids obtained from the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD, USA) drill cores</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ali, S.; Stute, M.; Torgersen, T.; Winckler, G.; Kennedy, B. M.</p> <p>2011-02-01</p> <p>4He accumulated in fluids is a well established geochemical tracer used to study crustal fluid dynamics. Direct fluid samples are not always collectable; therefore, a method to extract rare gases from matrix fluids of whole rocks by diffusion has been adapted. Helium was measured on matrix fluids extracted from sandstones and mudstones recovered during the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) drilling in California, USA. Samples were typically collected as subcores or from drillcore fragments. Helium concentration and isotope ratios were measured 4-6 times on each sample, and indicate a bulk 4He diffusion coefficient of 3.5 ± 1.3 × 10-8 cm2 s-1 at 21°C, compared to previously published diffusion coefficients of 1.2 × 10-18 cm2 s-1 (21°C) to 3.0 × 10-15 cm2 s-1 (150°C) in the sands and clays. Correcting the diffusion coefficient of 4Hewater for matrix porosity (˜3%) and tortuosity (˜6-13) produces effective diffusion coefficients of 1 × 10-8 cm2 s-1 (21°C) and 1 × 10-7 (120°C), effectively isolating pore fluid 4He from the 4He contained in the rock matrix. Model calculations indicate that <6% of helium initially dissolved in pore fluids was lost during the sampling process. Complete and quantitative extraction of the pore fluids provide minimum in situ porosity values for sandstones 2.8 ± 0.4% (SD, n = 4) and mudstones 3.1 ± 0.8% (SD, n = 4).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.T41A2100H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.T41A2100H"><span>Implications of Microstructural Studies of the SAFOD Gouge for the Strength and Deformation Mechanisms in the Creeping Segment of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hadizadeh, J.; Gratier, J. L.; Mittempergher, S.; Renard, F.; Richard, J.; di Toro, G.; Babaie, H. A.</p> <p>2010-12-01</p> <p>The San <span class="hlt">Andreas</span> Fault zone (SAF) in the vicinity of the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD)in central California is characterized by an average 21 mm/year aseismic creep and strain release through repeating M<3 earthquakes. Seismic inversion studies indicate that the ruptures occur on clusters of stationary patches making up 1% or less of the total fault surface area. The existence of these so-called asperity patches, although not critical in determining the fault strength, suggests interaction of different deformation mechanisms. What are the deformation mechanisms, and how do the mechanisms couple and factor into the current strength models for the SAF? The SAFOD provides core samples and geophysical data including cores from two shear zones where the main borehole casing is deforming. The studies so far show a weak fault zone with about 200m of low-permeability damage zone without anomalous temperature or high fluid pressure (Zoback et al. EOS 2010). To answer the above questions, we studied core samples and thin sections ranging in measured depths (MD) from 3059m to 3991m including gouge from borehole casing deformation zones. The methods of study included high resolution scanning and transmission electron microscopy, cathodoluminescence imaging, X-ray fluorescence mapping, and energy dispersive X-ray spectroscopy. The microstructural and analytical data suggest that deformation is by a coupling of cataclastic flow and pressure solution accompanied by widespread alteration of feldspar to clay minerals and other neomineralizations. The clay contents of the gouge and streaks of serpentinite are not uniformly distributed, but weakness of the creeping segment is likely to be due to intrinsically low frictional strength of the fault material. This conclusion, which is based on the overall ratio of clay/non-clay constituents and the presence of talc in the actively deforming zones, is consistent with the 0.3-0.45 coefficient of friction for the drill</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T41B2907Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T41B2907Y"><span>New Holocene Slip-rate Sites Along the Mojave San <span class="hlt">Andreas</span> Fault Near Palmdale, CA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Young, E. K.; Cowgill, E.; Scharer, K. M.</p> <p>2016-12-01</p> <p>The slip rate for the Mojave San <span class="hlt">Andreas</span> fault (MSAF) is poorly known: the long-term ( 413ky) geologic rate is as high as 37mm/yr [1] while the geodetic slip rate is as low as 15mm/yr [2]. To determine the Holocene-average rate for the MSAF, we investigated several offset landforms near Palmdale, CA using detailed surficial mapping and 14C analyses of charcoal collected from hand-dug excavations. Site X-12 preserves two offset markers sourced from a north-flowing catchment south of the fault. One is a terrace riser cut into older alluvium that is abutted by the eastern margin of a Qf3 fan and is displaced by 75 m. Offset of the riser and the alluvium generated a north-facing fault-scarp that was eroded to form a small fan on top of the eastern margin of the Qf3 fan. The second offset is a beheaded channel that is incised into the Qf3 fan north of the fault and is displaced 50m, and thus formed after the riser was offset by 25m. To date these landforms we opened 5 excavations at the site, with emphasis on the beheaded channel. Dates from within the Qf3 fan south/upstream of the fault and abandoned bedload in the beheaded channel give maximum and minimum ages for the channel incision of 1500calBP and 600calBP, respectively, implying a rate ≥33mm/yr. Future work seeks to date post-abandonment channel-fill deposits and the upper and lower bounds on the terrace riser. At a second site, Ranch Center, a north-flowing stream cut across a shutter ridge north of the fault and was then offset and deflected 80m before cutting a new channel across the fault. Future work here seeks to date the bases of the alluvial fans deposited by the abandoned and active channels to obtain maximum and minimum ages for the 80 m offset, respectively. Rates from these sites should help to better define the Holocene slip rate on the MSAF. 1) Matmon et al., 2005, GSAB. v. 117 p. 795 2) Becker et al., 2005, Geoph.. J. Int., v. 160 p. 634</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AGUFM.T24B..06C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AGUFM.T24B..06C"><span>Characterization of the San <span class="hlt">Andreas</span> Fault at Parkfield Using a Massive 3D VSP</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chavarria, J.; Goertz, A.; Karrenbach, M.; Milligan, P.; Paulsson, B.</p> <p>2005-12-01</p> <p>In preparation for the drilling of SAFOD's Phase II we installed an 80 level array of 3C seismometers inside the well. The goal of the array was to refine the existing velocity model to better locate the target events, and to monitor the local seismicity. The array, with sensors laying mostly within the deviated portion of the well, spans depths ranging from 2.7 to 1.5 km with levels every 15 m. It is this dense spacing what makes 3D VSP capable of bridging the gap between drill-hole observations and observations from the surface like 2D seismics. During April and May 2005 we recorded thirteen far offset shots surrounding the SAFOD site and target event area. Data from these shots was simultaneously recorded by the surface networks and used for better location of the target events. In addition to these, a zero offset shot at SAFOD was generated to refine the structure surrounding the well. The 1D velocity model inverted from the zero offset is representative of the current geologic model at SAFOD. The complexity of the velocity model for this segment of the fault can be inferred from deviations between the zero offset model and the shorter wavelength model derived from well logs. In addition to strong changes in velocity, both zero offset and far offset shots show the presence of strong scattered phases associated to the complex geologic structure of the San <span class="hlt">Andreas</span> Fault Zone. In addition to the active portion of the experiment we monitored the local seismicity (i.e. aftershocks from the Parkfield 2004 event) over a period of 13 days. During this period of time we recorded continuously at high sampling rates (4kHz) a large number of events, some of which were located by the surface networks and felt onsite. The quiet environment in the borehole enabled us to record microearthquakes that were not present in the NCEDC catalog. In some cases these small events were not even recorded along the entire array. Besides its high level of event detection, the high vector</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70192091','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70192091"><span>The Evergreen basin and the role of the Silver Creek fault in the San <span class="hlt">Andreas</span> fault system, San Francisco Bay region, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Jachens, Robert C.; Wentworth, Carl M.; Graymer, Russell W.; Williams, Robert; Ponce, David A.; Mankinen, Edward A.; Stephenson, William J.; Langenheim, Victoria</p> <p>2017-01-01</p> <p>The Evergreen basin is a 40-km-long, 8-km-wide Cenozoic sedimentary basin that lies mostly concealed beneath the northeastern margin of the Santa Clara Valley near the south end of San Francisco Bay (California, USA). The basin is bounded on the northeast by the strike-slip Hayward fault and an approximately parallel subsurface fault that is structurally overlain by a set of west-verging reverse-oblique faults which form the present-day southeastward extension of the Hayward fault. It is bounded on the southwest by the Silver Creek fault, a largely dormant or abandoned fault that splays from the active southern Calaveras fault. We propose that the Evergreen basin formed as a strike-slip pull-apart basin in the right step from the Silver Creek fault to the Hayward fault during a time when the Silver Creek fault served as a segment of the main route by which slip was transferred from the central California San <span class="hlt">Andreas</span> fault to the Hayward and other East Bay faults. The dimensions and shape of the Evergreen basin, together with palinspastic reconstructions of geologic and geophysical features surrounding it, suggest that during its lifetime, the Silver Creek fault transferred a significant portion of the ∼100 km of total offset accommodated by the Hayward fault, and of the 175 km of total San <span class="hlt">Andreas</span> system offset thought to have been accommodated by the entire East Bay fault system. As shown previously, at ca. 1.5–2.5 Ma the Hayward-Calaveras connection changed from a right-step, releasing regime to a left-step, restraining regime, with the consequent effective abandonment of the Silver Creek fault. This reorganization was, perhaps, preceded by development of the previously proposed basin-bisecting Mount Misery fault, a fault that directly linked the southern end of the Hayward fault with the southern Calaveras fault during extinction of pull-apart activity. Historic seismicity indicates that slip below a depth of 5 km is mostly transferred from the Calaveras</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.T23C2969K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.T23C2969K"><span>Geomorphological expression of a complex structural region: San <span class="hlt">Andreas</span> Fault through the San Gorgonio Pass, southern California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kendrick, K. J.; Matti, J. C.</p> <p>2015-12-01</p> <p>The San Gorgonio Pass (SGP) region of southern California is a locus of extensive Quaternary deformation surrounding a complex section of the San <span class="hlt">Andreas</span> Fault (SAF) zone. The geomorphology of the SGP region reflects the complicated history of geologic events in the formation of this structural 'knot'. Critical questions remain in assessing earthquake hazard for this region: What is the likelihood that rupture will propagate through the SGP? If rupture is able to propagate, what pathway will connect the various fault strands? To address these questions, we focus on the geology and geomorphology of the SGP region. We have identified fault-bounded blocks, and focus on three that are developed within crystalline bedrock: the Yucaipa Ridge block (YRB) block, the Kitching Peak block (KPB), and the Pisgah Peak block (PPB). The latter two blocks are positioned south of the YRB, and partially separated from each other by the San Bernardino strand; this strand cannot be mapped at the surface as an active connection between fault strands. Both KPB and PPB are bounded to the south by the San Gorgonio Pass Fault Zone. Morphometric analyses consistently demonstrate distinctions between KPB and PPB, though the bedrock lithologies are the same. Geologic mapping of the region highlights the differences in Quaternary units within the blocks. These geomorphic and geologic distinctions lead to our interpretation that KPB and PPB have experienced markedly different uplift histories that constrain the history of dextral slip on the SAF through SGP. Specifically, although the latest Quaternary geologic setting of SGP raises questions about modern slip transfer through the Pass, the contrasting uplift histories of KPB and PPB strongly suggest that earlier in Quaternary time SGP was not a barrier to slip transfer between the Coachella Valley to the SE and the San Bernardino Basin to the NW.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRB..123..583B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRB..123..583B"><span>Constraints on Friction, Dilatancy, Diffusivity, and Effective Stress From Low-Frequency Earthquake Rates on the Deep San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Beeler, N. M.; Thomas, Amanda; Bürgmann, Roland; Shelly, David</p> <p>2018-01-01</p> <p>Families of recurring low-frequency earthquakes (LFEs) within nonvolcanic tremor on the San <span class="hlt">Andreas</span> Fault in central California are sensitive to tidal stresses. LFEs occur at all levels of the tides, are strongly correlated and in phase with the 200 Pa shear stresses, and weakly and not systematically correlated with the 2 kPa tidal normal stresses. We assume that LFEs are small sources that repeatedly fail during shear within a much larger scale, aseismically slipping fault zone and consider two different models of the fault slip: (1) modulation of the fault slip rate by the tidal stresses or (2) episodic slip, triggered by the tides. LFEs are strongly clustered with duration much shorter than the semidiurnal tide; they cannot be significantly modulated on that time scale. The recurrence times of clusters, however, are many times longer than the semidiurnal, leading to an appearance of tidal triggering. In this context we examine the predictions of laboratory-observed triggered frictional (dilatant) fault slip. The undrained end-member model produces no sensitivity to the tidal normal stress, and slip onsets are in phase with the tidal shear stress. The tidal correlation constrains the diffusivity to be less than 1 × 10-6/s and the product of the friction and dilatancy coefficients to be at most 5 × 10-7, orders of magnitude smaller than observed at room temperature. In the absence of dilatancy the effective normal stress at failure would be about 55 kPa. For this model the observations require intrinsic weakness, low dilatancy, and lithostatic pore fluid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70028656','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70028656"><span>Slip on the San <span class="hlt">Andreas</span> fault at Parkfield, California, over two earthquake cycles, and the implications for seismic hazard</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Murray, J.; Langbein, J.</p> <p>2006-01-01</p> <p>Parkfield, California, which experienced M 6.0 earthquakes in 1934, 1966, and 2004, is one of the few locales for which geodetic observations span multiple earthquake cycles. We undertake a comprehensive study of deformation over the most recent earthquake cycle and explore the results in the context of geodetic data collected prior to the 1966 event. Through joint inversion of the variety of Parkfield geodetic measurements (trilateration, two-color laser, and Global Positioning System), including previously unpublished two-color data, we estimate the spatial distribution of slip and slip rate along the San <span class="hlt">Andreas</span> using a fault geometry based on precisely relocated seismicity. Although the three most recent Parkfield earthquakes appear complementary in their along-strike distributions of slip, they do not produce uniform strain release along strike over multiple seismic cycles. Since the 1934 earthquake, more than 1 m of slip deficit has accumulated on portions of the fault that slipped in the 1966 and 2004 earthquakes, and an average of 2 m of slip deficit exists on the 33 km of the fault southeast of Gold Hill to be released in a future, perhaps larger, earthquake. It appears that the fault is capable of partially releasing stored strain in moderate earthquakes, maintaining a disequilibrium through multiple earthquake cycles. This complicates the application of simple earthquake recurrence models that assume only the strain accumulated since the most recent event is relevant to the size or timing of an upcoming earthquake. Our findings further emphasize that accumulated slip deficit is not sufficient for earthquake nucleation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70019264','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019264"><span>Two-dimensional seismic image of the San <span class="hlt">Andreas</span> Fault in the Northern Gabilan Range, central California: Evidence for fluids in the fault zone</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Thurber, C.; Roecker, S.; Ellsworth, W.; Chen, Y.; Lutter, W.; Sessions, R.</p> <p>1997-01-01</p> <p>A joint inversion for two-dimensional P-wave velocity (Vp), P-to-S velocity ratio (Vp/Vs), and earthquake locations along the San <span class="hlt">Andreas</span> fault (SAF) in central California reveals a complex relationship among seismicity, fault zone structure, and the surface fault trace. A zone of low Vp and high Vp/Vs lies beneath the SAF surface trace (SAFST), extending to a depth of about 6 km. Most of the seismic activity along the SAF occurs at depths of 3 to 7 km in a southwest-dipping zone that roughly intersects the SAFST, and lies near the southwest edge of the low Vp and high Vp/Vs zones. Tests indicate that models in which this seismic zone is significantly closer to vertical can be confidently rejected. A second high Vp/Vs zone extends to the northeast, apparently dipping beneath the Diablo Range. Another zone of seismicity underlies the northeast portion of this Vp/Vs high. The high Vp/Vs zones cut across areas of very different Vp values, indicating that the high Vp/Vs values are due to the presence of fluids, not just lithology. The close association between the zones of high Vp/Vs and seismicity suggests a direct involvement of fluids in the faulting process. Copyright 1997 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.earthquakegeology.com/materials/proceedings/2016_Crestone.pdf','USGSPUBS'); return false;" href="http://www.earthquakegeology.com/materials/proceedings/2016_Crestone.pdf"><span>The Elizabeth Lake paleoseismic site: Rupture pattern constraints for the past ~800 years for the Mojave section of the south-central San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Bemis, Sean; Scharer, Katherine M.; Dolan, James F.; Rhodes, Ed</p> <p>2016-01-01</p> <p>The southern San <span class="hlt">Andreas</span> Fault in California has hosted two historic surface-rupturing earthquakes, the ~M7 1812 Wrightwood earthquake and the ~M7.9 1857 Fort Tejon earthquake (e.g., Sieh, 1978; Jacoby et al., 1988). Numerous paleoseismic studies have established chronologies of historic and prehistoric earthquakes at sites along the full length of the 1857 rupture (e.g., Sieh, 1978; Scharer et al., 2014). These studies provide an unparalleled opportunity to examine patterns of recent ruptures; however, at least two significant spatial gaps in high-quality paleoseismic sites remain. At ~100 km long each, these gaps contribute up to 100 km of uncertainty to paleo-rupture lengths and could also permit a surface rupture from an earthquake up to ~M7.2 to go undetected [using scaling relationships of Wells and Coppersmith (1994)]. Given the known occurrence of an ~M7 earthquake on this portion of the SAF (1812), it is critical to fill these gaps in order to better constrain paleo-rupture lengths and to increase the probability of capturing the full spatial record of surface rupturing earthquakes.   In this study, we target a new site within the 100 km long stretch of the San <span class="hlt">Andreas</span> Fault between the Frazier Mountain and Pallett Creek paleoseismic sites (Figure 1), near Elizabeth Lake, California. Prior excavations at the site during 1998-1999 encountered promising stratigraphy but these studies were hindered by shallow groundwater throughout the site. We began our current phase of investigations in 2012, targeting the northwestern end of a 40 x 350 m fault-parallel depression that defines the site (Figure 2). Subsequent investigations in 2013 and 2014 focused on the southeastern end of the depression where the fault trace is constrained between topographic highs and is proximal to an active drainage. In total, our paleoseismic investigations consist of 10 fault-perpendicular trenches that cross the depression (Figure 2) and expose a >2000 year depositional record</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.T21B0406K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.T21B0406K"><span>Fragmented Landscapes in the San Gorgonio Pass Region: Insights into Quaternary Strain History of the Southern San <span class="hlt">Andreas</span> Fault System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kendrick, K. J.; Matti, J. C.; Landis, G. P.; Alvarez, R. M.</p> <p>2006-12-01</p> <p>The San Gorgonio Pass (SGP) region is a zone of structural complexity within the southern San <span class="hlt">Andreas</span> Fault system that is characterized by (1) multiple strands of the San <span class="hlt">Andreas</span> Fault (SAF), (2) intense and diverse microseismicity, (3) contraction within the SGP fault zone (SGPfz), and (4) complex and diverse landforms - all a consequence of structural complications in the vicinity of the southeastern San Bernardino Mountains (SBM). Multiple strands of the SAF zone in the SGP region partition the landscape into discrete geomorphic/geologic domains, including: San Gorgonio Mountain (SGM), Yucaipa Ridge (YR), Kitching Peak (KP), Pisgah Peak (PP), and Coachella Valley (CV) domains. The morphology of each domain reflects the tectonic history unique to that region. Development of the SGP knot in the Mission Creek strand of the SAF (SAFmi) led to westward deflection of the SAFmi, juxtaposition of the KP, PP, and SGM domains, initiation of uplift of YR domain along thrust faults in headwaters of San Gorgonio River, and development of the San Jacinto Fault. Slip on the SAF diminished as a result, thereby allowing integrated drainage systems to develop in the greater SGP region. San Gorgonio River, Whitewater River, and Mission Creek are discrete drainages that transport sediment across the SGM, YR, PP, KP, and CV domains into alluvial systems peripheral to the SGP region. There, depositional units (San Timoteo Formation, upper member, deformed gravels of Whitewater River) all contain clasts of SBM-type and San Gabriel Mountain-type basement, thus constraining slip on the SAF in the SGP region. Middle and late Pleistocene slip on the Mill Creek strand of the SAF (SAFm) in the SGP region has attempted to bypass the SGP knot, and has disrupted landscapes established during SAFmi quiescence. Restoration of right-slip on the SAFm is key to deciphering landscape history. Matti and others (1985, 1992) proposed that a bi-lobed alluvial deposit in the Raywood Flats area has been</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AGUFM.S12F..10F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AGUFM.S12F..10F"><span>First Results from a Forward, 3-Dimensional Regional Model of a Transpressional San <span class="hlt">Andreas</span> Fault System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fitzenz, D. D.; Miller, S. A.