Sample records for east glacier park

  1. Alaska: Glaciers of Kenai Fjords National Park and Katmai National Park and Preserve

    NASA Technical Reports Server (NTRS)

    Giffens, Bruce A.; Hall, Dorothy K.; Chien, Janet Y. L.

    2014-01-01

    There are hundreds of glaciers in Kenai Fjords National Park (KEFJ) and Katmai National Park and Preserve (KATM) covering over 2,276 sq km of park land (ca. 2000). There are two primary glacierized areas in KEFJ (the Harding Icefield and the Grewingk-Yalik Glacier Complex) and three primary glacierized areas in KATM (the Mt. Douglas area, the Kukak Volcano to Mt. Katmai area, and the Mt. Martin area). Most glaciers in these parks terminate on land, though a few terminate in lakes. Only KEFJ has tidewater glaciers, which terminate in the ocean. Glacier mapping and analysis of the change in glacier extent has been accomplished on a decadal scale using satellite imagery, primarily Landsat data from the 1970s, 1980s, and from2000. Landsat Multispectral Scanner (MSS),Thematic Mapper (TM), and Enhanced Thematic Mapper Plus (ETM) imagery was used to map glacier extent on a park-wide basis. Classification of glacier ice using image-processing software, along with extensive manual editing, was employed to create Geographic Information System (GIS)outlines of the glacier extent for each park. Many glaciers that originate in KEFJ but terminate outside the park boundaries were also mapped. Results of the analysis show that there has been a reduction in the amount of glacier ice cover in the two parks over the study period. Our measurements show a reduction of approximately 21 sq km, or 1.5(from 1986 to 2000), and 76 sq km, or 7.7 (from19861987 to 2000), in KEFJ and KATM, respectively. This work represents the first comprehensive study of glaciers of KATM. Issues that complicate the mapping of glacier extent include debris cover(moraine and volcanic ash), shadows, clouds, fresh snow, lingering snow from the previous season, and differences in spatial resolution between the MSS,TM, or ETM sensors. Similar glacier mapping efforts in western Canada estimate mapping errors of 34. Measurements were also collected from a suite of glaciers in KEFJ and KATM detailing terminus positions

  2. Alaska: Glaciers of Kenai Fjords National Park and Katmai and Lake Clark National Parks and Preserve

    NASA Technical Reports Server (NTRS)

    Giffen, bruce A.; Hall, Dorothy K.; Chien, Janet Y. L.

    2011-01-01

    There are hundreds of glaciers in Kenai Fjords National Park (KEFJ) and Katmai National Park and Preserve (KATM) covering over 2276 sq km of park land (circa 2000). There are two primary glacierized areas in KEFJ -- the Harding Icefield and the Grewingk-Yalik Glacier Complex, and three primary glacierized areas in KATM - the Mt. Douglas area, the Kukak Volcano to Mt. Katmai area and the Mt. Martin area. Most glaciers in these parks terminate on land, though a few terminate in lakes. Only KEFJ has tidewater glaciers, which terminate in the ocean. Glacier mapping and analysis of the change in glacier extent has been accomplished on a decadal scale using satellite imagery, primarily Landsat data from the 1970s, 1980s, and from 2000. Landsat Multispectral Scanner (MSS), Thematic Mapper (TM) and Enhanced Thematic Mapper Plus (ETM+) imagery was used to map glacier extent on a park-wide basis. Classification of glacier ice using image processing software, along with extensive manual editing, was employed to create Geographic Information System (GIS) outlines of the glacier extent for each park. Many glaciers that originate in KEFJ but terminate outside the park boundaries were also mapped. Results of the analysis show that there has been a reduction in the amount of glacier ice cover in the two parks over the study period. Our measurements show a reduction of approximately 21 sq km, or -1.5% (from 1986 to 2000), and 76 sq km, or -7.7% (from 1986/87 to 2000), in KEFJ and KATM, respectively. This work represents the first comprehensive study of glaciers of KATM. Issues that complicate the mapping of glacier extent include: debris-cover (moraine and volcanic ash), shadows, clouds, fresh snow, lingering snow from the previous season, and differences in spatial resolution between the MSS and TM or ETM+ sensors. Similar glacier mapping efforts in western Canada estimate mapping errors of 3-4%. Measurements were also collected from a suite of glaciers in KEFJ and KATM detailing