</p> <p>2001-12-01</p> <p>We present preliminary results from a 3-dimensional fault interaction model, with the fault system specified by the geometry and tectonics of the San <span class="hlt">Andreas</span> Fault (SAF) system. We use the forward model for earthquake generation on interacting faults of Fitzenz and Miller [2001] that incorporates the analytical solutions of Okada [85,92], GPS-constrained tectonic loading, creep compaction and frictional dilatancy [Sleep and Blanpied, 1994, Sleep, 1995], and undrained poro-elasticity. The model fault system is centered at the Big Bend, and includes three large strike-slip faults (each discretized into multiple subfaults); 1) a 300km, right-lateral segment of the SAF to the North, 2) a 200km-long left-lateral segment of the Garlock fault to the East, and 3) a 100km-long right-lateral segment of the SAF to the South. In the initial configuration, three shallow-dipping faults are also included that correspond to the thrust belt sub-parallel to the SAF. Tectonic loading is decomposed into basal shear drag parallel to the plate boundary with a 35mm yr-1 plate velocity, and East-West compression approximated by a vertical dislocation surface applied at the far-field boundary resulting in fault-normal compression rates in the model space about 4mm yr-1. Our aim is to study the long-term seismicity characteristics, tectonic evolution, and fault interaction of this system. We find that overpressured faults through creep compaction are a necessary consequence of the tectonic loading, specifically where high normal stress acts on long straight fault segments. The optimal orientation of thrust faults is a function of the strike-slip behavior, and therefore results in a complex stress state in the elastic body. This stress state is then used to generate new fault surfaces, and preliminary results of dynamically generated faults will also be presented. Our long-term aim is to target measurable properties in or around fault zones, (e.g. pore pressures, hydrofractures, seismicity</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T41B2923M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T41B2923M"><span>Holocene geologic slip rate for the Mission Creek strand of the Southern San <span class="hlt">Andreas</span> Fault, northern Coachella Valley, CA.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Munoz, J. J.; Behr, W. M.; Sharp, W. D.; Fryer, R.; Gold, P. O.</p> <p>2016-12-01</p> <p>Slip on the southern San <span class="hlt">Andreas</span> fault in the northwestern Coachella Valley in Southern California is partitioned between three strands, the Mission Creek, Garnet Hill, and Banning strands. In the vicinity of the Indio Hills, the NW striking Mission Creek strand extends from the Indio Hills into the San Bernardino Mountains, whereas the Banning and Garnet Hill strands strike WNW and transfer slip into the San Gorgonio Pass region. Together, these three faults accommodate 20 mm/yr of right-lateral motion. Determining which strand accommodates the majority of fault slip and how slip rates on these strands have varied during the Quaternary is critical to seismic hazard assessment for the southern California region. Here we present a new Holocene geologic slip rate from an alluvial fan offset along the Mission Creek strand at the Three Palms site in the Indio Hills. Field mapping and remote sensing using the B4 LiDAR data indicates that the Three Palms fan is offset 57 +/- 3 meters. U-series dating on pedogenic carbonate rinds collected at 25-100 cm depth within the fan deposit constrain the minimum depositional age to 3.49 +/- 0.92 ka, yielding a maximum slip rate of 16 +6.1/-3.8 mm/yr. This Holocene maximum slip rate overlaps within errors with a previously published late Pleistocene slip rate of 12-22 mm/yr measured at Biskra Palms, a few kilometers to the south. Cosmogenic 10Be surface exposure samples were also collected from the fan surface to bracket the maximum depositional age. These samples have been processed and are currently awaiting AMS measurement.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002AGUFM.S21A0965M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002AGUFM.S21A0965M"><span>Examining Structural Controls on Earthquake Rupture Dynamics Along the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McGuire, J. J.; Ben-Zion, Y.</p> <p>2002-12-01</p> <p>Recent numerical simulations of dynamic rupture [Andrews and Ben-Zion, 1997; Harris and Day, 1997] have confirmed earlier analytical results [Weertman, 1980; Adams, 1995] that a contrast in elastic properties between the two sides of a fault will generate an interaction between the normal stress and fault slip that is not present in a homogeneous medium. It has been shown that for a range of frictional parameters and initial conditions, this interaction produces a statistical preference for unilateral rupture propagation in the direction of slip of the more compliant medium [Ben-Zion and Andrews, 1998; Cochard and Rice, 2000; Ben-Zion and Huang 2002]. Thus, the directivity of earthquake ruptures on large faults with well-developed material interfaces may be controlled by material contrasts of the rocks within and across the fault zone. One of the largest known velocity contrasts across a major crustal fault occurs along the Bear Valley section of the San <span class="hlt">Andreas</span> where high velocity materials on the SW side (P-velocity >5 km/s) are juxtaposed with low-velocity material on the NE side (P-velocity <4 km/s) down to a depth of about 4 km with a less dramatic contrast continuing to about 8 km [Thurber et al., 1997]. This boundary is strong enough to generate significant head-waves refracted along it that are recorded as the first arrivals at stations close to the fault on the NE side [McNally and McEvilly, 1977]. Rubin and Gillard [2000] and Rubin [2002] relocated the events in this region using NCSN waveform data and found that more than twice as many immediate aftershocks to small earthquakes occurred to the NW of the mainshock as to the SE, which they interpreted as being consistent with a preferred rupture direction to the SE. Their interpretation that aftershocks to microearthquakes occur preferentially in the direction opposite of rupture propagation has not been directly tested and is inconsistent with observations from moderate [Fletcher and Spudich, 1998] and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70041916','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70041916"><span>Fault zone structure from topography: signatures of en echelon fault slip at Mustang Ridge on the San <span class="hlt">Andreas</span> Fault, Monterey County, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>DeLong, Stephen B.; Hilley, George E.; Rymer, Michael J.; Prentice, Carol</p> <p>2010-01-01</p> <p>We used high-resolution topography to quantify the spatial distribution of scarps, linear valleys, topographic sinks, and oversteepened stream channels formed along an extensional step over on the San <span class="hlt">Andreas</span> Fault (SAF) at Mustang Ridge, California. This location provides detail of both creeping fault landform development and complex fault zone kinematics. Here, the SAF creeps 10–14 mm/yr slower than at locations ∼20 km along the fault in either direction. This spatial change in creep rate is coincident with a series of en echelon oblique-normal faults that strike obliquely to the SAF and may accommodate the missing deformation. This study presents a suite of analyses that are helpful for proper mapping of faults in locations where high-resolution topographic data are available. Furthermore, our analyses indicate that two large subsidiary faults near the center of the step over zone appear to carry significant distributed deformation based on their large apparent vertical offsets, the presence of associated sag ponds and fluvial knickpoints, and the observation that they are rotating a segment of the main SAF. Several subsidiary faults in the southeastern portion of Mustang Ridge are likely less active; they have few associated sag ponds and have older scarp morphologic ages and subdued channel knickpoints. Several faults in the northwestern part of Mustang Ridge, though relatively small, are likely also actively accommodating active fault slip based on their young morphologic ages and the presence of associated sag ponds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013GeoJI.194.1295P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013GeoJI.194.1295P"><span>Kinematics of rotating panels of E-W faults in the San <span class="hlt">Andreas</span> system: what can we tell from geodesy?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Platt, J. P.; Becker, T. W.</p> <p>2013-09-01</p> <p>Sets of E- to NE-trending sinistral and/or reverse faults occur within the San <span class="hlt">Andreas</span> system, and are associated with palaeomagnetic evidence for clockwise vertical-axis rotations. These structures cut across the trend of active dextral faults, posing questions as to how displacement is transferred across them. Geodetic data show that they lie within an overall dextral shear field, but the data are commonly interpreted to indicate little or no slip, nor any significant rate of rotation. We model these structures as rotating by bookshelf slip in a dextral shear field, and show that a combination of sinistral slip and rotation can produce the observed velocity field. This allows prediction of rates of slip, rotation, fault-parallel extension and fault-normal shortening within the panel. We use this method to calculate the kinematics of the central segment of the Garlock Fault, which cuts across the eastern California shear zone at a high angle. We obtain a sinistral slip rate of 6.1 ± 1.1 mm yr-1, comparable to geological evidence, but higher than most previous geodetic estimates, and a rotation rate of 4.0 ± 0.7° Myr-1 clockwise. The western Transverse Ranges transect a similar shear zone in coastal and offshore California, but at an angle of only 40°. As a result, the faults, which were sinistral when they were at a higher angle to the shear zone, have been reactivated in a dextral sense at a low rate, and the rate of rotation of the panel has decreased from its long-term rate of ˜5° to 1.6° ± 0.2° Myr-1 clockwise. These results help to resolve some of the apparent discrepancies between geological and geodetic slip-rate estimates, and provide an enhanced understanding of the mechanics of intracontinental transform systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.T51J..01C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.T51J..01C"><span>Velocity Gradient Across the San <span class="hlt">Andreas</span> Fault and Changes in Slip Behavior as Outlined by Full non Linear Tomography</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chiarabba, C.; Giacomuzzi, G.; Piana Agostinetti, N.</p> <p>2017-12-01</p> <p>The San <span class="hlt">Andreas</span> Fault (SAF) near Parkfield is the best known fault section which exhibit a clear transition in slip behavior from stable to unstable. Intensive monitoring and decades of studies permit to identify details of these processes with a good definition of fault structure and subsurface models. Tomographic models computed so far revealed the existence of large velocity contrasts, yielding physical insight on fault rheology. In this study, we applied a recently developed full non-linear tomography method to compute Vp and Vs models which focus on the section of the fault that exhibit fault slip transition. The new tomographic code allows not to impose a vertical seismic discontinuity at the fault position, as routinely done in linearized codes. Any lateral velocity contrast found is directly dictated by the data themselves and not imposed by subjective choices. The use of the same dataset of previous tomographic studies allows a proper comparison of results. We use a total of 861 earthquakes, 72 blasts and 82 shots and the overall arrival time dataset consists of 43948 P- and 29158 S-wave arrival times, accurately selected to take care of seismic anisotropy. Computed Vp and Vp/Vs models, which by-pass the main problems related to linarized LET algorithms, excellently match independent available constraints and show crustal heterogeneities with a high resolution. The high resolution obtained in the fault surroundings permits to infer lateral changes of Vp and Vp/Vs across the fault (velocity gradient). We observe that stable and unstable sliding sections of the SAF have different velocity gradients, small and negligible in the stable slip segment, but larger than 15 % in the unstable slip segment. Our results suggest that Vp and Vp/Vs gradients across the fault control fault rheology and the attitude of fault slip behavior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-05-05/pdf/2011-10925.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-05-05/pdf/2011-10925.pdf"><span>76 FR 25760 - Waiver of Aeronautical Land-Use Assurance <span class="hlt">Marion</span> Municipal Airport; <span class="hlt">Marion</span>, IN</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-05-05</p> <p>... sought to be purchased by the State of Indiana consists of approximately .683 acres. The FAA issued a... corner of said section, said southeast corner being designated as point ``80'' on said plat; thence South... said south line; thence North 0 degrees 26 minutes 29 seconds East 317.51 feet to point ``901...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.T11A2280G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.T11A2280G"><span>A kinematic model of patchy slip at depth explains observed tremor waveforms on the San <span class="hlt">Andreas</span> fault near Parkfield, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gottschaemmer, E.; Harrington, R. M.; Cochran, E. S.; Bohlen, T.</p> <p>2011-12-01</p> <p>Recent observations of both triggered and ambient tremor suggest that tremor results from simple shear-failure. Tremor episodes on the San <span class="hlt">Andreas</span> fault near Parkfield are thought to be comprised of clusters of individual events with frequencies between 2-8 Hz. Such low frequency earthquakes (LFEs) occur at depths where the frictional properties of the fault surface are primarily slip-strengthening with imbedded patches of slip weakening material that slip seismically when the surrounding fault creeps in a slow-slip event. Here we show new tremor waveforms from a temporary deployment of 13 broadband seismometers spaced at a maximum on the order of 30 km near Cholame, California are consistent with a series of small seismically slipping patches surrounded by an aseismic region along a fault surface. We model individual seismic events kinematically as small shear failures (M ~ 1) at depths exceeding 15 km. We use stress drop values of 1 MPa, based on a slip to fault area ratio. We simulate tremor recorded at the surface by our temporary array centered near Cholame, for frequencies up to 8 Hz using a staggered-grid finite-difference scheme to solve the elastic equations of motion, and the 3D velocity and density model from Thurber et al. (2006). Our simulations indicate that multiple seismically slipping patches in an aseismic region successfully recreate tremor characteristics observed in multiple studies, including individual tremor bursts, individual events, and episodic behavior. The kinematic model presented here will help to constrain the distribution and amplitude of the seismically slipping patches at depth, which will then be used in a dynamic model with variable frictional properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1980Tectp..67..221L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1980Tectp..67..221L"><span>Petrogenesis of cataclastic rocks within the San <span class="hlt">Andreas</span> fault zone of Southern California U.S.A.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lawford Anderson, J.; Osborne, Robert H.; Palmer, Donald F.</p> <p>1980-08-01</p> <p>This paper petrologically characterizes cataclastic rocks derived from four sites within the San <span class="hlt">Andreas</span> fault zone of southern California. In this area, the fault traverses an extensive plutonic and metamorphic terrane and the principal cataclastic rock formed at these upper crustal levels is unindurated gouge derived from a range of crystalline rocks including diorite, tonalite, granite, aplite, and pegmatite. The mineralogical nature of this gouge is decidedly different from the "clay gouge" reported by Wu (1975) for central California and is essentially a rock flour with a quartz, feldspar, biotite, chlorite, amphibole, epidote and oxide mineralogy representing the milled-down equivalent of the original rock. Clay development is minor (less than 4 wt. %) to nonexistent and is exclusively kaolinite. Alterations involve hematitic oxidation, chlorite alteration on biotite and amphibole, and local introduction of calcite. Electron microprobe analysis showed that in general the major minerals were not reequilibrated with the pressure—temperature regime imposed during cataclasis. Petrochemically, the form of cataclasis that we have investigated is largely an isochemical process. Some hydration occurs but the maximum amount is less than 2.2% added H 2O. Study of a 375 m deep core from a tonalite pluton adjacent to the fault showed that for Si, Al, Ti, Fe, Mg, Mn, K, Na, Li, Rb, and Ba, no leaching and/or enrichment occurred. Several samples experienced a depletion in Sr during cataclasis while lesser number had an enrichment of Ca (result of calcite veining). Texturally, the fault gouge is not dominated by clay-size material but consists largely of silt and fine sand-sized particles. An intriguing aspect of our work on the drill core is a general decrease in particulate size with depth (and confining pressure) with the predominate shifting sequentially from fine sand to silt-size material. The original fabric of these rocks is commonly not disrupted during the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.G41C..04L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.G41C..04L"><span>Groundwater withdrawal in the Central Valley, California: implications for San <span class="hlt">Andreas</span> Fault stressing and lithosphere rheology</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lundgren, P.; Liu, Z.; Ali, S. T.; Farr, T.; Faunt, C. C.</p> <p>2016-12-01</p> <p>Anthropogenic perturbations to crustal loading due to groundwater pumping are increasingly recognized as causing changes in nearby fault stresses. We present preliminary analysis of crustal unloading in the Central Valley (CV), California, for the period 2006-2010 to infer Coulomb stress changes on the central San <span class="hlt">Andreas</span> Fault (CSAF), lithospheric rheology, and system memory due to more than a century of groundwater withdrawal in the southern CV. We use data-driven unloading estimates to drive three-dimensional (3-D) finite element method models and compare model vertical surface deformation rates with observed GPS uplift rates outside the CV. Groundwater level changes are observed through well water elevation changes and through the resultant surface deformation (subsidence) by interferometric synthetic aperture radar (InSAR) and through broader scale changes in gravity from the GRACE satellite time variable gravity data [Famiglietti et al., 2011] that constrain the overall water volume changes. Combining InSAR with well-water data we are able to estimate the aquifer skeletal elastic and inelastic response and through a linear inversion derive the water volume (load) changes across the Central Valley and compare them with GRACE-inferred groundwater changes. Preliminary 3-D finite element method modeling that considers elastic and viscosity structure in the lithosphere gives three interesting results: 1) elastic models poorly fit the uplift rates near the SAF; 2) viscoelastic models that simulate different unloading histories show the past history of groundwater unloading has significant residual uplift rates and fault stress changes; 3) Coulomb stress change varies from inhibited on the locked (Carrizo) section to promoted on the creeping section of the SAF north of Parkfield. Thus, 3D models that account for lithosphere rheology, loading history viscous relaxation, have significant implications for longer-term time-dependent deformation, stress perturbation, and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.T42A..01B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.T42A..01B"><span>New slip rate estimates for the Mission Creek strand of the San <span class="hlt">Andreas</span> fault zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blisniuk, K.; Scharer, K. M.; Sharp, W. D.; Burgmann, R.; Rymer, M. J.; Williams, P. L.</p> <p>2013-12-01</p> <p>The potential for a large-magnitude earthquake (Mw ≥ 6.7) on the southern San <span class="hlt">Andreas</span> fault zone (SAFZ) is generally considered high (Working Group on California Earthquake Probabilities, 2007). However, the proportion of slip accommodated by each of its three major fault strands (Mission Creek, Banning, and Garnet Hill, from north to south) in the Indio Hills is poorly constrained. Each of these strands cut through San Gorgonio Pass west to the Los Angeles metropolitan region. To better assess the relative importance of these faults and their potential for a major earthquake, we dated offsets at two sites on the Mission Creek fault in the central Indio Hills, an offset channel at Pushawalla Canyon and an offset debris cone at a small unnamed canyon located ~1.5 km farther southeast. Previous work on this strand at Biskra Palms, in the southern Indio Hills, demonstrated a slip rate between 12 and 22 mm/yr, with a preferred rate of 14-17 mm/yr (Behr et al., GSAB, 2010). It is generally assumed that the slip rate on the Mission Creek fault decreases northwestwards from Biskra Palms (e.g. Fumal et al., BSSA, 2002) towards these two sites in the central Indio Hills. However, our initial results from uranium-series dating of pedogenic carbonate and 10Be cosmogenic exposure dating of surface clasts from deposits offset 1.3-1.6 km since ~70 ka and 44-50 m since ~2.5 ka indicate that during the late Pleistocene and Holocene slip on the Mission Creek fault in the central Indio Hills has occurred at a relatively constant and unexpectedly high rate of ~20 mm/yr. Combined with published paleoseismic studies for the Mission Creek fault, which show an average earthquake recurrence interval of 225 years for the past 5 events since 900 AD (Fumal et al., 2002), these data imply an average slip-per-event of ~4.5 m. The last earthquake to rupture this section of the Mission Creek fault occurred over 300 years ago (ca. 1690), which indicates that ca. 5.0 to 7.5 m of strain may have</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997SPIE.2941...75B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997SPIE.2941...75B"><span>Applications of a digital darkroom in the forensic laboratory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bullard, Barry D.; Birge, Brian</p> <p>1997-02-01</p> <p>Through a joint agreement with the Indiana-<span class="hlt">Marion</span> County Forensic Laboratory Services Agency, the Institute for Forensic Imaging conducted a pilot program to investigate crime lab applications of a digital darkroom. IFI installed and staffed a state-of-the-art digital darkroom in the photography laboratory of the Indianapolis-<span class="hlt">Marion</span> County crime lab located at Indianapolis, Indiana. The darkroom consisted of several high resolution color digital cameras, image processing computer, dye sublimation continuous tone digital printers, and CD-ROM writer. This paper describes the use of the digital darkroom in several crime lab investigations conducted during the program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70164444','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70164444"><span>Along-strike variations in fault frictional properties along the San <span class="hlt">Andreas</span> Fault near Cholame, California from joint earthquake and low-frequency earthquake relocations</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Harrington, Rebecca M.; Cochran, Elizabeth S.; Griffiths, Emily M.; Zeng, Xiangfang; Thurber, Clifford H.