  3. USA: Glacier National Park, Biosphere Reserve and GLORIA Site

    USGS Publications Warehouse

    Fagre, Daniel B.; Lee, Cathy; Schaaf, Thomas; Simmonds, Paul

    2004-01-01

    The area now managed as Glacier National Park was first set aside as a Forest Reserve in 1897 and then designated as a national park in 1910, six years before a national park service was created to oversee the growing number of parks that the US Congress was establishing. Waterton National Park was created by Canada immediately north of the US–Canada border during the same period. In 1932, a joint lobbying effort by private citizens and groups convinced both the United States and Canada to establish the world’s first trans-boundary park to explicitly underscore and symbolize the neighbourly relationship between these two countries. This became the world’s first ‘peace’ park and was named Waterton–Glacier International Peace Park. The combined park is managed collaboratively on many issues but each national park is separately funded and operates under different national statutes and laws. It was, however, jointly named a Biosphere Reserve in 1976 and a World Heritage Site in 1995. There have been recent efforts to significantly increase the size of Waterton National Park by adding publicly owned forests on the western side of the continental divide in British Columbia, Canada. For the purposes of this chapter, I will emphasize the US portion of the Waterton-Glacier International Peace Park and refer to it as the Glacier Mountain Biosphere Reserve (MBR).

  4. Alaska: Glaciers of Kenai Fjords National Park and Katmai National Park and Preserve (Chapter 12)

    NASA Technical Reports Server (NTRS)

    Giffen, Bruce A.; Hall, Dorothy K.; Chien, Janet Y.L.

    2007-01-01

    Much recent research points to the shrinkage of the Earth's small glaciers, however, few studies have been performed to quantify the amount of change over time. We measured glacier-extent changes in two national parks in southeastern Alaska. There are hundreds of glaciers in Kenai Fjords National Park (KEFJ) and Katmai National Park and Preserve (KATM) covering over 2373 sq km of parkland. There are two primary areas of glaciation in KEFJ - the Harding Icefield and the Grewingk-Yalik Glacier Complex, and three primary areas of glaciation in KATM - the Mt. Douglas area, the Kukak Volcano to Mt. Katmai area and the Mt. Martin area. We performed glacier mapping using satellite imagery, from the 1970s, 1980s, and from 2000. Results of the analysis show that there has been a reduction in the amount of glacier ice cover in the two parks over the study period, of approximately 22 sq km of ice, approximately - 1.6% from 1986 to 2000 (for KEFJ), and of approximately 76 sq km of glacier ice, or about -7.7% from 1986187 to 2000 (for KATM). In the future, measurements of surface elevation changes of these ice masses should be acquired; together with our extent-change measurements, the volume change of the ice masses can then be determined to estimate their contribution to sea-level rise. The work is a continuation of work done in KEFJ, but in KATM, our measurements represent the first comprehensive study of the glaciers in this remote, little-studied area.

  5. 5. GLACIER POINT ROAD VIEW AT SENTINEL DOME PARKING AREA. ...

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

    5. GLACIER POINT ROAD VIEW AT SENTINEL DOME PARKING AREA. LOOKING E. GIS: N-37 42 43.8 / W-119 35 12.1 - Glacier Point Road, Between Chinquapin Flat & Glacier Point, Yosemite Village, Mariposa County, CA

  6. 1. PARKING LOT AT GLACIER POINT. HALF DOME AT CENTER ...