</p> <p>2016-01-01</p> <p>Recent observations of low‐frequency earthquakes (LFEs) and tectonic tremor along the Parkfield–Cholame segment of the San <span class="hlt">Andreas</span> fault suggest slow‐slip earthquakes occur in a transition zone between the shallow fault, which accommodates slip by a combination of aseismic creep and earthquakes (<15  km depth), and the deep fault, which accommodates slip by stable sliding (>35  km depth). However, the spatial relationship between shallow earthquakes and LFEs remains unclear. Here, we present precise relocations of 34 earthquakes and 34 LFEs recorded during a temporary deployment of 13 broadband seismic stations from May 2010 to July 2011. We use the temporary array waveform data, along with data from permanent seismic stations and a new high‐resolution 3D velocity model, to illuminate the fine‐scale details of the seismicity distribution near Cholame and the relation to the distribution of LFEs. The depth of the boundary between earthquakes and LFE hypocenters changes along strike and roughly follows the 350°C isotherm, suggesting frictional behavior may be, in part, thermally controlled. We observe no overlap in the depth of earthquakes and LFEs, with an ∼5  km separation between the deepest earthquakes and shallowest LFEs. In addition, clustering in the relocated seismicity near the 2004 Mw 6.0 Parkfield earthquake hypocenter and near the northern boundary of the 1857 Mw 7.8 Fort Tejon rupture may highlight areas of frictional heterogeneities on the fault where earthquakes tend to nucleate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.T41A2092C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.T41A2092C"><span>Frictional and hydrologic behavior of the San <span class="hlt">Andreas</span> Fault: Insights from laboratory experiments on SAFOD cuttings and core</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Carpenter, B. M.; Marone, C.; Saffer, D. M.</p> <p>2010-12-01</p> <p>The debate concerning the apparent low strength of tectonic faults, including the San <span class="hlt">Andreas</span> Fault (SAF), continues to focus on: 1) low intrinsic friction resulting from mineralogy and/or fabric, and 2) decreased effective normal stress due to elevated pore pressure. Here we inform this debate with laboratory measurements of the frictional behavior and permeability of cuttings and core returned from the SAF at a vertical depth of 2.7 km. We conducted experiments on cuttings and core recovered during SAFOD Phase III drilling. All samples in this study are adjacent to and within the active fault zone penetrated at 10814.5 ft (3296m) measured depth in the SAFOD borehole. We sheared gouge samples composed of drilling cuttings in a double-direct shear configuration subject to true-triaxial loading under constant effective normal stress, confining pressure, and pore pressure. Intact wafers of material were sheared in a single-direct shear configuration under similar conditions of effective stress, confining pressure, and pore pressure. We also report on permeability measurements on intact wafers of wall rock and fault gouge prior to shearing. Initial results from experiments on cuttings show: 1) a weak fault (µ=~0.21) compared to the surrounding wall rock (µ=~0.35), 2) velocity strengthening behavior, (a-b > 0), consistent with aseismic slip, and 3) near zero healing rates in material from the active fault. XRD analysis on cuttings indicates the main mineralogical difference between fault rock and wall rock, is the presence of significant amounts of smectite within the fault rock. Taken together, the measured frictional behavior and clay mineral content suggest that the clay composition exhibits a basic control on fault behavior. Our results document the first direct evidence of weak material from an active fault at seismogenic depths. In addition, our results could explain why the SAF in central California fails aseismically and hosts only small earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S52B..05K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S52B..05K"><span>Creep avalanches on San <span class="hlt">Andreas</span> Fault and their underlying mechanism from 19 years of InSAR and seismicity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Khoshmanesh, M.; Shirzaei, M.</p> <p>2017-12-01</p> <p>Recent seismic and geodetic observations indicate that interseismic creep rate varies in both time and space. The spatial extent of creep determines the earthquake potential, while its temporal evolution, known as slow slip events (SSE), may trigger earthquakes. Although the conditions promoting fault creep are well-established, the mechanism for initiating self-sustaining and sometimes cyclic creep events is enigmatic. Here we investigate a time series of 19 years of surface deformation measured by radar interferometry between 1992 and 2011 along the Central San <span class="hlt">Andreas</span> Fault (CSAF) to constrain the temporal evolution of creep. We show that the creep rate along the CSAF has a sporadic behavior, quantified with a Gumbel-like probability distribution characterized by longer tail toward the extreme positive rates, which is signature of burst-like creep dynamics. Defining creep avalanches as clusters of isolated creep with rates exceeding the shearing rate of tectonic plates, we investigate the statistical properties of their size and length. We show that, similar to the frequency-magnitude distribution of seismic events, the distribution of potency estimated for creep avalanches along the CSAF follows a power law, dictated by the distribution of their along-strike lengths. We further show that an ensemble of concurrent creep avalanches which aseismically rupture isolated fault compartments form the semi-periodic SSEs observed along the CSAF. Using a rate and state friction model, we show that normal stress is temporally variable on the fault, and support this using seismic observations. We propose that, through a self-sustaining fault-valve behavior, compaction induced elevation of pore pressure within hydraulically isolated fault compartments, and subsequent frictional dilation is the cause for the observed episodic SSEs. We further suggest that the 2004 Parkfield Mw6 earthquake may have been triggered by the SSE on adjacent creeping segment, which increased Coulomb</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013GeoJI.195..130T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013GeoJI.195..130T"><span>Three-dimensional magnetotelluric inversion in practice—the electrical conductivity structure of the San <span class="hlt">Andreas</span> Fault in Central California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tietze, Kristina; Ritter, Oliver</p> <p>2013-10-01</p> <p>3-D inversion techniques have become a widely used tool in magnetotelluric (MT) data interpretation. However, with real data sets, many of the controlling factors for the outcome of 3-D inversion are little explored, such as alignment of the coordinate system, handling and influence of data errors and model regularization. Here we present 3-D inversion results of 169 MT sites from the central San <span class="hlt">Andreas</span> Fault in California. Previous extensive 2-D inversion and 3-D forward modelling of the data set revealed significant along-strike variation of the electrical conductivity structure. 3-D inversion can recover these features but only if the inversion parameters are tuned in accordance with the particularities of the data set. Based on synthetic 3-D data we explore the model space and test the impacts of a wide range of inversion settings. The tests showed that the recovery of a pronounced regional 2-D structure in inversion of the complete impedance tensor depends on the coordinate system. As interdependencies between data components are not considered in standard 3-D MT inversion codes, 2-D subsurface structures can vanish if data are not aligned with the regional strike direction. A priori models and data weighting, that is, how strongly individual components of the impedance tensor and/or vertical magnetic field transfer functions dominate the solution, are crucial controls for the outcome of 3-D inversion. If deviations from a prior model are heavily penalized, regularization is prone to result in erroneous and misleading 3-D inversion models, particularly in the presence of strong conductivity contrasts. A `good' overall rms misfit is often meaningless or misleading as a huge range of 3-D inversion results exist, all with similarly `acceptable' misfits but producing significantly differing images of the conductivity structures. Reliable and meaningful 3-D inversion models can only be recovered if data misfit is assessed systematically in the frequency</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.T21C0433T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.T21C0433T"><span>Elemental Geochemistry of Samples From Fault Segments of the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) Drill Hole</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tourscher, S. N.; Schleicher, A. M.; van der Pluijm, B. A.; Warr, L. N.</p> <p>2006-12-01</p> <p>Elemental geochemistry of mudrock samples from phase 2 drilling of the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) is presented from bore hole depths of 3066 m to 3169 m and from 3292 m to 3368 m, which contain a creeping section and main trace of the fault, respectively. In addition to preparation and analysis of whole rock sample, fault grains with neomineralized, polished surfaces were hand picked from well-washed whole rock samples, minimizing the potential contamination from drilling mud and steel shavings. The separated fractions were washed in deionized water, powdered using a mortar and pestle, and analyzed using an Inductively Coupled Plasma- Optical Emission Spectrometer for major and minor elements. Based on oxide data results, systematic differences in element concentrations are observed between the whole rock and fault rock. Two groupings of data points are distinguishable in the regions containing the main trace of the fault, a shallow part (3292- 3316 m) and a deeper section (3320-3368 m). Applying the isocon method, assuming Zr and Ti to be immobile elements in these samples, indicates a volume loss of more than 30 percent in the shallow part and about 23 percent in the deep part of the main trace. These changes are minimum estimates of fault-related volume loss, because the whole rock from drilling samples contains variable amount of fault rock as well. Minimum estimates for volume loss in the creeping section of the fault are more than 50 percent when using the isocon method, comparing whole rock to plucked fault rock. The majority of the volume loss in the fault rocks is due to the dissolution and loss of silica, potassium, aluminum, sodium and calcium, whereas (based on oxide data) the mineralized surfaces of fractures appear to be enriched in Fe and Mg. The large amount of element mobility within these fault traces suggests extensive circulation of hydrous fluids along fractures that was responsible for progressive dissolution and leaching</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA466336','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA466336"><span>Effects of Redundant Alerts on Platoon Leader Performance and Decision Making</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2007-04-01</p> <p>Effects of Redundant Alerts on Platoon Leader Performance and Decision Making by <span class="hlt">Andrea</span> S . Krausman, Rodger A. Pettitt, and Linda R. Elliott...and Decision Making <span class="hlt">Andrea</span> S . Krausman, Rodger A. Pettitt, and Linda R. Elliott Human Research and Engineering Directorate, ARL...Decision Making 5c. PROGRAM ELEMENT NUMBER 5d. PROJECT NUMBER 62716AH70 5e. TASK NUMBER 6. AUTHOR( S ) <span class="hlt">Andrea</span> S . Krausman, Rodger A</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70159233','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70159233"><span>Using a modified time-reverse imaging technique to locate low-frequency earthquakes on the San <span class="hlt">Andreas</span> Fault near Cholame, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Horstmann, Tobias; Harrington, Rebecca M.; Cochran, Elizabeth S.</p> <p>2015-01-01</p> <p>We present a new method to locate low-frequency earthquakes (LFEs) within tectonic tremor episodes based on time-reverse imaging techniques. The modified time-reverse imaging technique presented here is the first method that locates individual LFEs within tremor episodes within 5 km uncertainty without relying on high-amplitude P-wave arrivals and that produces similar hypocentral locations to methods that locate events by stacking hundreds of LFEs without having to assume event co-location. In contrast to classic time-reverse imaging algorithms, we implement a modification to the method that searches for phase coherence over a short time period rather than identifying the maximum amplitude of a superpositioned wavefield. The method is independent of amplitude and can help constrain event origin time. The method uses individual LFE origin times, but does not rely on a priori information on LFE templates and families.We apply the method to locate 34 individual LFEs within tremor episodes that occur between 2010 and 2011 on the San <span class="hlt">Andreas</span> Fault, near Cholame, California. Individual LFE location accuracies range from 2.6 to 5 km horizontally and 4.8 km vertically. Other methods that have been able to locate individual LFEs with accuracy of less than 5 km have mainly used large-amplitude events where a P-phase arrival can be identified. The method described here has the potential to locate a larger number of individual low-amplitude events with only the S-phase arrival. Location accuracy is controlled by the velocity model resolution and the wavelength of the dominant energy of the signal. Location results are also dependent on the number of stations used and are negligibly correlated with other factors such as the maximum gap in azimuthal coverage, source–station distance and signal-to-noise ratio.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.S23C4543X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.S23C4543X"><span>Imaging the Fine-Scale Structure of the San <span class="hlt">Andreas</span> Fault in the Northern Gabilan Range with Explosion and Earthquake Sources</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xin, H.; Thurber, C. H.; Zhang, H.; Wang, F.</p> <p>2014-12-01</p> <p>A number of geophysical studies have been carried out along the San <span class="hlt">Andreas</span> Fault (SAF) in the Northern Gabilan Range (NGR) with the purpose of characterizing in detail the fault zone structure. Previous seismic research has revealed the complex structure of the crustal volume in the NGR region in two-dimensions (Thurber et al., 1996, 1997), and there has been some work on the three-dimensional (3D) structure at a coarser scale (Lin and Roecker, 1997). In our study we use earthquake body-wave arrival times and differential times (P and S) and explosion arrival times (only P) to image the 3D P- and S-wave velocity structure of the upper crust along the SAF in the NGR using double-difference (DD) tomography. The earthquake and explosion data types have complementary strengths - the earthquake data have good resolution at depth and resolve both Vp and Vs structure, although only where there are sufficient seismic rays between hypocenter and stations, whereas the explosions contribute very good near-surface resolution but for P waves only. The original dataset analyzed by Thurber et al. (1996, 1997) included data from 77 local earthquakes and 8 explosions. We enlarge the dataset with 114 more earthquakes that occurred in the study area, obtain improved S-wave picks using an automated picker, and include absolute and cross-correlation differential times. The inversion code we use is the algorithm tomoDD (Zhang and Thurber, 2003). We assess how the P and S velocity models and earthquake locations vary as we alter the inversion parameters and the inversion grid. The new inversion results show clearly the fine-scale structure of the SAF at depth in 3D, sharpening the image of the velocity contrast from the southwest side to the northeast side.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA481600','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA481600"><span>Detection and Localization of Vibrotactile Signals in Moving Vehicles</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2008-05-01</p> <p>Detection and Localization of Vibrotactile Signals in Moving Vehicles by <span class="hlt">Andrea</span> S . Krausman and Timothy L. White ARL-TR-4463 May 2008...Proving Ground, MD 21005-5425 ARL-TR-4463 May 2008 Detection and Localization of Vibrotactile Signals in Moving Vehicles <span class="hlt">Andrea</span> S ...5e. TASK NUMBER 6. AUTHOR( S ) <span class="hlt">Andrea</span> S . Krausman and Timothy L. White (both of ARL) 5f. WORK UNIT NUMBER 7. PERFORMING ORGANIZATION</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeoJI.205.1326L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeoJI.205.1326L"><span>Automatic identification of fault zone head waves and direct P waves and its application in the Parkfield section of the San <span class="hlt">Andreas</span> Fault, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Zefeng; Peng, Zhigang</p> <p>2016-06-01</p> <p>Fault zone head waves (FZHWs) are observed along major strike-slip faults and can provide high-resolution imaging of fault interface properties at seismogenic depth. In this paper, we present a new method to automatically detect FZHWs and pick direct P waves secondary arrivals (DWSAs). The algorithm identifies FZHWs by computing the amplitude ratios between the potential FZHWs and DSWAs. The polarities, polarizations and characteristic periods of FZHWs and DSWAs are then used to refine the picks or evaluate the pick quality. We apply the method to the Parkfield section of the San <span class="hlt">Andreas</span> Fault where FZHWs have been identified before by manual picks. We compare results from automatically and manually picked arrivals and find general agreement between them. The obtained velocity contrast at Parkfield is generally 5-10 per cent near Middle Mountain while it decreases below 5 per cent near Gold Hill. We also find many FZHWs recorded by the stations within 1 km of the background seismicity (i.e. the Southwest Fracture Zone) that have not been reported before. These FZHWs could be generated within a relatively wide low velocity zone sandwiched between the fast Salinian block on the southwest side and the slow Franciscan Mélange on the northeast side. Station FROB on the southwest (fast) side also recorded a small portion of weak precursory signals before sharp P waves. However, the polarities of weak signals are consistent with the right-lateral strike-slip mechanisms, suggesting that they are unlikely genuine FZHW signals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T21B2808K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T21B2808K"><span>Evaluating the Possibility of a joint San <span class="hlt">Andreas</span>-Imperial Fault Rupture in the Salton Trough Region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kyriakopoulos, C.; Oglesby, D. D.; Meltzner, A. J.; Rockwell, T. K.</p> <p>2016-12-01</p> <p>A geodynamic investigation of possible earthquakes in a given region requires both field data and numerical simulations. In particular, the investigation of past earthquakes is also a fundamental part of understanding the earthquake potential of the Salton Trough region. Geological records from paleoseismic trenches inform us of past ruptures (length, magnitude, timing), while dynamic rupture models allow us to evaluate numerically the mechanics of such earthquakes. The two most recent events (Mw 6.4 1940 and Mw 6.9 1979) on the Imperial fault (IF) both ruptured up to the northern end of the mapped fault, giving the impression that rupture doesn't propagate further north. This result is supported by small displacements, 20 cm, measured at the Dogwood site near the end of the mapped rupture in each event. However, 3D paleoseismic data from the same site corresponding to the most recent pre-1940 event (1710 CE) and 5th (1635 CE) and 6th events back revealed up to 1.5 m of slip in those events. Since we expect the surface displacement to decrease toward the termination of a rupture, we postulate that in these earlier cases the rupture propagated further north than in 1940 or 1979. Furthermore, paleoseismic data from the Coachella site (Philibosian et al., 2011) on the San <span class="hlt">Andreas</span> fault (SAF) indicates slip events ca. 1710 CE and 1588-1662 CE. In other words, the timing of two large paleoseismic displacements on the IF cannot be distinguished from the timing of the two most recent events on the southern SAF, leaving a question: is it possible to have through-going rupture in the Salton Trough? We investigate this question through 3D dynamic finite element rupture modeling. In our work, we considered two scenarios: rupture initiated on the IF propagating northward, and rupture initiated on the SAF propagating southward. Initial results show that, in the first case, rupture propagates north of the mapped northern terminus of the IF only under certain pre</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.T41C0634D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.T41C0634D"><span>Fault zone structure and kinematics from lidar, radar, and imagery: revealing new details along the creeping San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>DeLong, S.; Donnellan, A.; Pickering, A.</p> <p>2017-12-01</p> <p>Aseismic fault creep, coseismic fault displacement, distributed deformation, and the relative contribution of each have important bearing on infrastructure resilience, risk reduction, and the study of earthquake physics. Furthermore, the impact of interseismic fault creep in rupture propagation scenarios, and its impact and consequently on fault segmentation and maximum earthquake magnitudes, is poorly resolved in current rupture forecast models. The creeping section of the San <span class="hlt">Andreas</span> Fault (SAF) in Central California is an outstanding area for establishing methodology for future scientific response to damaging earthquakes and for characterizing the fine details of crustal deformation. Here, we describe how data from airborne and terrestrial laser scanning, airborne interferometric radar (UAVSAR), and optical data from satellites and UAVs can be used to characterize rates and map patterns of deformation within fault zones of varying complexity and geomorphic expression. We are evaluating laser point cloud processing, photogrammetric structure from motion, radar interferometry, sub-pixel correlation, and other techniques to characterize the relative ability of each to measure crustal deformation in two and three dimensions through time. We are collecting new and synthesizing existing data from the zone of highest interseismic creep rates along the SAF where a transition from a single main fault trace to a 1-km wide extensional stepover occurs. In the stepover region, creep measurements from alignment arrays 100 meters long across the main fault trace reveal lower rates than those in adjacent, geomorphically simpler parts of the fault. This indicates that deformation is distributed across the en echelon subsidiary faults, by creep and/or stick-slip behavior. Our objectives are to better understand how deformation is partitioned across a fault damage zone, how it is accommodated in the shallow subsurface, and to better characterize the relative amounts of fault creep</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Tectp.719...66M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Tectp.719...66M"><span>Stratigraphic record of Pliocene-Pleistocene basin evolution and deformation within the Southern San <span class="hlt">Andreas</span> Fault Zone, Mecca Hills, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McNabb, James C.; Dorsey, Rebecca J.; Housen, Bernard A.; Dimitroff, Cassidy W.; Messé, Graham T.</p> <p>2017-11-01</p> <p>A thick section of Pliocene-Pleistocene nonmarine sedimentary rocks exposed in the Mecca Hills, California, provides a record of fault-zone evolution along the Coachella Valley segment of the San <span class="hlt">Andreas</span> fault (SAF). Geologic mapping, measured sections, detailed sedimentology, and paleomagnetic data document a 3-5 Myr history of deformation and sedimentation in this area. SW-side down offset on the Painted Canyon fault (PCF) starting 3.7 Ma resulted in deposition of the Mecca Conglomerate southwest of the fault. The lower member of the Palm Spring Formation accumulated across the PCF from 3.0 to 2.6 Ma during regional subsidence. SW-side up slip on the PCF and related transpressive deformation from 2.6 to 2.3 Ma created a time-transgressive angular unconformity between the lower and upper members of the Palm Spring Formation. The upper member accumulated in discrete fault-bounded depocenters until initiation of modern deformation, uplift, and basin inversion starting at 0.7 Ma. Some spatially restricted deposits can be attributed to the evolution of fault-zone geometric complexities. However, the deformation events at ca. 2.6 Ma and 0.7 Ma are recorded regionally along 80 km of the SAF through Coachella Valley, covering an area much larger than mapped fault-zone irregularities, and thus require regional explanations. We therefore conclude that late Cenozoic deformation and sedimentation along the SAF in Coachella Valley has been controlled by a combination of regional tectonic drivers and local deformation due to dextral slip through fault-zone complexities. We further propose a kinematic link between the 2.6-2.3 Ma angular unconformity and a previously documented but poorly dated reorganization of plate-boundary faults in the northern Gulf of California at 3.3-2.0 Ma. This analysis highlights the potential for high-precision chronologies in deformed terrestrial deposits to provide improved understanding of local- to regional-scale structural controls on basin</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2004/5206/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2004/5206/"><span>Thrust-induced collapse of mountains-an example from the "Big Bend" region of the San <span class="hlt">Andreas</span> Fault, western transverse ranges, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kellogg, Karl S.</p> <p>2005-01-01</p> <p>Mount Pinos and Frazier Mountain are two prominent mountains just south of the San <span class="hlt">Andreas</span> fault in the western Transverse Ranges of southern California, a region that has undergone rapid Quaternary contraction and uplift. Both mountains are underlain, at least in part, by thrusts that place granitic and gneissic rocks over sedimentary rocks as young as Pliocene. Broad profiles and nearly flat summits of each mountain have previously been interpreted as relicts of a raised erosion surface. However, several features bring this interpretation into question. First, lag or stream gravels do not mantle the summit surfaces. Second, extensive landslide deposits, mostly pre?Holocene and deeply incised, mantle the flanks of both mountains. Third, a pervasive fracture and crushed?rock network pervades the crystalline rocks underlying both mountains. The orientation of the fractures, prominent in roadcuts on Mount Pinos, is essentially random. 'Hill?and?saddle' morphology characterizes ridges radiating from the summits, especially on Mount Pinos; outcrops are sparse on the hills and are nonexistent in the saddles, suggesting fractures are concentrated in the saddles. Latest movement on the thrusts underlying the two mountain massifs is probably early Quaternary, during which the mountains were uplifted to considerably higher (although unknown) elevations than at present. A model proposes that during thrusting, ground accelerations in the hanging wall, particularly near thrust tips, were high enough to pervasively fracture the hanging?wall rocks, thereby weakening them and producing essentially an assemblage of loose blocks. Movement over flexures in the fault surface accentuated fracturing. The lowered shear stresses necessary for failure, coupled with deep dissection and ongoing seismic activity, reduced gravitational potential by spreading the mountain massifs, triggering flanking landslides and producing broad, flat?topped mountains. This study developed from mapping in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..1213775G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..1213775G"><span>Seismic reflection images of the central California coast ranges and the tremor region of the San-<span class="hlt">Andreas</span>-Fault system near Cholame</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gutjahr, Stine; Buske, Stefan</p> <p>2010-05-01</p> <p>The SJ-6 seismic reflection profile was acquired in 1981 over a distance of about 180 km from Morro Bay to the Sierra Nevada foothills in South Central California. The profile runs across several prominent fault systems, e.g. the Riconada Fault (RF) in the western part as well as the San <span class="hlt">Andreas</span> Fault (SAF) in its central part. The latter includes the region of increased tremor activity near Cholame, as reported recently by several authors. We have recorrelated the original field data to 26 seconds two-way traveltime which allows us to image the crust and uppermost mantle down to approximately 40 km depth. A 3D tomographic velocity model derived from local earthquake data (Thurber et al., 2006) was used and Kirchhoff prestack depth migration as well as Fresnel-Volume-Migration were applied to the data set. Both imaging techniques were implemented in 3D by taking into account the true shot and receiver locations. The imaged subsurface volume itself was divided into three separate parts to correctly account for the significant kink in the profile line near the SAF. The most prominent features in the resulting images are areas of high reflectivity down to 30 km depth in particular in the central western part of the profile corresponding to the Salinian Block between the RF and the SAF. In the southwestern part strong reflectors can be identified that are dipping slightly to the northeast at depths of around 15-25 km. The eastern part consists of west dipping sediments at depths of 2-10 km that form a syncline structure in the west of the eastern part. The resulting images are compared to existing interpretations (Trehu and Wheeler, 1987; Wentworth and Zoback, 1989; Bloch et al., 1993) and discussed in the frame of the suggested tremor locations in that area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-06-22/pdf/2011-15624.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-06-22/pdf/2011-15624.pdf"><span>76 FR 36558 - Illinois; Amendment No. 1 to Notice of a Major Disaster Declaration</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-06-22</p> <p>... disaster by the President in his declaration of June 7, 2011. Alexander, Franklin, Gallatin, Hardin... Public Assistance (already designated for Individual Assistance). Hamilton, Jefferson, <span class="hlt">Marion</span>, Union...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70047748','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70047748"><span>Chemical controls on fault behavior: weakening of serpentinite sheared against quartz-bearing rocks and its significance for fault creep in the San <span class="hlt">Andreas</span> system</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Moore, Diane E.; Lockner, David A.</p> <p>2013-01-01</p> <p>The serpentinized ultramafic rocks found in many plate-tectonic settings commonly are juxtaposed against crustal rocks along faults, and the chemical contrast between the rock types potentially could influence the mechanical behavior of such faults. To investigate this possibility, we conducted triaxial experiments under hydrothermal conditions (200-350°C), shearing serpentinite gouge between forcing blocks of granite or quartzite. In an ultramafic chemical environment, the coefficient of friction, µ, of lizardite and antigorite serpentinite is 0.5-0.6, and µ increases with increasing temperature over the tested range. However, when either lizardite or antigorite serpentinite is sheared against granite or quartzite, strength is reduced to µ ~ 0.3, with the greatest strength reductions at the highest temperatures (temperature weakening) and slowest shearing rates (velocity strengthening). The weakening is attributed to a solution-transfer process that is promoted by the enhanced solubility of serpentine in pore fluids whose chemistry has been modified by interaction with the quartzose wall rocks. The operation of this process will promote aseismic slip (creep) along serpentinite-bearing crustal faults at otherwise seismogenic depths. During short-term experiments serpentine minerals reprecipitate in low-stress areas, whereas in longer experiments new Mg-rich phyllosilicates crystallize in response to metasomatic exchanges across the serpentinite-crustal rock contact. Long-term shear of serpentinite against crustal rocks will cause the metasomatic mineral assemblages, which may include extremely weak minerals such as saponite or talc, to play an increasingly important role in the mechanical behavior of the fault. Our results may explain the distribution of creep on faults in the San <span class="hlt">Andreas</span> system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.eatright.org/food/nutrition/nutrition-facts-and-food-labels/teach-your-teen-about-nutrition-facts-panels','NIH-MEDLINEPLUS'); return false;" href="https://www.eatright.org/food/nutrition/nutrition-facts-and-food-labels/teach-your-teen-about-nutrition-facts-panels"><span>Teach Your Teen about Food Labels</span></a></p> <p><a target="_blank" href="http://medlineplus.gov/">MedlinePlus</a></p> <p></p> <p></p> <p>... Your Teen about Nutrition Facts Panels By <span class="hlt">Andrea</span> Johnson, RD, CSP, LDN Published September 8, 2017 Antonio_ ... it will pay off! Reviewed July 2017 <span class="hlt">Andrea</span> Johnson, RD, CSP, LDN, is a clinical dietitian at ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T43D3077W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T43D3077W"><span>Testing the shorter and variable recurrence interval hypothesis along the Cholame segment of the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Williams, A.; Arrowsmith, R.; Rockwell, T. K.; Akciz, S. O.; Grant Ludwig, L.</p> <p>2016-12-01</p> <p>The Cholame segment of the San <span class="hlt">Andreas</span> Fault interacts with the Parkfield segment to the northwest with its creep and M6 earthquakes and the locked Carrizo segment to the southeast. Although offset reconstructions exist for this 75 km reach, rupture behavior is poorly characterized, limiting seismic hazard evaluation. Here we present new paleoseismic results from 2 fault perpendicular 26 m long trenches connected by a 15 m long fault parallel trench. The site is located south of the Parkfield segment 20 km southeast of Highway 46. Site geomorphology is characterized by several 50 m offset drainages northwest of the trenches, small shutter ridges and sag ponds, and alluvial fans crossing the fault. Fault zone stratigraphy consists of alternating finely bedded sands, silts, and gravels, and bioturbated soil horizons. The strata record 3-4 earthquakes and possible deformation post-1857, similar to the LY4 site 38 km southeast. E4, E3 and the most recent earthquake (MRE) are well supported by evidence of decreasing vertical offset up-sequence, capped fissure fill and colluvial wedges, which create small horst and graben structures. Units display vertical offsets ranging from 60 cm at the base to 12 cm near the MRE horizon, small colluvial wedges, and sag deposits within the 4 m wide fault zone. E2—the penultimate-is less certain, supported only by the decreasing offset in the stratigraphic sequence. The E4 event horizon is a gradational clayey silt sag deposit capped by discontinuous gravel, 18 cm at its thickest point and extending 4.8 m across the fault zone. The E3 and E2 event horizons are capped by thin bedded silty clay, and bounded by discontinuous burn horizons. The MRE horizon extends 6 m across the main fault zone, and consists of a silty clay sag deposit capped by very fine, bedded sand and coarse gravel, 22 cm at its thickest point and overlying a burn horizon. If the MRE is indeed the 1857 event, it has significant potential in correlation with the high</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2013-03-05/pdf/2013-04974.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2013-03-05/pdf/2013-04974.pdf"><span>78 FR 14356 - National Register of Historic Places; Notification of Pending Nominations and Related Actions</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2013-03-05</p> <p>... SW. Marconi Ave., Portland, 13000119 RHODE ISLAND Providence County Edgewood Historic District--Shaw Plat, (Edgewood Neighborhood, Cranston, R.I. MPS) Shaw & <span class="hlt">Marion</span> Aves., parts of Narragansett Blvd...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70026358','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70026358"><span>New insights on stress rotations from a forward regional model of the San <span class="hlt">Andreas</span> fault system near its Big Bend in southern California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fitzenz, D.D.; Miller, S.A.</p> <p>2004-01-01</p> <p>Understanding the stress field surrounding and driving active fault systems is an important component of mechanistic seismic hazard assessment. We develop and present results from a time-forward three-dimensional (3-D) model of the San <span class="hlt">Andreas</span> fault system near its Big Bend in southern California. The model boundary conditions are assessed by comparing model and observed tectonic regimes. The model of earthquake generation along two fault segments is used to target measurable properties (e.g., stress orientations, heat flow) that may allow inferences on the stress state on the faults. It is a quasi-static model, where GPS-constrained tectonic loading drives faults modeled as mostly sealed viscoelastic bodies embedded in an elastic half-space subjected to compaction and shear creep. A transpressive tectonic regime develops southwest of the model bend as a result of the tectonic loading and migrates toward the bend because of fault slip. The strength of the model faults is assessed on the basis of stress orientations, stress drop, and overpressures, showing a departure in the behavior of 3-D finite faults compared to models of 1-D or homogeneous infinite faults. At a smaller scale, stress transfers from fault slip transiently induce significant perturbations in the local stress tensors (where the slip profile is very heterogeneous). These stress rotations disappear when subsequent model earthquakes smooth the slip profile. Maps of maximum absolute shear stress emphasize both that (1) future models should include a more continuous representation of the faults and (2) that hydrostatically pressured intact rock is very difficult to break when no material weakness is considered. Copyright 2004 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990JGR....95.1139E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990JGR....95.1139E"><span>Crustal strain near the Big Bend of the San <span class="hlt">Andreas</span> Fault: Analysis of the Los Padres-Tehachapi Trilateration Networks, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eberhart-Phillips, Donna; Lisowski, Michael; Zoback, Mark D.</p> <p>1990-02-01</p> <p>In the region of the Los Padres-Tehachapi geodetic network, the San <span class="hlt">Andreas</span> fault (SAF) changes its orientation by over 30° from N40°W, close to that predicted by plate motion for a transform boundary, to N73°W. The strain orientation near the SAF is consistent with right-lateral shear along the fault, with maximum shear rate of 0.38±0.01 μrad/yr at N63°W. In contrast, away from the SAF the strain orientations on both sides of the fault are consistent with the plate motion direction, with maximum shear rate of 0.19±0.01 μrad/yr at N44°W. The strain rate does not drop off rapidly away from the fault, and thus the area is fit by either a broad shear zone below the SAF or a single fault with a relatively deep locking depth. The fit to the line length data is poor for locking depth d less than 25 km. For d of 25 km a buried slip rate of 30 ± 6 mm/yr is estimated. We also estimated buried slip for models that included the Garlock and Big Pine faults, in addition to the SAF. Slip rates on other faults are poorly constrained by the Los Padres-Tehachapi network. The best fitting Garlock fault model had computed left-lateral slip rate of 11±2 mm/yr below 10 km. Buried left-lateral slip of 15±6 mm/yr on the Big Pine fault, within the Western Transverse Ranges, provides significant reduction in line length residuals; however, deformation there may be more complicated than a single vertical fault. A subhorizontal detachment on the southern side of the SAF cannot be well constrained by these data. We investigated the location of the SAF and found that a vertical fault below the surface trace fits the data much better than either a dipping fault or a fault zone located south of the surface trace.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.S33A0849N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.S33A0849N"><span>Imaging Stress Transients and Fault Zone Processes with Crosswell Continuous Active-Source Seismic Monitoring at the San <span class="hlt">Andreas</span> Fault Observatory at Depth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Niu, F.; Taira, T.; Daley, T. M.; Marchesini, P.; Robertson, M.; Wood, T.</p> <p>2017-12-01</p> <p>Recent field and laboratory experiments identify seismic velocity changes preceding microearthquakes and rock failure (Niu et al., 2008, Nature; Scuderi et al., 2016, NatureGeo), which indicates that a continuous monitoring of seismic velocity might provide a mean of understanding of the earthquake nucleation process. Crosswell Continuous Active-Source Seismic Monitoring (CASSM) using borehole sources and sensors has proven to be an effective tool for measurements of seismic velocity and its temporal variation at seismogenic depth (Silver, et al, 2007, BSSA; Daley, et al, 2007, Geophysics). To expand current efforts on the CASSM development, in June 2017 we have begun to conduct a year-long CASSM field experiment at the San <span class="hlt">Andreas</span> Fault Observatory at Depth (SAFOD) in which the preceding field experiment detected the two sudden velocity reductions approximately 10 and 2 hours before microearthquakes (Niu et al., 2008, Nature). We installed a piezoelectric source and a three-component accelerometer at the SAFOD pilot and main holes ( 1 km depth) respectively. A seismic pulse was fired from the piezoelectric source four times per second. Each waveform was recorded 150-ms-long data with a sampling rate of 48 kHz. During this one-year experiment, we expect to have 10-15 microearthquakes (magnitude 1-3) occurring near the SAFOD site, and the data collected from the new experiment would allow us to further explore a relation between velocity changes and the Parkfield seismicity. Additionally, the year-long data provide a unique opportunity to study long-term velocity changes that might be related to seasonal stress variations at Parkfield (Johnson et al., 2017, Science). We will report on initial results of the SAFOD CASSM experiment and operational experiences of the CASSM development.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-08-05/pdf/2011-19908.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-08-05/pdf/2011-19908.pdf"><span>76 FR 47637 - Kansas Disaster #KS-00055</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-08-05</p> <p>... disaster declaration on 07/29/2011, Private Non- Profit organizations that provide essential services of... Counties: Barton, Clay, Cloud, Hamilton, Jewell, Lincoln, Logan, Lyon, <span class="hlt">Marion</span>, Mitchell, Morton, Osage...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0407.photos.379585p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0407.photos.379585p/"><span>11. Photo image looking northwest of the relocation office and ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>11. Photo image looking northwest of the relocation office and other administrative buildings. - Fort Benjamin Harrison, East Fifty-sixth Street (Aultman Avenue) & Glenn Road, Lawrence, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AGUFM.T24B..03T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AGUFM.T24B..03T"><span>Strength of the San <span class="hlt">Andreas</span> Fault Zone: Insight From SAFOD Cuttings and Core</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tembe, S.; Lockner, D. A.; Solum, J. G.; Morrow, C. A.; Wong, T.; Moore, D. E.</p> <p>2005-12-01</p> <p>Cuttings acquired during drilling of the SAFOD scientific hole near Parkfield, California offer a continuous physical record of the lithology across the San <span class="hlt">Andreas</span> fault (SAF) zone and provide the only complete set of samples available for laboratory testing. Guided by XRD clay mineral analysis and velocity and gamma logs, we selected washed cuttings from depths spanning the main hole from 1.85 to 3.0 km true vertical depth. Cuttings were chosen to represent primary lithologic units as well as significant shear zones, including candidates for the currently active SAF. To determine frictional properties triaxial sliding tests were conducted on cylindrical granite blocks containing sawcuts inclined at 30° and filled with 1 mm-thick sample gouge layers. Tests were run at constant effective normal stresses of 10 and 40 MPa and constant pore pressure of 1 MPa. Samples were sheared up to 10.4 mm at room temperature and velocities of 1, 0.1 and 0.01 μm/s. Stable sliding behavior and overall strain hardening were observed in all tests. The coefficient of friction typically showed a modest decrease with increasing effective normal stress and mostly velocity strengthening was observed. Preliminary results yield coefficients of friction, μ, which generally fell into two clusters spanning the range of 0.45 to 0.8. The higher values of friction (~0.7 - 0.8) corresponded to quartzofeldspathic samples derived from granodiorites and arkoses encountered in the drill hole. Lower values of friction (0.45 - 0.55) were observed at depth intervals interpreted as shear zones based on enriched clay content, reduced seismic velocities and increased gamma radiation. Arguments for a weak SAF suggest coseismic frictional strength of μ = 0.1 to 0.2 yet the actual fault zone materials studied here appear consistently stronger. At least two important limitations exist for inferring in-situ fault strength from cuttings. (1) Clays and weak minerals are preferentially lost during drilling and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title7-vol10/pdf/CFR-2012-title7-vol10-sec1210-501.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title7-vol10/pdf/CFR-2012-title7-vol10-sec1210-501.pdf"><span>7 CFR 1210.501 - Realignment of districts.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2012&page.go=Go">Code of Federal Regulations, 2012 CFR</a></p> <p></p> <p>2012-01-01</p> <p>..., Gulf, Hamilton, Hernando, Holmes, Jackson, Jefferson, Lafayette, Leon, Levy, Liberty, Madison, <span class="hlt">Marion</span>..., Virginia, Vermont, Wisconsin, West Virginia, and Washington, DC. (e) District 5—The State of California. (f...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title7-vol10/pdf/CFR-2014-title7-vol10-sec1210-501.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title7-vol10/pdf/CFR-2014-title7-vol10-sec1210-501.pdf"><span>7 CFR 1210.501 - Realignment of districts.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-01-01</p> <p>..., Gulf, Hamilton, Hernando, Holmes, Jackson, Jefferson, Lafayette, Leon, Levy, Liberty, Madison, <span class="hlt">Marion</span>..., Virginia, Vermont, Wisconsin, West Virginia, and Washington, DC. (e) District 5—The State of California. (f...</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title7-vol10/pdf/CFR-2013-title7-vol10-sec1210-501.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title7-vol10/pdf/CFR-2013-title7-vol10-sec1210-501.pdf"><span>7 CFR 1210.501 - Realignment of districts.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2013&page.go=Go">Code of Federal Regulations, 2013 CFR</a></p> <p></p> <p>2013-01-01</p> <p>..., Gulf, Hamilton, Hernando, Holmes, Jackson, Jefferson, Lafayette, Leon, Levy, Liberty, Madison, <span class="hlt">Marion</span>..., Virginia, Vermont, Wisconsin, West Virginia, and Washington, DC. (e) District 5—The State of California. (f...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020516','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020516"><span>Evolution of the Gorda Escarpment, San <span class="hlt">Andreas</span> fault and Mendocino triple junction from multichannel seismic data collected across the northern Vizcaino block, offshore northern California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Godfrey, N.J.; Meltzer, A.S.; Klemperer, S.L.; Trehu, A.M.; Leitner, B.; Clarke, S.H.; Ondrus, A.