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

    1. PARKING LOT AT GLACIER POINT. HALF DOME AT CENTER REAR. LOOKING NE. GIS: N-36 43 45.8 / W-119 34 14.1 - Glacier Point Road, Between Chinquapin Flat & Glacier Point, Yosemite Village, Mariposa County, CA

  7. Modeled climate-induced glacier change in Glacier National Park, 1850-2100

    USGS Publications Warehouse

    Hall, M.H.P.; Fagre, D.B.

    2003-01-01

    The glaciers in the Blackfoot-Jackson Glacier Basin of Glacier National Park, Montana, decreased in area from 21.6 square kilometers (km2) in 1850 to 7.4 km2 in 1979. Over this same period global temperatures increased by 0.45??C (?? 0. 15??C). We analyzed the climatic causes and ecological consequences of glacier retreat by creating spatially explicit models of the creation and ablation of glaciers and of the response of vegetation to climate change. We determined the melt rate and spatial distribution of glaciers under two possible future climate scenarios, one based on carbon dioxide-induced global warming and the other on a linear temperature extrapolation. Under the former scenario, all glaciers in the basin will disappear by the year 2030, despite predicted increases in precipitation; under the latter, melting is slower. Using a second model, we analyzed vegetation responses to variations in soil moisture and increasing temperature in a complex alpine landscape and predicted where plant communities are likely to be located as conditions change.

  8. Topographic context of glaciers and perennial snowfields, Glacier National Park, Montana

    NASA Astrophysics Data System (ADS)

    Allen, Thomas R.

    1998-01-01

    Equilibrium-line altitudes (ELAs) of modem glaciers in the northern Rocky Mountains are known to correspond with regional climate, but strong subregional gradients such as across the Continental Divide in Glacier National Park, Montana, also exert topoclimatic influences on the ELA. This study analyzed the relationships between glacier and snowfield morphology, ELA, and surrounding topography. Glaciers and perennial snowfields were mapped using multitemporal satellite data from the Landsat Thematic Mapper and aerial photography within an integrated Geographic Information System (GIS). Relationships between glacier morphology and ELA were investigated using discriminant analysis. Four morphological categories of perennial snow and ice patches were identified: cirque glacier, niche glacier, ice cap, and snowfield. ELA was derived from overlaid glacier boundaries and Digital Elevation Models (DEMs) within the GIs. DEMs provided topographic variables and models of solar radiation and wind exposure/shelteredness. Regression analysis showed the effects of exposure; on snow accumulation, the strong influence of local topography through upslope zone morphology such as cirque backwalls, and the tendency for glaciers with high ELAs to exhibit compactness in morphology. Results highlight the relatively compact shape and larger area of glaciers adjacent to the Continental Divide. Discriminant analysis correctly predicted the type of glacier morphology in more than half the observations using factored variables of glacier shape, elevation range, and upslope area.

  9. Flood estimates for ungaged streams in Glacier and Yellowstone National Parks, Montana

    USGS Publications Warehouse

    Omang, R.J.; Parrett, Charles; Hull, J.A.

    1983-01-01

    Estimates of 100-year discharges were made at 59 sites in Glacier National Park and 21 sites in Yellowstone National Park to assist the National Park Services in quantifying stream inflow and outflow in the Parks. The estimates were made using regression equations previously developed for Montana. The resulting 100-year discharges are listed in tables; the discharges ranged from 260 to 53,200 cu ft/s in Glacier National Park and from 110 to 27,900 cu ft/s in Yellowstone National Park. (USGS)

  10. Adapting to climate change at Glacier National Park, Montana, USA (Invited)

    NASA Astrophysics Data System (ADS)

    Fagre, D. B.