</p> <p>1998-01-01</p> <p>The Gorda Escarpment is a north facing scarp immediately south of the Mendocino transform fault (the Gorda/Juan de Fuca-Pacific plate boundary) between 126??W and the Mendocino triple junction. It elevates the seafloor at the northern edge of the Vizcaino block, part of the Pacific plate, ??? 1.5 km above the seafloor of the Gorda/Juan de Fuca plate to the north. Stratigraphy interpreted from multichannel seismic data across and close to the Gorda Escarpment suggests that the escarpment is a relatively recent pop-up feature caused by north-south compression across the plate boundary. Close to 126??W. the Vizcaino block acoustic basement shallows and is overlain by sediments that thin north toward the Gorda Escarpment. These sediments are tilted south and truncated at the seafloor. By contrast, in a localized region at the eastern end of the Gorda Escarpment, close to the Mendocino triple junction, the top of acoustic basement dips north and is overlain by a 2-km-thick wedge of pre-11 Ma sedimentary rocks that thickens north, toward the Gorda Escarpment. This wedge of sediments is restricted to the northeast corner of the Vizcaino block. Unless the wedge of sediments was a preexisting feature on the Vizcaino block before it was transferred from the North American to the Pacific plate, the strong spatial correlation between the sedimentary wedge and the triple junction suggests the entire Vizcaino block, with the San <span class="hlt">Andreas</span> at its eastern boundary, has been part of the Pacific plate since significantly before 11 Ma.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/al0682.photos.006301p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/al0682.photos.006301p/"><span>24. Historic American Buildings Survey Alex Bush, Photographer, January 9, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>24. Historic American Buildings Survey Alex Bush, Photographer, January 9, 1937 INTERIOR VIEW OF N. W. CORNER SMOKE HOUSE - Kenworthy Hall, State Highway 14 (Greensboro Road), <span class="hlt">Marion</span>, Perry County, AL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060436p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060436p/"><span>2. OVERALL VIEW OF THE MEDICAL CENTER (ESPECIALLY BUILDING 1) ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>2. OVERALL VIEW OF THE MEDICAL CENTER (ESPECIALLY BUILDING 1) FROM THE 'CAMPUS' GROUNDS; LOOKING SW. (Harms) - Veterans Administration Medical Center, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2010-09-09/pdf/2010-22215.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2010-09-09/pdf/2010-22215.pdf"><span>75 FR 54914 - Notice Pursuant to the National Cooperative Research and Production Act of 1993-Robotics...</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2010-09-09</p> <p>...; Esys Integration Corporation, Auburn Hills, MI; JADI, Inc., Troy, MI; Mobile Robots Inc., Amherst, NH... Alto, CA; Robot Worx, <span class="hlt">Marion</span>, OH; RPU Technology, Inc., Needham, MA; Scientific Systems Company, Inc...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0059.photos.065204p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0059.photos.065204p/"><span>32. Historic American Buildings Survey Photocopy of 1927 Bretzman Photograph, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>32. Historic American Buildings Survey Photocopy of 1927 Bretzman Photograph, Courtesy Tislow, Hunter & Associates, Inc. NORTHEAST CORNER OF COSMETIC ROOM - Indiana Theatre, 134 West Washington Street, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890064818&hterms=rock+islands&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DThe%2Brock%2Bislands','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890064818&hterms=rock+islands&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DThe%2Brock%2Bislands"><span>Kaersutite-bearing xenoliths and megacrysts in volcanic rocks from the Funk Seamount in the souhtwest Indian Ocean</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reid, Arch M.; Le Roex, Anton P.</p> <p>1988-01-01</p> <p>The petrography, mineral chemistry, and whole-rock compositions of volcanic rocks dredged from the Funk Seamount, located 60 km NW of <span class="hlt">Marion</span> Island in the southwestern Indian Ocean, are presented together with the mineral chemistry of their inclusions. On the basis of these characteristics, the possible relationships between the Funk Seamount's volcanic rocks and the megacrysts and xenoliths in these rocks are discussed. It is argued that the Funk Seamount lavas derive from a similar mantle source region as that of the <span class="hlt">Marion</span> Island and Prince Edward Island hotspot lavas. The geochemical signature of these lavas implies derivation from a source that is enriched (e.g., in Ti, K, P, and Nb) over the depleted mantle source regions for the adjacent mid-ocean ridge basalts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226092p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226092p/"><span>GENERAL PERSPECTIVE VIEW OF THE "KEEP OREGON GREEN" ASSOCIATION BUILDING ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>GENERAL PERSPECTIVE VIEW OF THE "KEEP OREGON GREEN" ASSOCIATION BUILDING (K.O.G. BUILDING), VIEW LOOKING SOUTHWEST ACROSS STATE STREET. - Oregon State Forester's Office Complex, 2600 State Street, Salem, <span class="hlt">Marion</span>, OR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-05-17/pdf/2011-12051.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-05-17/pdf/2011-12051.pdf"><span>76 FR 28435 - Change in Bank Control Notices; Acquisitions of Shares of a Bank or Bank Holding Company</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-05-17</p> <p>... June 1, 2011. A. Federal Reserve Bank of Minneapolis (Jacqueline G. King, Community Affairs Officer) 90.... Short, University Heights, Ohio; Colleen V. Short, Edina, Minnesota; and the trustees (<span class="hlt">Marion</span> D. Short...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-05-26/pdf/2011-13030.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-05-26/pdf/2011-13030.pdf"><span>76 FR 30737 - Arkansas; Amendment No. 3 to Notice of a Major Disaster Declaration</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-05-26</p> <p>..., Office of Response and Recovery, Federal Emergency Management Agency, 500 C Street, SW., Washington, DC..., <span class="hlt">Marion</span>, Nevada, Newton, Perry, Pike, Polk, Searcy, Sharp, Van Buren, White, and Yell Counties for Public...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhDT.......146T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhDT.......146T"><span>Transform Faults and Lithospheric Structure: Insights from Numerical Models and Shipboard and Geodetic Observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takeuchi, Christopher S.</p> <p></p> <p>In this dissertation, I study the influence of transform faults on the structure and deformation of the lithosphere, using shipboard and geodetic observations as well as numerical experiments. I use marine topography, gravity, and magnetics to examine the effects of the large age-offset Andrew Bain transform fault on accretionary processes within two adjacent segments of the Southwest Indian Ridge. I infer from morphology, high gravity, and low magnetization that the extremely cold and thick lithosphere associated with the Andrew Bain strongly suppresses melt production and crustal emplacement to the west of the transform fault. These effects are counteracted by enhanced temperature and melt production near the <span class="hlt">Marion</span> Hotspot, east of the transform fault. I use numerical models to study the development of lithospheric shear zones underneath continental transform faults (e.g. the San <span class="hlt">Andreas</span> Fault in California), with a particular focus on thermomechanical coupling and shear heating produced by long-term fault slip. I find that these processes may give rise to long-lived localized shear zones, and that such shear zones may in part control the magnitude of stress in the lithosphere. Localized ductile shear participates in both interseismic loading and postseismic relaxation, and predictions of models including shear zones are within observational constraints provided by geodetic and surface heat flow data. I numerically investigate the effects of shear zones on three-dimensional postseismic deformation. I conclude that the presence of a thermally-activated shear zone minimally impacts postseismic deformation, and that thermomechanical coupling alone is unable to generate sufficient localization for postseismic relaxation within a ductile shear zone to kinematically resemble that by aseismic fault creep (afterslip). I find that the current record geodetic observations of postseismic deformation do not provide robust discriminating power between candidate linear and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226045p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226045p/"><span>GENERAL PERSPECTIVE VIEW OF THE COMPLEX, VIEW LOOKING NORTHWEST AT ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>GENERAL PERSPECTIVE VIEW OF THE COMPLEX, VIEW LOOKING NORTHWEST AT THE OREGON FORESTER'S OFFICE BUILDING, WITH MILL CREEK IN THE FOREGROUND. - Oregon State Forester's Office Complex, 2600 State Street, Salem, <span class="hlt">Marion</span>, OR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060438p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060438p/"><span>4. OVERALL VIEW OF THE MEDICAL CENTER (ESPECIALLY BUILDING 2 ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. OVERALL VIEW OF THE MEDICAL CENTER (ESPECIALLY BUILDING 2 & BUILDING 1) FROM THE 'CAMPUS' GROUNDS; LOOKING NE. (Harms) - Veterans Administration Medical Center, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060437p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0575.photos.060437p/"><span>3. OVERALL VIEW OF THE MEDICAL CENTER (ESPECIALLY BUILDING 2 ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>3. OVERALL VIEW OF THE MEDICAL CENTER (ESPECIALLY BUILDING 2 & BUILDING 1) FROM THE 'CAMPUS' GROUNDS; LOOKING NW. (Harms) - Veterans Administration Medical Center, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0382.photos.379531p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0382.photos.379531p/"><span>1. General view looking east from the south side of ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>1. General view looking east from the south side of the 300 block of West Washington Street - Washington Street Downtown Corridor, Washington Street between Alabama Street & Senate Avenue, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0059.photos.065203p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0059.photos.065203p/"><span>31. Historic American Buildings Survey Photocopy of 1927 Bretzman Photograph, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>31. Historic American Buildings Survey Photocopy of 1927 Bretzman Photograph, Courtesy Tislow, Hunter & Associates, Inc. WOMEN'S LOUNGE, LOOKING EAST INTO COSMETIC ROOM - Indiana Theatre, 134 West Washington Street, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.T21B2815D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.T21B2815D"><span>Southern San <span class="hlt">Andreas</span> Fault Slip History Refined Using Pliocene Colorado River Deposits in the Western Salton Trough</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dorsey, R. J.; Bennett, S. E. K.; Housen, B. A.</p> <p>2016-12-01</p> <p>Tectonic reconstructions of Pacific-North America plate motion in the Salton Trough region (Bennett et al., 2016) are constrained by: (1) late Miocene volcanic rocks that record 255 +/-10 km of transform offset across the northern Gulf of California since 6 Ma (average 42 mm/yr; Oskin and Stock, 2003); and (2) GPS data that show modern rates of 50-52 mm/yr between Pacific and North America plates, and 46-48 mm/yr between Baja California (BC) and North America (NAM) (Plattner et al., 2007). New data from Pliocene Colorado River deposits in the Salton Trough provide an important additional constraint on the geologic history of slip on the southern San <span class="hlt">Andreas</span> Fault (SAF). The Arroyo Diablo Formation (ADF) in the San Felipe Hills SW of the Salton Sea contains abundant cross-bedded channel sandstones deformed in the dextral Clark fault zone. The ADF ranges in age from 4.3 to 2.8 Ma in the Fish Creek-Vallecito basin, and in the Borrego Badlands its upper contact with the Borrego Formation is 2.9 Ma based on our new magnetostratigraphy. ADF paleocurrent data from a 20-km wide, NW-oriented belt near Salton City record overall transport to the SW (corrected for bedding dip, N=165), with directions ranging from NW to SE. Spatial domain analysis reveals radial divergence of paleoflow to the: W and NW in the NW domain; SW in the central domain; and S in the SE domain. Data near Borrego Sink, which restores to south of Salton City after removing offset on the San Jacinto fault zone, show overall transport to the SE. Pliocene patterns of radial paleoflow divergence strongly resemble downstream bifurcation of fluvial distributary channels on the modern Colorado River delta SW of Yuma, and indicate that Salton City has translated 120-130 km NW along the SAF since 3 Ma. We propose a model in which post-6 Ma BC-NAM relative motion gradually accelerated to 50 mm/yr by 4 Ma, continued at 50 mm/yr from 4-1 Ma, and decreased to 46 mm/yr from 1-0 Ma (split equally between the SAF and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70041940','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70041940"><span>Timing of large earthquakes during the past 500 years along the Santa Cruz Mountains segment of the San <span class="hlt">Andreas</span> fault at Mill Canyon, near Watsonville, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fumal, Thomas E.</p> <p>2012-01-01</p> <p>A paleoseismic investigation across the Santa Cruz Mountains section of the San <span class="hlt">Andreas</span> fault at Mill Canyon indicates that four surface‐rupturing earthquakes have occurred there during the past ~500  years. At this site, right‐lateral fault slip has moved a low shutter ridge across the mouth of the canyon, ponding latest Holocene sediments. These alluvial deposits are deformed along a narrow zone of faulting. There is excellent evidence for a 1906 (M 7.8) and three earlier earthquakes consisting of well‐developed fissures, scarps, and colluvial wedges. Deformation resulting from the earlier earthquakes is comparable to that from 1906, suggesting they also were large‐magnitude events. The earthquake prior to 1906 occurred either about A.D. 1750 (1711–1770) or A.D. 1855 (1789–1904), depending on assumptions incorporated into two alternative OxCal models. If the later age range is correct, then the earthquake may have been a historical early‐to‐mid‐nineteenth‐century earthquake, possibly the A.D. 1838 earthquake. Both models are viable, and there is no way to select one over the other with the available data. Two earlier earthquakes occurred about A.D. 1690 (1660–1720) and A.D. 1522 (1454–1605). Using OxCal, recalculation of the age of the reported penultimate earthquake reported from the Grizzly Flat site, located about 10 km northwest of Mill Canyon, indicates it occurred about A.D. 1105–1545, earlier than any of the past three earthquakes, and possibly correlates to the fourth earthquake at Mill Canyon.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0053.photos.065154p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0053.photos.065154p/"><span>7. Historic American Buildings Survey Photocopy taken from Sanborn Map ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>7. Historic American Buildings Survey Photocopy taken from Sanborn Map at Assessor's Office City-County Building, Indianapolis, Indiana PLAN OF FACTORY - Duesenberg Automobile Company Factory, West Washington & Harding Streets, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226047p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226047p/"><span>GENERAL PERSPECTIVE VIEW OF THE COMPLEX, VIEW LOOKING WEST AT ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>GENERAL PERSPECTIVE VIEW OF THE COMPLEX, VIEW LOOKING WEST AT THE MOTORPOOL, AUTOMOTIVE SHOP AND FOREST PROTECTION BUILDING (FROM LEFT TO RIGHT). - Oregon State Forester's Office Complex, 2600 State Street, Salem, <span class="hlt">Marion</span>, OR</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0466.photos.581274p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0466.photos.581274p/"><span>SECTION 1, WITH BIVOUAC OF THE DEAD TABLET IN FOREGROUND ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>SECTION 1, WITH BIVOUAC OF THE DEAD TABLET IN FOREGROUND AND FLAGPOLE IN BACKGROUND. VIEW TO EAST. - Crown Hill Cemetery, Crown Hill National Cemetery, 700 West Thirty-eighth Street, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol3/pdf/CFR-2010-title40-vol3-sec52-776.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol3/pdf/CFR-2010-title40-vol3-sec52-776.pdf"><span>40 CFR 52.776 - Control strategy: Particulate matter.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-07-01</p> <p>... Approval—The complete Indiana plan for Clark, Dearborn, Dubois, <span class="hlt">Marion</span> (except for coke batteries), St..., Vandenburgh County; 6-1-17, Clark County; 6-1-18, St. Joseph County; 6-2, Particulate Emissions Limitations...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol3/pdf/CFR-2011-title40-vol3-sec52-776.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol3/pdf/CFR-2011-title40-vol3-sec52-776.pdf"><span>40 CFR 52.776 - Control strategy: Particulate matter.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... Approval—The complete Indiana plan for Clark, Dearborn, Dubois, <span class="hlt">Marion</span> (except for coke batteries), St..., Vandenburgh County; 6-1-17, Clark County; 6-1-18, St. Joseph County; 6-2, Particulate Emissions Limitations...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2010-10-08/pdf/2010-25410.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2010-10-08/pdf/2010-25410.pdf"><span>75 FR 62363 - Agency Information Collection Activities: Proposed Collection; Comment Request-Supplemental...</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2010-10-08</p> <p>... other forms of information technology. Comments may be sent to: <span class="hlt">Andrea</span> Gold, Chief, Retailer Management... INFORMATION CONTACT: <span class="hlt">Andrea</span> Gold, Chief, Retailer Management and Issuance Branch at (703) 305-2456... Reconciliation. Form Number: FNS-46. OMB Number: 0584-0080. Expiration Date: February 28, 2011. Type of Request...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060451p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060451p/"><span>12. DETAIL VIEW OF DEDICATION PLAQUE & ITS SURROUNDING TABERNACLE ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>12. DETAIL VIEW OF DEDICATION PLAQUE & ITS SURROUNDING TABERNACLE FRAME; MAIN LOBBY OF BUILDING 1, LOOKING E. (Ryan) - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0382.photos.379574p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0382.photos.379574p/"><span>44. Detail view looking east along sidewalk on the south ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>44. Detail view looking east along sidewalk on the south side of the 300 block of East Washington Street - Washington Street Downtown Corridor, Washington Street between Alabama Street & Senate Avenue, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.epa.gov/in/american-lead-action-memorandum','PESTICIDES'); return false;" href="https://www.epa.gov/in/american-lead-action-memorandum"><span>American Lead Action Memorandum</span></a></p> <p><a target="_blank" href="http://www.epa.gov/pesticides/search.htm">EPA Pesticide Factsheets</a></p> <p></p> <p></p> <p>ACTION MEMORANDUM— Request for a Time-Critical Removal Action andExemption from the $2 Million and 12-Month Statutory Limits at the AmericanLead Site, Indianapolis, <span class="hlt">Marion</span> County, Indiana (Site ID #B56J)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Tecto..36.2863M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Tecto..36.2863M"><span>Two-Phase Exhumation of the Santa Rosa Mountains: Low- and High-Angle Normal Faulting During Initiation and Evolution of the Southern San <span class="hlt">Andreas</span> Fault System</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mason, Cody C.; Spotila, James A.; Axen, Gary; Dorsey, Rebecca J.; Luther, Amy; Stockli, Daniel F.</p> <p>2017-12-01</p> <p>Low-angle detachment fault systems are important elements of oblique-divergent plate boundaries, yet the role detachment faulting plays in the development of such boundaries is poorly understood. The West Salton Detachment Fault (WSDF) is a major low-angle normal fault that formed coeval with localization of the Pacific-North America plate boundary in the northern Salton Trough, CA. Apatite U-Th/He thermochronometry (AHe; <fi>n</fi> = 29 samples) and thermal history modeling of samples from the Santa Rosa Mountains (SRM) reveal that initial exhumation along the WSDF began at circa 8 Ma, exhuming footwall material from depths of >2 to 3 km. An uplifted fossil (Miocene) helium partial retention zone is present in the eastern SRM, while a deeper crustal section has been exhumed along the Pleistocene high-angle Santa Rosa Fault (SFR) to much higher elevations in the southwest SRM. Detachment-related vertical exhumation rates in the SRM were 0.15-0.36 km/Myr, with maximum fault slip rates of 1.2-3.0 km/Myr. Miocene AHe isochrons across the SRM are consistent with northeast crustal tilting of the SRM block and suggest that the post-WSDF vertical exhumation rate along the SRF was 1.3 km/Myr. The timing of extension initiation in the Salton Trough suggests that clockwise rotation of relative plate motions that began at 8 Ma is associated with initiation of the southern San <span class="hlt">Andreas</span> system. Pleistocene regional tectonic reorganization was contemporaneous with an abrupt transition from low- to high-angle faulting and indicates that local fault geometry may at times exert a fundamental control on rock uplift rates along strike-slip fault systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/pa4056.photos.223005p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/pa4056.photos.223005p/"><span>PERSPECTIVE VIEW FROM SOUTHWEST OF (LEFT TO RIGHT): 1903 SERVANTS' ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>PERSPECTIVE VIEW FROM SOUTHWEST OF (LEFT TO RIGHT): 1903 SERVANTS' QUARTERS, 1824 STONE KITCHEN, 1789 BRICK HOUSE (BEHIND BALCONY), AND 1901 SOUTH BEDROOM WING. - Friendship Hill, 223 New Geneva Road, Point <span class="hlt">Marion</span>, Fayette County, PA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226048p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226048p/"><span>GENERAL PERSPECTIVE VIEW OF THE COMPLEX, VIEW LOOKING EAST AT ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>GENERAL PERSPECTIVE VIEW OF THE COMPLEX, VIEW LOOKING EAST AT THE FOREST PROTECTION BUILDING, THE MOTORPOOL, AND THE AUTOMOTIVE SHOP (FROM LEFT TO RIGHT). - Oregon State Forester's Office Complex, 2600 State Street, Salem, <span class="hlt">Marion</span>, OR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0045.photos.065141p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0045.photos.065141p/"><span>8. Historic American Buildings Survey PHOTOCOPY CA. 1916 OF SLEEPING ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>8. Historic American Buildings Survey PHOTOCOPY CA. 1916 OF SLEEPING PORCH, AT NORTH END CAPABLE OF SLEEPING SIX. Original in Collections of Marian College - James A. Allison Mansion, 3200 Coldspring Road, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-08-15/pdf/2011-20713.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-08-15/pdf/2011-20713.pdf"><span>76 FR 50443 - Proposed Flood Elevation Determinations</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-08-15</p> <p>... the of Hull, Town of intersection of Brant <span class="hlt">Marion</span>, Town of Beach Avenue and Ocean Mattapoisett, Town..., MA 02043. Town of Hull Maps are available for inspection at the Town Hall, 253 Atlantic Avenue, Hull...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-09-21/pdf/2011-24160.