    2009-12-01

    The impact of climate change on mountain watersheds has been studied at Glacier National Park, Montana since 1991. Despite a 14% increase in annual precipitation, glaciers have receded, snow packs have diminished, and late season stream discharge has declined. Snow melts one month earlier in the spring, leading to earlier hydrologic peaks and tree invasions of subalpine meadows. This has been largely driven by annual temperature increases that are 2-3 times greater than the global average for the past century. How do scientists and park managers adapt? Although stopping the glaciers from disappearing is not a management option, park staff have embarked on an aggressive education and interpretation effort to use melting glaciers as the segue into dialog about climate change. Media such as podcasts, handouts, posters, visitor center displays and roadside signage complement interpretive ranger-led talks about climate change and incorporate the latest glacial data from ongoing research. With few historic data on most animal populations, Glacier Park staff and other scientists are unable to assess the impacts of climate change to resources that the public cares about. They have recently initiated alpine wildlife monitoring programs to track populations of potentially climate-sensitive organisms such as the American pika (Ochotona princeps). Recognizing that climate change increases the frequency and severity of extreme weather events, design specifications for reconstruction of an alpine highway were adjusted to include larger culverts and hardened rock walls. Species that are dependent on cold water will be at risk as glaciers and snowfields disappear but managers cannot control these processes. However, they are proactively reducing other stressors to sensitive native fish species by removing exotic, introduced species that are competitors. In addition to these adaptation measures, Glacier Park has implemented shuttles, fleet conversions and enhanced building

  11. A climatic handbook for Glacier National Park-with data for Waterton Lakes National Park

    Treesearch

    Arnold I. Finklin

    1986-01-01

    A climatic description of the Glacier-Waterton Lakes Park area; mainly covers Glacier. Contains numerous tables, graphs, and maps showing the year-round pattern of climatic elements and 10-day details during fire season. Data analysis includes frequency distributions in addition to average values. Examines relationship of averages to topography, weather correlations...

  12. 36 CFR 13.1130 - Is commercial fishing authorized in the marine waters of Glacier Bay National Park?

    Code of Federal Regulations, 2013 CFR

    2013-07-01

    ... authorized in the marine waters of Glacier Bay National Park? 13.1130 Section 13.1130 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1130 Is commercial...

  13. 36 CFR 13.1130 - Is commercial fishing authorized in the marine waters of Glacier Bay National Park?

    Code of Federal Regulations, 2014 CFR

    2014-07-01

    ... authorized in the marine waters of Glacier Bay National Park? 13.1130 Section 13.1130 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1130 Is commercial...

  14. 36 CFR 13.1130 - Is commercial fishing authorized in the marine waters of Glacier Bay National Park?

    Code of Federal Regulations, 2011 CFR

    2011-07-01

    ... authorized in the marine waters of Glacier Bay National Park? 13.1130 Section 13.1130 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1130 Is commercial...

  15. 36 CFR 13.1130 - Is commercial fishing authorized in the marine waters of Glacier Bay National Park?

    Code of Federal Regulations, 2012 CFR

    2012-07-01

    ... authorized in the marine waters of Glacier Bay National Park? 13.1130 Section 13.1130 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1130 Is commercial...

  16. 36 CFR 13.1130 - Is commercial fishing authorized in the marine waters of Glacier Bay National Park?

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... authorized in the marine waters of Glacier Bay National Park? 13.1130 Section 13.1130 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1130 Is commercial...

  17. Examples of fire restoration in Glacier National Park

    Treesearch

    Laurie Kurth

    1996-01-01

    Covering just over 1 million acres, Glacier National Park straddles the Continental Divide in northwestern Montana. Diverse vegetation communities include moist western cedar- western hemlock (Thuja plicata - Tsuga heterophylla) old growth forests similar to those of the Pacific Coast, dry western grasslands and prairies, dense...

  18. Geology of Glacier National Park and the Flathead Region, Northwestern Montana

    USGS Publications Warehouse

    Ross, Clyde P.