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-09-21/pdf/2011-24160.pdf"><span>76 FR 58535 - National Register of Historic Places; Notification of Pending Nominations and Related Actions</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-09-21</p> <p>... Railroad District, (Post-World War II Development in Ottumwa, IA 1944-1959 MPS), Main St. to BNSFRR tracks... <span class="hlt">Marion</span> County 1927 Hillsboro Water Tower, Lots 10 & 11, Blk. 2, Hill's 2nd Addition., Hillsboro, 11000727...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2013-01-15/pdf/2013-00676.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2013-01-15/pdf/2013-00676.pdf"><span>78 FR 2950 - Opportunity To Comment on the Applicants for the Indianapolis, Indiana Area</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2013-01-15</p> <p>...; Johnson; Madison, west of State Route 13 and south of State Route 132; <span class="hlt">Marion</span>; Monroe; Morgan; and Shelby... Administration. [FR Doc. 2013-00676 Filed 1-14-13; 8:45 am] BILLING CODE 3410-KD-P ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018498','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018498"><span>1957 Gobi-Altay, Mongolia, earthquake as a prototype for southern California's most devastating earthquake</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Bayarsayhan, C.; Bayasgalan, A.; Enhtuvshin, B.; Hudnut, K.W.; Kurushin, R.A.; Molnar, P.; Olziybat, M.</p> <p>1996-01-01</p> <p>The 1957 Gobi-Altay earthquake was associated with both strike-slip and thrust faulting, processes similar to those along the San <span class="hlt">Andreas</span> fault and the faults bounding the San Gabriel Mountains just north of Los Angeles, California. Clearly, a major rupture either on the San <span class="hlt">Andreas</span> fault north of Los Angeles or on the thrust faults bounding the Los Angeles basin poses a serious hazard to inhabitants of that area. By analogy with the Gobi-Altay earthquake, we suggest that simultaneous rupturing of both the San <span class="hlt">Andreas</span> fault and the thrust faults nearer Los Angeles is a real possibility that amplifies the hazard posed by ruptures on either fault system separately.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.S21B2506R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.S21B2506R"><span>Permeability and of the San <span class="hlt">Andreas</span> Fault core and damage zone from SAFOD drill core</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rathbun, A. P.; Fry, M.; Kitajima, H.; Song, I.; Carpenter, B. M.; Marone, C.; Saffer, D. M.</p> <p>2012-12-01</p> <p>Quantifying fault-rock permeability is important toward understanding both the regional hydrologic behavior of fault zones, and poro-elastic processes that may affect faulting and earthquake mechanics by mediating effective stress. These include persistent fluid overpressures hypothesized to reduce fault strength, as well as dynamic processes that may occur during earthquake slip, including thermal pressurization and dilatancy hardening. To date, studies of permeability on fault rocks and gouge from plate-boundary strike-slip faults have mainly focused on samples from surface outcrops. We report on permeability tests conducted on the host rock, damage zone, and a major actively creeping fault strand (Central Deformation Zone, CDZ) of the San <span class="hlt">Andreas</span> Fault (SAF), obtained from coring across the active SAF at ~2.7 km depth as part of SAFOD Phase III. We quantify permeability on subsamples oriented both perpendicular and parallel to the coring axis, which is nearly perpendicular to the SAF plane, to evaluate permeability anisotropy. The fault strand samples were obtained from the CDZ, which accommodates significant creep, and hosts ~90% of the observed casing deformation measured between drilling phases. The CDZ is 2.6 m thick with a matrix grain size < 10 μm and ~5% vol. clasts, and contains ~80% clay, of which ~90% is smectite. We also tested damage zone samples taken from adjacent core sections within a few m on either side of the CDZ. Permeability experiments were conducted in a triaxial vessel, on samples 25.4 mm in diameter and ~20-35 mm in length. We conducted measurements under isotropic stress conditions, at effective stress (Pc') of ~5-70 MPa. We measure permeability using a constant head flow-through technique. At the highest Pc', low permeability of the CDZ and damage zone necessitates using a step loading transient method and is in good agreement with permeabilities obtained from flow-through experiments. We quantify compression behavior by monitoring</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.T41D1598F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.T41D1598F"><span>Transpressional deformation style and AMS fabrics adjacent to the southernmost segment of the San <span class="hlt">Andreas</span> fault, Durmid Hill, CA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>French, M.; Wojtal, S. F.; Housen, B.</p> <p>2006-12-01</p> <p>In the Salton Trough, the trace of the San <span class="hlt">Andreas</span> Fault (SAF) ends where it intersects the NNW-trending Brawley seismic zone at Durmid Hill (DH). The topographic relief of DH is a product of faulting and folding of Pleistocene Borrego Formation strata (Babcock, 1974). Burgmann's (1991) detailed mapping and analysis of the western part of DH showed that the folds and faults accommodate transpression. Key to Burgmann's work was the recognition that the ~2m thick Bishop Ash, a prominent marker horizon, has been elongated parallel to the hinges of folds and boudinaged. We are mapping in detail the eastern portion of DH, nearer to the trace of the SAF. Folds in the eastern part of DH are tighter and thrust faulting is more prominent, consistent with greater shortening magnitude oblique to the SAF. Boudinage of the ash layer again indicates elongation parallel to fold hinges and subparallel to the SAF. The Bishop Ash locally is <1m thick along fold limbs in eastern DH, suggesting that significant continuous deformation accompanied the development of map-scale features. We measured anisotropy of magnetic susceptibility (AMS) fabrics in the Bishop Ash in order to assess continuous deformation in the Ash at DH. Because the Bishop Ash at DH is altered, consisting mainly of silica glass and clay minerals, samples from DH have significantly lower magnetic susceptibilities than Bishop Ash samples from elsewhere in the Salton Trough. With such low susceptibilities, there is significant scatter in the orientation of magnetic foliation and lineation in our samples. Still, in some Bishop samples within 1 km of the SAF, magnetic foliation is consistent with fold-related flattening. Magnetic lineation in these samples is consistently sub-parallel to fold hinges, parallel to the elongation direction inferred from boudinage. Even close to the trace of the SAF, this correlation breaks down in map-scale zones where fold hinge lines change attitude, fold shapes change, and the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226042p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226042p/"><span>TOPOGRAPHIC VIEW OF THE STATE FORESTER'S COMPLEX, VIEW LOOKING SOUTHWEST ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>TOPOGRAPHIC VIEW OF THE STATE FORESTER'S COMPLEX, VIEW LOOKING SOUTHWEST FROM STATE STREET, WITH THE K.O.G BUILDING (KEEP OREGON GREEN) IN THE FOREGROUND. - Oregon State Forester's Office Complex, 2600 State Street, Salem, <span class="hlt">Marion</span>, OR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0407.photos.379583p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0407.photos.379583p/"><span>9. View looking south showing additional housing and administrative buildings. ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>9. View looking south showing additional housing and administrative buildings. Taken from same side of the street as the view above. - Fort Benjamin Harrison, East Fifty-sixth Street (Aultman Avenue) & Glenn Road, Lawrence, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ghr.nlm.nih.gov/condition/north-american-indian-childhood-cirrhosis','NIH-MEDLINEPLUS'); return false;" href="https://ghr.nlm.nih.gov/condition/north-american-indian-childhood-cirrhosis"><span>Genetics Home Reference: North American Indian childhood cirrhosis</span></a></p> <p><a target="_blank" href="http://medlineplus.gov/">MedlinePlus</a></p> <p></p> <p></p> <p>... Chagnon P, Fujiwara TM, Morgan K, Richter A, Hudson TJ, Mitchell GA. Localization of a recessive gene ... J, <span class="hlt">Marion</span> JF, Drouin E, Rasquin-Weber A, Hudson TJ, Richter A. A missense mutation (R565W) in ...</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0410.photos.319159p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0410.photos.319159p/"><span>4. AN IMAGE, LOOKING NORTH WEST, OF THE TRACK, BALLAST ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. AN IMAGE, LOOKING NORTH WEST, OF THE TRACK, BALLAST AND CONCRETE BALUSTRADES, SHOWING HOW THE BRIDGE WAS DESIGNED TO TAKE TWO SETS OF TRACK. - Vandalia Railroad Bridge, Spanning U.S. Route 40, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol5/pdf/CFR-2010-title47-vol5-sec90-614.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title47-vol5/pdf/CFR-2010-title47-vol5-sec90-614.pdf"><span>47 CFR 90.614 - Segments of the 806-824/851-869 MHz band for non-border areas.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-10-01</p> <p>..., Henry, Houston, Jackson, Jefferson, Lamar, Lauderdale, Lawrence, Lee, Limestone, Lowndes, Macon, Madison..., Jasper, Jeff Davis, Jefferson, Jenkins, Johnson, Jones, Lamar, Lanier, Laurens, Lee, Liberty, Lincoln..., Lawrence, Leake, Lee, Lincoln, Lowndes, Madison, <span class="hlt">Marion</span>, Monroe, Montgomery, Neshoba, Newton, Noxubee...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title47-vol5/pdf/CFR-2011-title47-vol5-sec90-614.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title47-vol5/pdf/CFR-2011-title47-vol5-sec90-614.pdf"><span>47 CFR 90.614 - Segments of the 806-824/851-869 MHz band for non-border areas.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-10-01</p> <p>..., Henry, Houston, Jackson, Jefferson, Lamar, Lauderdale, Lawrence, Lee, Limestone, Lowndes, Macon, Madison..., Jasper, Jeff Davis, Jefferson, Jenkins, Johnson, Jones, Lamar, Lanier, Laurens, Lee, Liberty, Lincoln..., Lawrence, Leake, Lee, Lincoln, Lowndes, Madison, <span class="hlt">Marion</span>, Monroe, Montgomery, Neshoba, Newton, Noxubee...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70019175','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019175"><span>Three-dimensional upper crustal velocity structure beneath San Francisco Peninsula, California</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Parsons, T.; Zoback, M.L.</p> <p>1997-01-01</p> <p>This paper presents new seismic data from, and crustal models of the San Francisco Peninsula. In much of central California the San <span class="hlt">Andreas</span> fault juxtaposes the Cretaceous granitic Salinian terrane on its west and the Late Mesozoic/Early Tertiary Franciscan Complex on its east. On San Francisco Peninsula, however, the present-day San <span class="hlt">Andreas</span> fault is completely within a Franciscan terrane, and the Pilarcitos fault, located southwest of the San <span class="hlt">Andreas</span>, marks the Salinian-Franciscan boundary. This circumstance has evoked two different explanations: either the Pilarcitos is a thrust fault that has pushed Franciscan rocks over Salinian rocks or the Pilarcitos is a transform fault that has accommodated significant right-lateral slip. In an effort to better resolve the subsurface structure of the peninsula faults, we established a temporary network of 31 seismographs arrayed across the San <span class="hlt">Andreas</span> fault and the subparallel Pilarcitos fault at ???1-2 km spacings. These instruments were deployed during the first 6 months of 1995 and recorded local earthquakes, air gun sources set off in San Francisco Bay, and explosive sources. Travel times from these sources were used to augment earthquake arrival times recorded by the Northern California Seismic Network and were inverted for three-dimensional velocity structure. Results show lateral velocity changes at depth (???0.5-7 km) that correlate with downward vertical projections of the surface traces of the San <span class="hlt">Andreas</span> and Pilarcitos faults. We thus interpret the faults as high-angle to vertical features (constrained to a 70??-110?? dip range). From this we conclude that the Pilarcitos fault is probably an important strike-slip fault that accommodated much of the right-lateral plate boundary strain on the peninsula prior to the initiation of the modern-day San <span class="hlt">Andreas</span> fault in this region sometime after about 3.0 m.y. ago.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226043p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226043p/"><span>TOPOGRAPHIC VIEW THE STATE FORESTER'S COMPLEX, VIEW LOOKING SOUTH FROM ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>TOPOGRAPHIC VIEW THE STATE FORESTER'S COMPLEX, VIEW LOOKING SOUTH FROM STATE STREET, WITH THE NATIONAL REGISTER-LISTED OREGON STATE FORESTER'S OFFICE BUILDING TO THE LEFT OF VIEW. - Oregon State Forester's Office Complex, 2600 State Street, Salem, <span class="hlt">Marion</span>, OR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226061p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/or0533.photos.226061p/"><span>PERSPECTIVE VIEW OF THE NATIONAL REGISTERLISTED OREGON STATE FORESTER'S OFFICE ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>PERSPECTIVE VIEW OF THE NATIONAL REGISTER-LISTED OREGON STATE FORESTER'S OFFICE BUILDING AND STONEMASONRY WALL AND FLAGSTONE SIDEWALK ALONG STATE STREET, VIEW LOOKING SOUTH FROM STATE STREET. - Oregon State Forester's Office Complex, 2600 State Street, Salem, <span class="hlt">Marion</span>, OR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/pa4056.photos.223006p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/pa4056.photos.223006p/"><span>PERSPECTIVE VIEW FROM SOUTHEAST OF (LEFT TO RIGHT): 1901 SOUTH ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>PERSPECTIVE VIEW FROM SOUTHEAST OF (LEFT TO RIGHT): 1901 SOUTH BEDROOM WING, 1789 BRICK HOUSE, 1798 FRAME HOUSE (ALL BEHIND TWO-STORY PORCH), AND 1823 STONE HOUSE . - Friendship Hill, 223 New Geneva Road, Point <span class="hlt">Marion</span>, Fayette County, PA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Henning&id=EJ1105413','ERIC'); return false;" href="https://eric.ed.gov/?q=Henning&id=EJ1105413"><span>I Remember Highlander</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Williams-Hawkins, Maria</p> <p>2016-01-01</p> <p>"I Remember Highlander" reflects on the life choices of <span class="hlt">Marion</span> Barry and Herman Henning Jr., fraternity brothers who sought the same goal but took different paths. The essay examines cultural and family situations that shaped lives and decisions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70155114','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70155114"><span>Tectonic activity as a significant source of crustal tetrafluoromethane emissions to the atmosphere: observations in groundwaters along the San <span class="hlt">Andreas</span> Fault</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Deeds, Daniel A.; Kulongoski, Justin T.; Muhle, Jens; Weiss, Ray F.</p> <p>2015-01-01</p> <p>Tetrafluoromethane (CF4) concentrations were measured in 14 groundwater samples from the Cuyama Valley, Mil Potrero and Cuddy Valley aquifers along the Big Bend section of the San <span class="hlt">Andreas</span> Fault System (SAFS) in California to assess whether tectonic activity in this region is a significant source of crustal CF4 to the atmosphere. Dissolved CF4 concentrations in all groundwater samples but one were elevated with respect to estimated recharge concentrations including entrainment of excess air during recharge (CreCre; ∼30 fmol kg−1 H2O), indicating subsurface addition of CF4 to these groundwaters. Groundwaters in the Cuyama Valley contain small CF4 excesses (0.1–9 times CreCre), which may be attributed to an in situ release from weathering and a minor addition of deep crustal CF4 introduced to the shallow groundwater through nearby faults. CF4 excesses in groundwaters within 200 m of the SAFS are larger (10–980 times CreCre) and indicate the presence of a deep crustal flux of CF4 that is likely associated with the physical alteration of silicate minerals in the shear zone of the SAFS. Extrapolating CF4 flux rates observed in this study to the full extent of the SAFS (1300 km × 20–100 km) suggests that the SAFS potentially emits (0.3–1)×10−1 kg(0.3–1)×10−1 kg CF4 yr−1 to the Earth's surface. For comparison, the chemical weathering of ∼7.5×104 km2∼7.5×104 km2 of granitic rock in California is estimated to release (0.019–3.2)×10−1 kg(0.019–3.2)×10−1 kg CF4 yr−1. Tectonic activity is likely an important, and potentially the dominant, driver of natural emissions of CF4 to the atmosphere. Variations in preindustrial atmospheric CF4 as observed in paleo-archives such as ice cores may therefore represent changes in both continental weathering and tectonic activity, including changes driven by variations in continental ice cover during glacial–interglacial transitions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70192474','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70192474"><span>Subsurface geometry of the San <span class="hlt">Andreas</span> fault in southern California: Results from the Salton Seismic Imaging Project (SSIP) and strong ground motion expectations</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fuis, Gary S.; Bauer, Klaus; Goldman, Mark R.; Ryberg, Trond; Langenheim, Victoria; Scheirer, Daniel S.; Rymer, Michael J.; Stock, Joann M.; Hole, John A.; Catchings, Rufus D.; Graves, Robert; Aagaard, Brad T.</p> <p>2017-01-01</p> <p>The San <span class="hlt">Andreas</span> fault (SAF) is one of the most studied strike‐slip faults in the world; yet its subsurface geometry is still uncertain in most locations. The Salton Seismic Imaging Project (SSIP) was undertaken to image the structure surrounding the SAF and also its subsurface geometry. We present SSIP studies at two locations in the Coachella Valley of the northern Salton trough. On our line 4, a fault‐crossing profile just north of the Salton Sea, sedimentary basin depth reaches 4 km southwest of the SAF. On our line 6, a fault‐crossing profile at the north end of the Coachella Valley, sedimentary basin depth is ∼2–3  km">∼2–3  km and centered on the central, most active trace of the SAF. Subsurface geometry of the SAF and nearby faults along these two lines is determined using a new method of seismic‐reflection imaging, combined with potential‐field studies and earthquakes. Below a 6–9 km depth range, the SAF dips ∼50°–60°">∼50°–60° NE, and above this depth range it dips more steeply. Nearby faults are also imaged in the upper 10 km, many of which dip steeply and project to mapped surface fault traces. These secondary faults may join the SAF at depths below about 10 km to form a flower‐like structure. In Appendix D, we show that rupture on a northeast‐dipping SAF, using a single plane that approximates the two dips seen in our study, produces shaking that differs from shaking calculated for the Great California ShakeOut, for which the southern SAF was modeled as vertical in most places: shorter‐period (T<1  s">T<1  s) shaking is increased locally by up to a factor of 2 on the hanging wall and is decreased locally by up to a factor of 2 on the footwall, compared to shaking calculated for a vertical fault.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=motion+AND+comic&id=ED070280','ERIC'); return false;" href="https://eric.ed.gov/?q=motion+AND+comic&id=ED070280"><span>Off With Their Heads! A Serio-Comic Tale of Hollywood.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Marion, Frances</p> <p></p> <p>Frances <span class="hlt">Marion</span> was one of Hollywood's leading screenplay writers from the silent era through the 1940's. This informal memoir concentrates on the personalities she knew, the development of major motion pictures, her friends, and her own career. (JK)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0696.photos.060473p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0696.photos.060473p/"><span>3. DETAIL VIEW OF THE MAIN ENTRY OF BUILDING 13, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>3. DETAIL VIEW OF THE MAIN ENTRY OF BUILDING 13, SHOWING THE ORIGINAL LIGHT FIXTURES AND THE EGYPTIAN MOTIF DECORATION; LOOKING SSW. (Ryan) - Veterans Administration Medical Center, Building No. 13, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://rosap.ntl.bts.gov/view/dot/34137','DOTNTL'); return false;" href="https://rosap.ntl.bts.gov/view/dot/34137"><span>Aggressive driving enforcement : evaluation of two demonstration programs</span></a></p> <p><a target="_blank" href="http://ntlsearch.bts.gov/tris/index.do">DOT National Transportation Integrated Search</a></p> <p></p> <p>2004-03-24</p> <p>This report presents the results of a study conducted to assess the effects of two programs that were implemented to reduce the incidence of aggressive driving. The programs were conducted by the <span class="hlt">Marion</span> County Traffic Safety Partnership (a consortium...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://rosap.ntl.bts.gov/view/dot/1718','DOTNTL'); return false;" href="https://rosap.ntl.bts.gov/view/dot/1718"><span>Aggressive driving enforcement : evaluations of two demonstration programs</span></a></p> <p><a target="_blank" href="http://ntlsearch.bts.gov/tris/index.do">DOT National Transportation Integrated Search</a></p> <p></p> <p>2004-03-01</p> <p>This report presents the results of a study conducted to assess the effects of two programs that were implemented to reduce the incidence of aggressive driving. The programs were conducted by the <span class="hlt">Marion</span> County Traffic Safety Partnership (a consortium...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title12-vol1/pdf/CFR-2014-title12-vol1-sec7-4007.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title12-vol1/pdf/CFR-2014-title12-vol1-sec7-4007.pdf"><span>12 CFR 7.4007 - Deposit-taking.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2014&page.go=Go">Code of Federal Regulations, 2014 CFR</a></p> <p></p> <p>2014-01-01</p> <p>... <span class="hlt">Marion</span> County, N.A. v. Nelson, Florida Insurance Commissioner, et al. 517 U.S. 25 (1996): (1) Contracts... al. 517 U.S. 25 (1996), or that is made applicable by Federal law. [69 FR 1916, Jan. 13, 2004, as...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title12-vol1/pdf/CFR-2012-title12-vol1-sec7-4007.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title12-vol1/pdf/CFR-2012-title12-vol1-sec7-4007.pdf"><span>12 CFR 7.4007 - Deposit-taking.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2012&page.go=Go">Code of Federal Regulations, 2012 CFR</a></p> <p></p> <p>2012-01-01</p> <p>... <span class="hlt">Marion</span> County, N.A. v. Nelson, Florida Insurance Commissioner, et al. 517 U.S. 25 (1996): (1) Contracts... al. 517 U.S. 25 (1996), or that is made applicable by Federal law. [69 FR 1916, Jan. 13, 2004, as...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title12-vol1/pdf/CFR-2013-title12-vol1-sec7-4007.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title12-vol1/pdf/CFR-2013-title12-vol1-sec7-4007.pdf"><span>12 CFR 7.4007 - Deposit-taking.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2013&page.go=Go">Code of Federal Regulations, 2013 CFR</a></p> <p></p> <p>2013-01-01</p> <p>... <span class="hlt">Marion</span> County, N.A. v. Nelson, Florida Insurance Commissioner, et al. 517 U.S. 25 (1996): (1) Contracts... al. 517 U.S. 25 (1996), or that is made applicable by Federal law. [69 FR 1916, Jan. 13, 2004, as...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.cancer.net/blog/2015-06/choosing-breast-prosthesis-survivors-perspective','NIH-MEDLINEPLUS'); return false;" href="https://www.cancer.net/blog/2015-06/choosing-breast-prosthesis-survivors-perspective"><span>Choosing a Breast Prosthesis: A Survivor's Perspective</span></a></p> <p><a target="_blank" href="http://medlineplus.gov/">MedlinePlus</a></p> <p></p> <p></p> <p>... Q: Do you have any tips for other women who might be trying to decide if a breast prosthesis is right for them? <span class="hlt">Andrea</span>: You can always meet with ... out what kind of surgical options might be right for you. Q: Is there anything else you think women with breast cancer should know? <span class="hlt">Andrea</span>: I think ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://ngmdb.usgs.gov/Prodesc/proddesc_68889.htm','USGSPUBS'); return false;" href="http://ngmdb.usgs.gov/Prodesc/proddesc_68889.htm"><span>Chapter E. The Loma Prieta, California, Earthquake of October 17, 1989 - Geologic Setting and Crustal Structure</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Wells, Ray E.</p> <p>2004-01-01</p> <p>Although some scientists considered the Ms=7.1 Loma Prieta, Calif., earthquake of 1989 to be an anticipated event, some aspects of the earthquake were surprising. It occurred 17 km beneath the Santa Cruz Mountains along a left-stepping restraining bend in the San <span class="hlt">Andreas</span> fault system. Rupture on the southwest-dipping fault plane consisted of subequal amounts of right-lateral and reverse motion but did not reach the surface. In the area of maximum uplift, severe shaking and numerous ground cracks occurred along Summit Road and Skyland Ridge, several kilometers south of the main trace of the San <span class="hlt">Andreas</span> fault. The relatively deep focus of the earthquake, the distribution of ground failure, the absence of throughgoing surface rupture on the San <span class="hlt">Andreas</span> fault, and the large component of uplift raised several questions about the relation of the 1989 Loma Prieta earthquake to the San <span class="hlt">Andreas</span> fault: Did the earthquake actually occur on the San <span class="hlt">Andreas</span> fault? Where exactly is the San <span class="hlt">Andreas</span> fault in the heavily forested Santa Cruz Mountains, and how does the fault relate to ground ruptures that occurred there in 1989 and 1906? What is the geometry of the San <span class="hlt">Andreas</span> fault system at depth, and how does it relate to the major crustal blocks identified by geologic mapping? Subsequent geophysical and geologic investigations of crustal structure in the Loma Prieta region have addressed these and other questions about the relation of the earthquake to geologic structures observed in the southern Santa Cruz Mountains. The diverse papers in this chapter cover several topics: geologic mapping of the region, potential- field and electromagnetic modeling of crustal structure, and the velocity structure of the crust and mantle in and below the source region for the earthquake. Although these papers were mostly completed between 1992 and 1997, they provide critical documentation of the crustal structure of the Loma Prieta region. Together, they present a remarkably coherent, three</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020136','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020136"><span>Seismotectonics of the Loma Prieta, California, region determined from three-dimensional Vp, Vp/Vs, and seismicity</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Eberhart-Phillips, D.; Michael, A.J.</p> <p>1998-01-01</p> <p>Three-dimensional Vp and Vp/Vs velocity models for the Loma Prieta region were developed from the inversion of local travel time data (21,925 P arrivals and 1,116 S arrivals) from earthquakes, refraction shots, and blasts recorded on 1700 stations from the Northern California Seismic Network and numerous portable seismograph deployments. The velocity and density models and microearthquake hypocenters reveal a complex structure that includes a San <span class="hlt">Andreas</span> fault extending to the base of the seismogenic layer. A body with high Vp extends the length of the rupture and fills the 5 km wide volume between the Loma Prieta mainshock rupture and the San <span class="hlt">Andreas</span> and Sargent faults. We suggest that this body controls both the pattern of background seismicity on the San <span class="hlt">Andreas</span> and Sargent faults and the extent of rupture during the mainshock, thus explaining how the background seismicity outlined the along-strike and depth extent of the mainshock rupture on a different fault plane 5 km away. New aftershock focal mechanisms, based on three-dimensional ray tracing through the velocity model, support a heterogeneous postseismic stress field and can not resolve a uniform fault normal compression. The subvertical (or steeply dipping) San <span class="hlt">Andreas</span> fault and the fault surfaces that ruptured in the 1989 Loma Prieta earthquake are both parts of the San <span class="hlt">Andreas</span> fault zone and this section of the fault zone does not have a single type of characteristic event.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.T51J..03D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.T51J..03D"><span>Identifying Fault Connections of the Southern Pacific-North American Plate Boundary Using Triggered Slip and Crustal Velocities</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Donnellan, A.; Grant Ludwig, L.; Rundle, J. B.; Parker, J. W.; Granat, R.; Heflin, M. B.; Pierce, M. E.; Wang, J.; Gunson, M.; Lyzenga, G. A.</p> <p>2017-12-01</p> <p>The 2010 M7.2 El Mayor - Cucapah earthquake caused extensive triggering of slip on faults proximal to the Salton Trough in southern California. Triggered slip and postseismic motions that have continued for over five years following the earthquake highlight connections between the El Mayor - Cucapah rupture and the network of faults that branch out along the southern Pacific - North American Plate Boundary. Coseismic triggering follows a network of conjugate faults from the northern end of the rupture to the Coachella segment of the southernmost San <span class="hlt">Andreas</span> fault. Larger aftershocks and postseismic motions favor connections to the San Jacinto and Elsinore faults further west. The 2012 Brawley Swarm can be considered part of the branching on the Imperial Valley or east side of the plate boundary. Cluster analysis of long-term GPS velocities using Lloyds Algorithm, identifies bifurcation of the Pacific - North American plate boundary; The San Jacinto fault joins with the southern San <span class="hlt">Andreas</span> fault, and the Salton Trough and Coachella segment of the San <span class="hlt">Andreas</span> fault join with the Eastern California Shear Zone. The clustering analysis does not identify throughgoing deformation connecting the Coachella segment of the San <span class="hlt">Andreas</span> fault with the rest of the San <span class="hlt">Andreas</span> fault system through the San Gorgonio Pass. This observation is consistent with triggered slip from both the 1992 Landers and 2010 El Mayor - Cucapah earthquakes that follows the plate boundary bifurcation and with paleoseismic evidence of smaller earthquakes in the San Gorgonio Pass.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.epa.gov/ia/country-club-estates-llc-clean-water-act-public-notice','PESTICIDES'); return false;" href="https://www.epa.gov/ia/country-club-estates-llc-clean-water-act-public-notice"><span>Country Club Estates, LLC - Clean Water Act Public Notice</span></a></p> <p><a target="_blank" href="http://www.epa.gov/pesticides/search.htm">EPA Pesticide Factsheets</a></p> <p></p> <p></p> <p>The EPA is providing notice of an Administrative Penalty Assessment in the form of an Expedited Storm Water Settlement Agreement against Country Club Estates, LLC, a business located at 3415 Mulberry Dr., <span class="hlt">Marion</span>, Iowa, for alleged violations at Tower Tenac</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=302112','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=302112"><span>Blackberry (Rubus spp.)-Virus Diseases</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>Many viruses have been found in blackberries in the Pacific Northwest. Blackberry calico virus (a carlavirus) is universally present in older commercial 'Thornless Loganberry' fields. Similar calico diseases occur in field-run '<span class="hlt">Marion</span>', 'Chehalem', 'Olallie', and 'Waldo' blackberries. Other virus di...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060442p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060442p/"><span>3. N ELEVATION OF BUILDING 1'S E WING, SHOWING THE ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>3. N ELEVATION OF BUILDING 1'S E WING, SHOWING THE PILASTERS, TERRA COTTA PANELS, AND THE EGYPTIAN MOTIF DECORATIVE CORNICE ELEMENTS; LOOKING S. (Ryan) - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0695.photos.060469p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0695.photos.060469p/"><span>3. DETAIL VIEW OF S ENTRY DOOR TO BUILDING 8, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>3. DETAIL VIEW OF S ENTRY DOOR TO BUILDING 8, SHOWING ONE OF THE ORIGINAL LIGHT FIXTURES AND THE EGYPTIAN MOTIF DECORATION; LOOKING N (Ryan) - Veterans Administration Medical Center, Building No. 8, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2012-09-05/pdf/2012-21782.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2012-09-05/pdf/2012-21782.pdf"><span>77 FR 54600 - Mississippi; Emergency and Related Determinations</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2012-09-05</p> <p>.... Specifically, you are authorized to provide assistance for emergency protective measures (Category B), limited..., Hinds, Jackson, Jasper, Jefferson Davis, Jones, Lamar, Lauderdale, Lawrence, Lincoln, Madison, <span class="hlt">Marion</span>... protective measures (Category B), limited to direct Federal assistance, under the Public Assistance program...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-201009130010HQ.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-201009130010HQ.html"><span>NASA Education Stakeholder's Summit</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2010-09-12</p> <p><span class="hlt">Marion</span> C. Blakey, President and CEO, Aerospace Industries Association, addresses guests at the NASA Education Stakeholders’ Summit One Stop Shopping Initiative (OSSI), Monday, Sep. 13, 2010, at the Westfields Marriott Conference Center in Chantilly, VA. (Photo Credit: NASA/Carla Cioffi)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/26350','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/26350"><span>Timber resources of northwest Oregon.</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Mary A. Mei</p> <p>1979-01-01</p> <p>This report presents statistics from a 1976 timber resource inventory of 10 counties in northwest Oregon: Clackamas, Clatsop, Columbia, Hood River, <span class="hlt">Marion</span>, Multnomah, Polk, Tillamook, Washington, and Yamhill. Tables presented are of forest area, timber volume, growth, mortality, and harvest.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/al0682.color.314266c/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/al0682.color.314266c/"><span>91. VIEW OF THE SOUTHWEST CORNER WHERE THE TOWER JOINS ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>91. VIEW OF THE SOUTHWEST CORNER WHERE THE TOWER JOINS THE WEST GABLE & THE BRICK STEPS LEAD UP TO A SMALL VERANDAH (DUPLICATE OF HABS No. AL-765-34) - Kenworthy Hall, State Highway 14 (Greensboro Road), <span class="hlt">Marion</span>, Perry County, AL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.T41A2870J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.T41A2870J"><span>Character and Implications of a Newly Identified Creeping Strand of the San <span class="hlt">Andreas</span> fault NE of Salton Sea, Southern California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Janecke, S. U.; Markowski, D.</p> <p>2015-12-01</p> <p>The overdue earthquake on the Coachella section, San <span class="hlt">Andreas</span> fault (SAF), the model ShakeOut earthquake, and the conflict between cross-fault models involving the Extra fault array and mapped shortening in the Durmid Hill area motivate new analyses at the southern SAF tip. Geologic mapping, LiDAR, seismic reflection, magnetic and gravity datasets, and aerial photography confirm the existence of the East Shoreline strand (ESS) of the SAF southwest of the main trace of the SAF. We mapped the 15 km long ESS, in a band northeast side of the Salton Sea. Other data suggest that the ESS continues N to the latitude of the Mecca Hills, and is >35 km long. The ESS cuts and folds upper Holocene beds and appears to creep, based on discovery of large NW-striking cracks in modern beach deposits. The two traces of the SAF are parallel and ~0.5 to ~2.5 km apart. Groups of east, SE, and ENE-striking strike-slip cross-faults connect the master dextral faults of the SAF. There are few sinistral-normal faults that could be part of the Extra fault array. The 1-km wide ESS contains short, discontinuous traces of NW-striking dextral-oblique faults. These en-echelon faults bound steeply dipping Pleistocene beds, cut out section, parallel tight NW-trending folds, and produced growth folds. Beds commonly dip toward the ESS on both sides, in accord with persistent NE-SW shortening across the ESS. The dispersed fault-fold structural style of the ESS is due to decollements in faulted mud-rich Pliocene to Holocene sediment and ramps and flats along the strike-slip faults. A sheared ladder-like geometric model of the two master dextral strands of the SAF and their intervening cross-faults, best explains the field relationships and geophysical datasets. Contraction across >40 km2 of the southernmost SAF zone in the Durmid Hills suggest that interaction of active structures in the SAF zone may inhibit the nucleation of large earthquakes in this region. The ESS may cross the northern Coachella</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRB..123..815S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRB..123..815S"><span>"3D_Fault_Offsets," a Matlab Code to Automatically Measure Lateral and Vertical Fault Offsets in Topographic Data: Application to San <span class="hlt">Andreas</span>, Owens Valley, and Hope Faults</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stewart, N.; Gaudemer, Y.; Manighetti, I.; Serreau, L.; Vincendeau, A.; Dominguez, S.; Mattéo, L.; Malavieille, J.</p> <p>2018-01-01</p> <p>Measuring fault offsets preserved at the ground surface is of primary importance to recover earthquake and long-term slip distributions and understand fault mechanics. The recent explosion of high-resolution topographic data, such as Lidar and photogrammetric digital elevation models, offers an unprecedented opportunity to measure dense collections of fault offsets. We have developed a new Matlab code, 3D_Fault_Offsets, to automate these measurements. In topographic data, 3D_Fault_Offsets mathematically identifies and represents nine of the most prominent geometric characteristics of common sublinear markers along faults (especially strike slip) in 3-D, such as the streambed (minimum elevation), top, free face and base of channel banks or scarps (minimum Laplacian, maximum gradient, and maximum Laplacian), and ridges (maximum elevation). By calculating best fit lines through the nine point clouds on either side of the fault, the code computes the lateral and vertical offsets between the piercing points of these lines onto the fault plane, providing nine lateral and nine vertical offset measures per marker. Through a Monte Carlo approach, the code calculates the total uncertainty on each offset. It then provides tools to statistically analyze the dense collection of measures and to reconstruct the prefaulted marker geometry in the horizontal and vertical planes. We applied 3D_Fault_Offsets to remeasure previously published offsets across 88 markers on the San <span class="hlt">Andreas</span>, Owens Valley, and Hope faults. We obtained 5,454 lateral and vertical offset measures. These automatic measures compare well to prior ones, field and remote, while their rich record provides new insights on the preservation of fault displacements in the morphology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.S53A1817Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.S53A1817Z"><span>Identification of repeating earthquakes and spatio-temporal variations of fault zone properties around the Parkfield section of the San <span class="hlt">Andreas</span> fault and the central Calaveras fault</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, P.; Peng, Z.</p> <p>2008-12-01</p> <p>We systemically identify repeating earthquakes and investigate spatio-temporal variations of fault zone properties associated with the 2004 Mw6.0 Parkfield earthquake along the Parkfield section of the San <span class="hlt">Andreas</span> fault, and the 1984 Mw6.2 Morgan Hill earthquake along the central Calaveras fault. The procedure for identifying repeating earthquakes is based on overlapping of the source regions and the waveform similarity, and is briefly described as follows. First, we estimate the source radius of each event based on a circular crack model and a normal stress drop of 3 MPa. Next, we compute inter-hypocentral distance for events listed in the relocated catalog of Thurber et al. (2006) around Parkfield, and Schaff et al. (2002) along the Calaveras fault. Then, we group all events into 'initial' clusters by requiring the separation distance between each event pair to be less than the source radius of larger event, and their magnitude difference to be less than 1. Next, we calculate the correlation coefficients between every event pair within each 'initial' cluster using a 3-s time window around the direct P waves for all available stations. The median value of the correlation coefficients is used as a measure of similarity between each event pair. We drop an event if the median similarity to the rest events in that cluster is less than 0.9. After identifying repeating clusters in both regions, our next step is to apply a sliding window waveform cross-correlation technique (Niu et al., 2003; Peng and Ben-Zion, 2006) to calculate the delay time and decorrelation index for each repeating cluster. By measuring temporal changes in waveforms of repeating clusters at different locations and depth, we hope to obtain a better constraint on spatio-temporal variations of fault zone properties and near-surface layers associated with the occurrence of major earthquakes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70094691','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70094691"><span>Volatile fluxes through the Big Bend section of the San <span class="hlt">Andreas</span> Fault, California: helium and carbon-dioxide systematics</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kulongoski, Justin T.; Hilton, David R.; Barry, Peter H.; Esser, Bradley K.; Hillegonds, Darren; Belitz, Kenneth</p> <p>2013-01-01</p> <p>To investigate the source of volatiles and their relationship to the San <span class="hlt">Andreas</span> Fault System (SAFS), 18 groundwater samples were collected from wells near the Big Bend section of the SAFS in southern California and analyzed for helium and carbon abundance and isotopes. Concentrations of 4He, corrected for air-bubble entrainment, vary from 4.15 to 62.7 (× 10− 8) cm3 STP g− 1 H2O. 3He/4He ratios vary from 0.09 to 3.52 RA (where RA = air 3He/4He), consistent with up to 44% mantle helium in samples. A subset of 10 samples was analyzed for the major volatile phase (CO2) — the hypothesized carrier phase of the helium in the mantle–crust system: CO2/3He ratios vary from 0.614 to 142 (× 1011), and δ13C (CO2) values vary from − 21.5 to − 11.9‰ (vs. PDB). 3He/4He ratios and CO2 concentrations are highest in the wells located in the Mil Potrero and Cuddy valleys adjacent to the SAFS. The elevated 3He/4He ratios are interpreted to be a consequence of a mantle volatile flux though the SAFS diluted by radiogenic He produced in the crust. Samples with the highest 3He/4He ratios also had the lowest CO2/3He ratios. The combined helium isotope, He–CO2 elemental relationships, and δ13C (CO2) values of the groundwater volatiles reveal a mixture of mantle and deep crustal (metamorphic) fluid origins. The flux of fluids into the seismogenic zone at high hydrostatic pressure may cause fault rupture, and transfer volatiles into the shallow crust. We calculate an upward fluid flow rate of 147 mm a− 1 along the SAFS, up to 37 times higher than previous estimates (Kennedy et al., 1997). However, using newly identified characteristics of the SAFS, we calculate a total flux of 3He along the SAFS of 7.4 × 103 cm3 STP a− 1 (0.33 mol 3He a− 1), and a CO2 flux of 1.5 × 1013 cm3STP a− 1 (6.6 × 108 mol a− 1), ~ 1% of previous estimates. Lower fluxes along the Big Bend section of the SAFS suggest that the flux of mantle volatiles alone is insufficient to cause the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.S41B2441H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.S41B2441H"><span>Applying time-reverse-imaging techniques to locate individual low-frequency earthquakes on the San <span class="hlt">Andreas</span> fault near Cholame, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Horstmann, T.; Harrington, R. M.; Cochran, E.; Shelly, D. R.</p> <p>2013-12-01</p> <p>Observations of non-volcanic tremor have become ubiquitous in recent years. In spite of the abundance of observations, locating tremor remains a difficult task because of the lack of distinctive phase arrivals. Here we use time-reverse-imaging techniques that do not require identifying phase arrivals to locate individual low-frequency-earthquakes (LFEs) within tremor episodes on the San <span class="hlt">Andreas</span> fault near Cholame, California. Time windows of 1.5-second duration containing LFEs are selected from continuously recorded waveforms of the local seismic network filtered between 1-5 Hz. We propagate the time-reversed seismic signal back through the subsurface using a staggered-grid finite-difference code. Assuming all rebroadcasted waveforms result from similar wave fields at the source origin, we search for wave field coherence in time and space to obtain the source location and origin time where the constructive interference is a maximum. We use an interpolated velocity model with a grid spacing of 100 m and a 5 ms time step to calculate the relative curl field energy amplitudes for each rebroadcasted seismogram every 50 ms for each grid point in the model. Finally, we perform a grid search for coherency in the curl field using a sliding time window, and taking the absolute value of the correlation coefficient to account for differences in radiation pattern. The highest median cross-correlation coefficient value over at a given grid point indicates the source location for the rebroadcasted event. Horizontal location errors based on the spatial extent of the highest 10% cross-correlation coefficient are on the order of 4 km, and vertical errors on the order of 3 km. Furthermore, a test of the method using earthquake data shows that the method produces an identical hypocentral location (within errors) as that obtained by standard ray-tracing methods. We also compare the event locations to a LFE catalog that locates the LFEs from stacked waveforms of repeated LFEs</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.T51B2584R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.T51B2584R"><span>Salton Seismic Imaging Project Line 5—the San <span class="hlt">Andreas</span> Fault and Northern Coachella Valley Structure, Riverside County, California</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rymer, M. J.; Fuis, G.; Catchings, R. D.; Goldman, M.; Tarnowski, J. M.; Hole, J. A.; Stock, J. M.; Matti, J. C.