    1959-01-01

    This report summarizes available data on two adjacent and partly overlapping regions in northwestern Montana. The first of these is Glacier National Park plus small areas east and west of the park. The second is here called, for convenience, the Flathead region; it embraces the mountains from the southern tip of Glacier Park to latitude 48 deg north and between the Great Plains on the east and Flathead Valley on the west. The fieldwork under the direction of the writer was done in 1948, 1949, 1950, and 1951, with some work in 1952 and 1953. The two regions together include parts of the Swan, Flathead, Livingstone, and Lewis Ranges. They are drained largely by branches of the Flathead River. On the east and north, however, they are penetrated by tributaries of the Missouri River and in addition by streams that flow into Canada. Roads and highways reach the borders of the regions; but there are few roads in the regions and only two highways cross them. The principal economic value of the assemblage of mountains described in the present report is as a collecting ground for snow to furnish the water used in the surrounding lowlands and as a scenic and wildlife recreation area. A few metallic deposits and lignitic coal beds are known, but these have not proved to be important and cannot, as far as can now be judged, be expected to become so. No oil except minor seeps has yet been found, and most parts of the two regions covered do not appear geologically favorable to the presence of oil in commercial quantities. The high, Hungry Horse Dam on which construction was in progress during the fieldwork now floods part of the Flathead region and will greatly influence the future of that region. The rocks range in age from Precambrian to Recent. The thickest units belong to the Belt series of Precambrian age, and special attention was paid to them. As a result, it is clear that at least the upper part of the series shows marked lateral changes within short distances. This fact

  19. Grinnell and Sperry Glaciers, Glacier National Park, Montana: A record of vanishing ice

    USGS Publications Warehouse

    Johnson, Arthur

    1980-01-01

    Grinnell and Sperry Glaciers, in Glacier National Park, Mont., have both shrunk considerably since their discovery in 1887 and 1895, respectively. This shrinkage, a reflection of climatic conditions, is evident when photographs taken at the time of discovery are compared with later photographs. Annual precipitation and terminus-recession measurements, together with detailed systematic topographic mapping since 1900, clearly record the changes in the character and size of these glaciers. Grinnell Glacier decreased in area from 530 acres in 1900 to 315 acres in 1960 and to 298 acres in 1966. Between 1937 and 1969 the terminus receded nearly 1,200 feet. Periodic profile measurements indicate that in 1969 the surface over the main part of the glacier was 25-30 feet lower than in 1950. Observations from 1947 to 1969 indicate annual northeastward movement ranging from 32 to 52 feet and generally averaging 35-45 feet. The annual runoff at the glacier is estimated to be 150 inches, of which approximately 6 inches represents reduction in glacier volume. The average annual runoff at a gaging station on Grinnell Creek 1.5 miles downvalley from the glacier for the 20-year period, 1949-69, was 100 inches. The average annual precipitation over the glacier was probably 120-150 inches. Sperry Glacier occupied 800 acres in 1901; by 1960 it covered only 287 acres, much of its upper part having disappeared from the enclosing cirque. From 1938 to 1969 certain segments of the terminus receded more than 1,000 feet. Profile measurements dating from 1949 indicate a lowering of the glacier surface below an altitude of 7,500 feet, but a fairly constant or slightly increased elevation of the surface above an altitude of 7,500 feet. Along one segment of the 1969 terminus the ice had been more than 100 feet thick in 1950. According to observations during 1949-69, average annual downslope movement was less than 15 feet per year in the central part of the glacier and slightly more rapid toward

  20. 2011 Updates on the Long-term Glacier Monitoring Program in Denali National Park and Preserve

    NASA Astrophysics Data System (ADS)

    Burrows, R. A.; Adema, G. W.; Herreid, S. J.; Arendt, A. A.; Larsen, C. F.

    2011-12-01

    The area of Denali National Park and Preserve (DENA) dominated by ice is vast, with glaciers covering 3,780 km^2, approximately one sixth of the park's area. They are integral components of the region's hydrologic, ecologic, and geologic systems - with changes to the glacier systems driving the dependent ecosystems. The National Park Service (NPS) conducts long term monitoring of glaciers in Denali with a variety of methods at a range of spatial and temporal scales. This includes seasonal mass balance and surface movement data collection, annual searches for surging glaciers, and decadal areal extent mapping and volume change estimates of all glaciers in the park. If a glacier surge is detected, the event is documented via photography and surface measurements, when possible. In addition, more intensive ground-based GPS surveys of termini and ice surface elevations are conducted on ten study glaciers every 5-10 years, on a rotating basis. Many of the glaciers are located in designated Wilderness, hence the use of mechanized transport is reduced as much as possible. Monitoring objectives are accomplished by park staff and with cooperative agreements with other agencies and universities. Research to understand the context of the long term data is encouraged and supported as much as possible by the NPS and has recently yielded significant results. The year 2011 marks the 20th anniversary of glacier mass balance monitoring on Kahiltna and Traleika Glaciers, located on the south and north sides of Mt. McKinley respectively. A single "index" site near the ELA of each glacier provides an index of winter, summer, and net balances each year as well as flow velocities and changes in surface elevation. Long-term net balance trends are positive from 1991-2003, and negative since 2003, including the 2009-2010 balance year. The average flow velocity at the Kahiltna index site is 200 +/- 21 m/year with a neutral to slightly negative trend, while on Traleika average velocity is 67

  1. 6. VIEW OF NORTH END OF EAST DAM, LOOKING SOUTH. ...

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

    6. VIEW OF NORTH END OF EAST DAM, LOOKING SOUTH. (View is taken from lakeside with lowered water level. This view encompasses the same area as MT-88-A-5 above.) - Three Bears Lake & Dams, East Dam, North of Marias Pass, East Glacier Park, Glacier County, MT

  2. Observations of paraglacial processes on glacier forelands in Aoraki/Mount Cook National Park, Southern Alps, New Zealand

    NASA Astrophysics Data System (ADS)

    Winkler, Stefan

    2015-04-01

    The large and extensively debris-covered valley glaciers in Aoraki/Mount Cook National Park immediate east of the Main Divide in the Southern Alps of New Zealand experienced a substantial frontal retreat and vertical downwasting during the past few decades, often connected with the development of a proglacial lake and retreat by calving. Their Holocene glacier forelands are characterised by huge lateral moraines and multi-ridged lateral moraine systems alongside smaller terminal moraines and frontal outwash heads. Placed within a very dynamic general geomorphological regime of various efficient process-systems, these Holocene glacier forelands are currently affected by substantial paraglacial modification. These paraglacial processes have already caused some consequences for the touristic infrastructure in the area and are likely to cause further problems for the accessibility of established tramping routes, tourist huts, and lookouts in the near and medium future. One of the first steps in a project under development presented here is a detailed visual comparison of changes documented during the past 15 Years on the glacier forelands of Hooker, Mueller and Tasman Glaciers in Aoraki/Mount Cook National Park. It reveals considerable erosion especially on the proximal slopes of the lateral moraines by gully development and retreat of erosion scars at their crest in order of several metres in just a few years. Different processes contribute to high erosion rates, among others rill erosion connected to mid-slope springs that only are temporarily active following substantial rainfall events, efficient gully incision, and slumping. Although any quantification of the actual erosion rates is just preliminary and further studies are necessary in order to make reliable predictions for future development, the amount of paraglacial erosion in this environment is very high compared to other regions and highlights the current importance of the paraglacial process-system in the