</p> <p>2012-12-01</p> <p>The Salton Seismic Imaging Project (SSIP) is a large-scale, active- and passive-source seismic project designed to image the San <span class="hlt">Andreas</span> Fault (SAF) and the adjacent basins (Imperial and Coachella Valleys) in southern California. Here, we focus on SSIP Line 5, one of four 2-D NE-SW-oriented seismic profiles that were acquired across the Coachella Valley. The 38-km-long SSIP-Line-5 seismic profile extends from the Santa Rosa Ranges to the Little San Bernardino Mountains and crosses both strands of the SAF, the Mission Creek (MCF) and Banning (BF) strands, near Palm Desert. Data for Line 5 were generated from nine buried explosive sources (most spaced about 2 to 8 km apart) and were recorded on approximately 281 Texan seismographs (average spacing 138 m). First-arrival refractions were used to develop a refraction tomographic velocity image of the upper crust along the seismic profile. The seismic data were also stacked and migrated to develop low-fold reflection images of the crust. From the surface to about 8 km depth, P-wave velocities range from about 2 km/s to more than 7.5 km/s, with the lowest velocities within a well-defined (~2-km-deep, 15-km-wide) basin (< 4 km/s), and the highest velocities below the transition from the Coachella Valley to the Santa Rosa Ranges on the southwest and within the Little San Bernardino Mountains on the northeast. The MCF and BF strands of the SAF bound an approximately 2.5-km-wide horst-type structure on the northeastern side of the Coachella Valley, beneath which the upper crust is characterized by a pronounced low-velocity zone that extends to the bottom of the velocity image. Rocks within the low-velocity zone have significantly lower velocities than those to the northeast and the southwest at the same depths. Conversely, the velocities of rocks on both sides of the Coachella Valley are greater than 7 km/s at depths exceeding about 4 km. The relatively narrow zone of shallow high-velocity rocks between the surface traces of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2012-04-20/pdf/2012-9499.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2012-04-20/pdf/2012-9499.pdf"><span>77 FR 23791 - Oregon Disaster #OR-00042</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2012-04-20</p> <p>... SMALL BUSINESS ADMINISTRATION [Disaster Declaration 13060 and 13061] Oregon Disaster OR-00042... Administrative declaration of a disaster for the State of OREGON dated 04/02/2012. Incident: Severe Winter Storm... the disaster: Primary Counties: <span class="hlt">Marion</span>. Contiguous Counties: Oregon: Clackamas, Jefferson, Linn, Polk...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://rosap.ntl.bts.gov/view/dot/22722','DOTNTL'); return false;" href="https://rosap.ntl.bts.gov/view/dot/22722"><span>Research notes : polymer modified emulsions for chip seals.</span></a></p> <p><a target="_blank" href="http://ntlsearch.bts.gov/tris/index.do">DOT National Transportation Integrated Search</a></p> <p></p> <p>1991-12-01</p> <p>The Research Unit is conducting a study of chip seal emulsions using asphalts containing polymers on test sections that were built in 1987 on Oregon Route 22 near Stayton in <span class="hlt">Marion</span> County. A commonly used emulsion in the 1987 OSHD Specifications for ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=327770&Lab=NRMRL&keyword=management+AND+projects&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=327770&Lab=NRMRL&keyword=management+AND+projects&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>NATIONAL STORMWATER CALCULATOR: LOW IMPACT DEVELOPMENT STORMWATER CONTROL COST ESTIMATION PROGRAMMING & FUTURE ENHANCEMENTS</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>National Stormwater Calculator: Low Impact Development Stormwater Control Cost Estimation Programming & Future EnhancementsJason Berner1; Michael Tryby1; Scott Struck2, Dan Pankani2, <span class="hlt">Marion</span> Deerhake3, Michelle Simon11. USEPA2. GeoSyntec, Inc.3. RTI, Inc.The National Stormwater Ca...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060453p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060453p/"><span>14. OVERALL VIEW OF ELEVATOR LOBBY OF BUILDING 1, SHOWING ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>14. OVERALL VIEW OF ELEVATOR LOBBY OF BUILDING 1, SHOWING THE CONTINUATION OF THE EGYPTIAN MOTIF DECORATIVE DETAIL; LOOKING NNW TOWARD THE MAIN LOBBY. (Ryan & Harms) - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2012-09-07/pdf/2012-22101.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2012-09-07/pdf/2012-22101.pdf"><span>77 FR 55220 - Mississippi; Major Disaster and Related Determinations</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2012-09-07</p> <p>... removal and emergency protective measures (Categories A and B), including direct Federal assistance, under..., Lawrence, Lincoln, Madison, <span class="hlt">Marion</span>, Montgomery, Pearl River, Perry, Pike, Rankin, Stone, Walthall, Wayne, Wilkinson, and Yazoo Counties for debris removal and emergency protective measures (Categories A and B...</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-03-25/pdf/2011-7063.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-03-25/pdf/2011-7063.pdf"><span>76 FR 16846 - Ohio Disaster # OH-00026</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-03-25</p> <p>... SMALL BUSINESS ADMINISTRATION [Disaster Declaration 12492 and 12493] Ohio Disaster OH-00026 AGENCY... declaration of a disaster for the State of OHIO dated 03/18/2011. Incident: Severe Storms and Flooding... disaster: Primary Counties: Auglaize, <span class="hlt">Marion</span>. Contiguous Counties: Ohio: Allen, Crawford, Darke, Delaware...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70030759','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70030759"><span>Paleoearthquakes on the southern San <span class="hlt">Andreas</span> Fault, Wrightwood, California, 3000 to 1500 B.C.: A new method for evaluating paleoseismic evidence and earthquake horizons</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Scharer, K.M.; Weldon, R.J.; Fumal, T.E.; Biasi, G.P.</p> <p>2007-01-01</p> <p>We present evidence of 11-14 earthquakes that occurred between 3000 and 1500 B.C. on the San <span class="hlt">Andreas</span> fault at the Wrightwood paleoseismic site. Earthquake evidence is presented in a novel form in which we rank (high, moderate, poor, or low) the quality of all evidence of ground deformation, which are called "event indicators." Event indicator quality reflects our confidence that the morphologic and sedimentologic evidence can be attributable to a ground-deforming earthquake and that the earthquake horizon is accurately identified by the morphology of the feature. In four vertical meters of section exposed in ten trenches, we document 316 event indicators attributable to 32 separate stratigraphic horizons. Each stratigraphic horizon is evaluated based on the sum of rank (Rs), maximum rank (Rm), average rank (Ra), number of observations (Obs), and sum of higher-quality event indicators (Rs>1). Of the 32 stratigraphic horizons, 14 contain 83% of the event indicators and are qualified based on the number and quality of event indicators; the remaining 18 do not have satisfactory evidence for further consideration. Eleven of the 14 stratigraphic horizons have sufficient number and quality of event indicators to be qualified as "probable" to "very likely" earthquakes; the remaining three stratigraphic horizons are associated with somewhat ambiguous features and are qualified as "possible" earthquakes. Although no single measurement defines an obvious threshold for designation as an earthquake horizon, Rs, Rm, and Rs>1 correlate best with the interpreted earthquake quality. Earthquake age distributions are determined from radio-carbon ages of peat samples using a Bayesian approach to layer dating. The average recurrence interval for the 10 consecutive and highest-quality earthquakes is 111 (93-131) years and individual intervals are ??50% of the average. With comparison with the previously published 14-15 earthquake record between A.D. 500 and present, we find no evidence</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70048205','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70048205"><span>S-wave triggering of tremor beneath the Parkfield, California, section of the San <span class="hlt">Andreas</span> fault by the 2011 Tohoku, Japan earthquake: observations and theory</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hill, David P.; Peng, Zhigang; Shelly, David R.; Aiken, Chastity</p> <p>2013-01-01</p> <p>The dynamic stresses that are associated with the energetic seismic waves generated by the Mw 9.0 Tohoku earthquake off the northeast coast of Japan triggered bursts of tectonic tremor beneath the Parkfield section of the San <span class="hlt">Andreas</span> fault (SAF) at an epicentral distance of ∼8200  km. The onset of tremor begins midway through the ∼100‐s‐period S‐wave arrival, with a minor burst coinciding with the SHSH arrival, as recorded on the nearby broadband seismic station PKD. A more pronounced burst coincides with the Love arrival, followed by a series of impulsive tremor bursts apparently modulated by the 20‐ to 30‐s‐period Rayleigh wave. The triggered tremor was located at depths between 20 and 30 km beneath the surface trace of the fault, with the burst coincident with the S wave centered beneath the fault 30 km northwest of Parkfield. Most of the subsequent activity, including the tremor coincident with the SHSH arrival, was concentrated beneath a stretch of the fault extending from 10 to 40 km southeast of Parkfield. The seismic waves from the Tohoku epicenter form a horizontal incidence angle of ∼14°, with respect to the local strike of the SAF. Computed peak dynamic Coulomb stresses on the fault at tremor depths are in the 0.7–10 kPa range. The apparent modulation of tremor bursts by the small, strike‐parallel Rayleigh‐wave stresses (∼0.7  kPa) is likely enabled by pore pressure variations driven by the Rayleigh‐wave dilatational stress. These results are consistent with the strike‐parallel dynamic stresses (δτs) associated with the S, SHSH, and surface‐wave phases triggering small increments of dextral slip on the fault with a low friction (μ∼0.2). The vertical dynamic stresses δτd do not trigger tremor with vertical or oblique slip under this simple Coulomb failure model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G43A0900W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G43A0900W"><span>New constraints on slip rates and locking depths of the San <span class="hlt">Andreas</span> Fault System from Sentinel-1A InSAR and GAGE GPS observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ward, L. A.; Smith-Konter, B. R.; Higa, J. T.; Xu, X.; Tong, X.; Sandwell, D. T.</p> <p>2017-12-01</p> <p>After over a decade of operation, the EarthScope (GAGE) Facility has now accumulated a wealth of GPS and InSAR data, that when successfully integrated, make it possible to image the entire San <span class="hlt">Andreas</span> Fault System (SAFS) with unprecedented spatial coverage and resolution. Resulting surface velocity and deformation time series products provide critical boundary conditions needed for improving our understanding of how faults are loaded across a broad range of temporal and spatial scales. Moreover, our understanding of how earthquake cycle deformation is influenced by fault zone strength and crust/mantle rheology is still developing. To further study these processes, we construct a new 4D earthquake cycle model of the SAFS representing the time-dependent 3D velocity field associated with interseismic strain accumulation, co-seismic slip, and postseismic viscoelastic relaxation. This high-resolution California statewide model, spanning the Cerro Prieto fault to the south to the Maacama fault to the north, is constructed on a 500 m spaced grid and comprises variable slip and locking depths along 42 major fault segments. Secular deep slip is prescribed from the base of the locked zone to the base of the elastic plate while episodic shallow slip is prescribed from the historical earthquake record and geologic recurrence intervals. Locking depths and slip rates for all 42 fault segments are constrained by the newest GAGE Facility geodetic observations; 3169 horizontal GPS velocity measurements, combined with over 53,000 line-of-sight (LOS) InSAR velocity observations from Sentinel-1A, are used in a weighted least-squares inversion. To assess slip rate and locking depth sensitivity of a heterogeneous rheology model, we also implement variations in crustal rigidity throughout the plate boundary, assuming a coarse representation of shear modulus variability ranging from 20-40 GPa throughout the (low rigidity) Salton Trough and Basin and Range and the (high rigidity) Central</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70028715','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70028715"><span>Introduction to the special issue on the 2004 Parkfield earthquake and the Parkfield earthquake prediction experiment</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Harris, R.A.; Arrowsmith, J.R.</p> <p>2006-01-01</p> <p>The 28 September 2004 M 6.0 Parkfield earthquake, a long-anticipated event on the San <span class="hlt">Andreas</span> fault, is the world's best recorded earthquake to date, with state-of-the-art data obtained from geologic, geodetic, seismic, magnetic, and electrical field networks. This has allowed the preearthquake and postearthquake states of the San <span class="hlt">Andreas</span> fault in this region to be analyzed in detail. Analyses of these data provide views into the San <span class="hlt">Andreas</span> fault that show a complex geologic history, fault geometry, rheology, and response of the nearby region to the earthquake-induced ground movement. Although aspects of San <span class="hlt">Andreas</span> fault zone behavior in the Parkfield region can be modeled simply over geological time frames, the Parkfield Earthquake Prediction Experiment and the 2004 Parkfield earthquake indicate that predicting the fine details of future earthquakes is still a challenge. Instead of a deterministic approach, forecasting future damaging behavior, such as that caused by strong ground motions, will likely continue to require probabilistic methods. However, the Parkfield Earthquake Prediction Experiment and the 2004 Parkfield earthquake have provided ample data to understand most of what did occur in 2004, culminating in significant scientific advances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70022027','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70022027"><span>Slip rates on San Francisco Bay area faults from anelastic deformation of the continental lithosphere</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Geist, E.L.; Andrews, D.J.</p> <p>2000-01-01</p> <p>Long-term slip rates on major faults in the San Francisco Bay area are predicted by modeling the anelastic deformation of the continental lithosphere in response to regional relative plate motion. The model developed by Bird and Kong [1994] is used to simulate lithospheric deformation according to a Coulomb frictional rheology of the upper crust and a dislocation creep rheology at depth. The focus of this study is the long-term motion of faults in a region extending from the creeping section of the San <span class="hlt">Andreas</span> fault to the south up to the latitude of Cape Mendocino to the north. Boundary conditions are specified by the relative motion between the Pacific plate and the Sierra Nevada - Great Valley microplate [Argus and Gordon, 2000]. Rheologic-frictional parameters are specified as independent variables, and prediction errors are calculated with respect to geologic estimates of slip rates and maximum compressive stress directions. The model that best explains the region-wide observations is one in which the coefficient of friction on all of the major faults is less than 0.15, with the coefficient of friction for the San <span class="hlt">Andreas</span> fault being approximately 0.09, consistent with previous inferences of San <span class="hlt">Andreas</span> fault friction. Prediction error increases with lower fault friction on the San <span class="hlt">Andreas</span>, indicating a lower bound of ??SAF > 0.08. Discrepancies with respect to previous slip rate estimates include a higher than expected slip rate along the peninsula segment of the San <span class="hlt">Andreas</span> fault and a slightly lower than expected slip rate along the San Gregorio fault.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=278289','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=278289"><span>Water, weed, and nutrient management practices in organic blackberries</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>The purpose of our study is to investigate the effects of organic management on plant and soil water and nutrient relations, plant growth, yield, and fruit quality in an organic trailing blackberry production system. Treatments include: cultivar ('<span class="hlt">Marion</span>' and 'Black Diamond'); irrigation (post-harve...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA277808','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA277808"><span>FAA Aviation Forecasts, Fiscal Years 1994-2005</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1994-03-01</p> <p>ID (BOI) Fort Lauderdale Executive, FL (FXE) Idaho Falls Fanning Field, ID (IDA) Fort Myers Page Field, FL (FMY) Lewiston , ID (LWS) Fort Myers...Pacoima/Whitman, California (WHP) 3. Lakeland, Florida (LAL) 4. Valdosta Municipal, Georgia (VLD) 5. Halley, Idaho (SUN) 6. <span class="hlt">Marion</span> Williamson County</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/in0466.photos.581272p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/in0466.photos.581272p/"><span>SOUTHWESTERN CORNER OF SECTION 1, WITH (L TO R) BIVOUAC ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>SOUTHWESTERN CORNER OF SECTION 1, WITH (L TO R) BIVOUAC OF THE DEAD TABLET, INVERTED CANNON AND UNION WOMAN’S RELIEF CORPS #44 MONUMENT IN FOREGROUND. VIEW TO NORTHEAST. - Crown Hill Cemetery, Crown Hill National Cemetery, 700 West Thirty-eighth Street, Indianapolis, <span class="hlt">Marion</span> County, IN</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060443p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/il0693.photos.060443p/"><span>4. DETAIL VIEW OF NW CORNER OF BUILDING 1'S W ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. DETAIL VIEW OF NW CORNER OF BUILDING 1'S W WING, SHOWING THE PILASTERS, TERRA COTTA PANELS, AND THE EGYPTIAN MOTIF DECORATIVE CORNICE ELEMENTS; LOOKING ENE. (Harms) - Veterans Administration Medical Center, Building No. 1, Old State Route 13 West, <span class="hlt">Marion</span>, Williamson County, IL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-11-14/pdf/2011-29177.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-11-14/pdf/2011-29177.pdf"><span>76 FR 70361 - Approval and Promulgation of Air Quality Implementation Plans; Indiana; Redesignation of the...</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-11-14</p> <p>... ENVIRONMENTAL PROTECTION AGENCY 40 CFR Parts 52 and 81 [EPA-R05-OAR-2009-0839; FRL-9489-6... Area to Attainment of the 1997 Annual Standard for Fine Particulate Matter AGENCY: Environmental... request to redesignate the Indianapolis, Indiana nonattainment area (Hamilton, Hendricks, Johnson, <span class="hlt">Marion</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mo1259.photos.196968p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mo1259.photos.196968p/"><span>37. Photographic copy of the original construction drawing, 1934, by ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>37. Photographic copy of the original construction drawing, 1934, by Sverdrup and Parcel, Consulting Engineers, from microfilm copy at Bridge Division, Missouri Highway and Transportation Department. Stress sheet, continuous span - Mark Twain Memorial Bridge, Spanning Mississippi River at US Route 36, Hannibal, <span class="hlt">Marion</span> County, MO</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol17/pdf/CFR-2010-title40-vol17-sec81-315.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title40-vol17/pdf/CFR-2010-title40-vol17-sec81-315.pdf"><span>40 CFR 81.315 - Indiana.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2010&page.go=Go">Code of Federal Regulations, 2010 CFR</a></p> <p></p> <p>2010-07-01</p> <p>... Unclassifiable/Attainment Cass County Unclassifiable/Attainment Clark County Unclassifiable/Attainment Clay...: Vanderburgh County Attainment Indianapolis Area: <span class="hlt">Marion</span> County Attainment Louisville Area: Clark County 10/23... LaPorte CO., IN: LaPorte County 7/19/07 Attainment. Louisville, KY-IN: Clark County. Floyd County July 19...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-315.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-315.pdf"><span>40 CFR 81.315 - Indiana.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... Unclassifiable/Attainment Cass County Unclassifiable/Attainment Clark County Unclassifiable/Attainment Clay...: Vanderburgh County Attainment Indianapolis Area: <span class="hlt">Marion</span> County Attainment Louisville Area: Clark County 10/23... LaPorte CO., IN: LaPorte County 7/19/07 Attainment. Louisville, KY-IN: Clark County. Floyd County July 19...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2011-08-30/pdf/2011-22184.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2011-08-30/pdf/2011-22184.pdf"><span>76 FR 53927 - Kansas; Major Disaster and Related Determinations</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collection.action?collectionCode=FR">Federal Register 2010, 2011, 2012, 2013, 2014</a></p> <p></p> <p>2011-08-30</p> <p>... determinations. DATES: Effective Date: July 29, 2011. FOR FURTHER INFORMATION CONTACT: Peggy Miller, Office of... Public Assistance and Hazard Mitigation will be limited to 75 percent of the total eligible costs...: Barton, Clay, Cloud, Hamilton, Jewell, Lincoln, Logan, Lyon, <span class="hlt">Marion</span>, Mitchell, Morton, Osage, Osborne...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/al0682.photos.320922p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/al0682.photos.320922p/"><span>47. VIEW ON THE ROOF, LOOKING EAST TO WEST FROM ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>47. VIEW ON THE ROOF, LOOKING EAST TO WEST FROM THE SOUTHWEST CORNER OF THE MAIN ROOF (NOTE CHIMNEY TO NORTH & PEAK OF METAL ROOF COVERING THE EAST SIDE BAY OF OCTAGONAL ROOMS) - Kenworthy Hall, State Highway 14 (Greensboro Road), <span class="hlt">Marion</span>, Perry County, AL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eosweb.larc.nasa.gov/project/misr/gallery/mystery_9_answer','SCIGOV-ASDC'); return false;" href="https://eosweb.larc.nasa.gov/project/misr/gallery/mystery_9_answer"><span>Mystery #9 Answer</span></a></p> <p><a target="_blank" href="http://eosweb.larc.nasa.gov/">Atmospheric Science Data Center </a></p> <p></p> <p>2013-04-22</p> <p>... automated temperature and pressure data.   C.   A nuclear bomb test occurred within 3000 kilometers of the island.   D. ... weather station was established in 1977 •   A nuclear bomb was detonated between Bouvetøya and <span class="hlt">Marion</span> Island in 1979 ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=252514&keyword=ted&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=252514&keyword=ted&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>An Integrated Assessment of Sediment Remediation in a Midwestern U.S. Stream Using Sediment Chemistry, Water Quality, Bioassessment and Fish Biomarkers</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>A comprehensive biological, sediment and water quality study of the lower Little Scioto River near <span class="hlt">Marion</span>, Ohio, USA was undertaken in July 2007 to evaluate the effectiveness of removal of creosote-contaminated sediment. The study area covered 7.5 river miles (RMs) of the river, ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=anthropology+AND+poland&id=EJ716110','ERIC'); return false;" href="https://eric.ed.gov/?q=anthropology+AND+poland&id=EJ716110"><span>Chronicle of Higher Education. Volume 51, Number 23, February 11, 2005</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Chronicle of Higher Education, 2005</p> <p>2005-01-01</p> <p>"Chronicle of Higher Education" presents an abundant source of news and information for college and university faculty members and administrators. This February 11, 2005 issue of "Chronicle of Higher Education" includes the following articles: (1) "A Giant Eye on the Stars" (Lloyd, <span class="hlt">Marion</span>); (2) "Taiwanese…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-310.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title40-vol17/pdf/CFR-2011-title40-vol17-sec81-310.pdf"><span>40 CFR 81.310 - Florida.</span></a></p> <p><a target="_blank" href="http://www.gpo.gov/fdsys/browse/collectionCfr.action?selectedYearFrom=2011&page.go=Go">Code of Federal Regulations, 2011 CFR</a></p> <p></p> <p>2011-07-01</p> <p>... County Indian River County Jackson County Jefferson County Lafayette County Lake County Lee County Leon... County Lee County Leon County Levy County Liberty County Madison County Manatee County <span class="hlt">Marion</span> County... Indian River County Jackson County Jefferson County Lafayette County Lake County Lee County Leon County...</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>