  3. The Rocks and Fossils of Glacier National Park: The Story of Their Origin and History

    USGS Publications Warehouse

    Ross, Clyde P.; Rezak, Richard

    1959-01-01

    The story of Glacier National Park begins about 500 million years ago, at a time when there were no mountains in the region - only a vast, exceedingly shallow sea, bordered by desolate plains. The sand, clay, and mud, in part very limy, that were laid down in this sea eventually hardened into the rocks that are now known as the Belt series. These are the principal rocks in the park. Scattered through these rocks are crinkled, limy masses of many forms, the remains of deposits made by colonies of algae. After the Belt series was laid down, successive seas slowly advanced and retreated through long ages across what is now Glacier National Park, burying the Belt rocks under younger ones. After another very long time, a gentle uplift, the forerunner of later events, brought this part of the continent above the reach of sea water for the last time. Much later, some 50 million years ago, the disturbance became far more intense. To climax this upheaval, a mass of rock thousands of feet thick and hundreds of miles long was shoved eastward for 35 miles or more. This tremendous dislocation, well exposed along the eastern boundary of the park, is known as the Lewis overthrust. When the rocks of the region emerged from the sea they began to be attacked by erosion. As successive periods of crustal movement and erosion continued, the younger rocks were slowly stripped off the Belt series and sculpture of the latter by weather and water shaped the early Rocky Mountains. The final episode in the park's geologic past was the ice age, beginning about a million years ago. Repeated advances and retreats of the great glaciers in the high valleys accentuated the mountain terrain and developed the scenic grandeur that is now Glacier National Park. One may say that the park is still in the ice age, for some glaciers still exist. The present report, companion to two more technical reports on the region, informally presents the story of the park's development through past eras for readers

  4. The mosquitoes and chaoborids of Glacier and Yellowstone National Parks with new records and Ochlerotatus nevadensis, a new state record for Montana.

    PubMed

    Nielsen, Lewis T

    2012-03-01

    The known mosquito fauna of Glacier National Park, Montana, and Yellowstone National Park, Wyoming, is reported with new records, including a list of the species of Chaoboridae known from both parks. Ochlerotatus nevadensis (= Aedes nevadensis) from Glacier National Park is a new record for the state of Montana.

  5. Post-Little Landscape and Glacier Change in Glacier Bay National Park: Documenting More than a Century of Variability with Repeat Photography

    NASA Astrophysics Data System (ADS)

    Molnia, B. F.; Karpilo, R. D.; Pranger, H. S.

    2004-12-01

    Historical photographs, many dating from the late-19th century are being used to document landscape and glacier change in the Glacier Bay area. More than 350 pre-1980 photographs that show the Glacier Bay landscape and glacier termini positions have been acquired by the authors. Beginning in 2003, approximately 150 of the sites from which historical photographs had been made were revisited. At each site, elevation and latitude and longitude were recorded using WAAS-enabled GPS. Compass bearings to photographic targets were also determined. Finally, using the historical photographs as a composition guide, new photographs were exposed using digital imaging and film cameras. In the laboratory, 21st century images and photographs were compared with corresponding historical photographs to determine, and to better understand rates, timing, and mechanics of Glacier Bay landscape evolution, as well as to clarify the response of specific glaciers to changing climate and environment. The comparisons clearly document rapid vegetative succession throughout the bay; continued retreat of larger glaciers in the East Arm of the bay; a complex pattern of readvance and retreat of the larger glaciers in the West Arm of the bay, coupled with short-term fluctuations of its smaller glaciers; transitions from tidewater termini to stagnant, debris-covered termini; fiord sedimentation and erosion; development of outwash and talus features; and many other dramatic changes. As might be expected, 100-year-plus photo comparisons show significant changes throughout the Glacier Bay landscape, especially at the southern ends of East and West Arms. Surprisingly, recent changes, occurring since the late-1970s were equally dramatic, especially documenting the rapid thinning and retreat of glaciers in upper Muir Inlet.

  6. Grizzly bear habitat research in Glacier National Park, Montana

    USGS Publications Warehouse

    Martinka, C.J.; Kendall, K.C.

    1986-01-01

    Grizzly bear habitat research began in 1967 and is continuing in Glacier National Park, MT. Direct observations and fecal analysis revealed a relatively definable pattern of habitat use by the bears. Habitat data were subsequently used to develop management models and explore the relationships between grizzlies and park visitors. Current research strategy is based on the concept that humans are an integral components of grizzly bear habitat. A geographic information system is being developed to assist in the application of habitat data. In addition, the behavioral response of grizzlies to annual changes in food production is being studied. Management that addresses bears, humans, and their habitat as a system is proposed.

  7. Ecological Succession in the Pleistocene in Glacier National Park, Montana, in Relation to Current Successional Stages in the Western Mountains of the U.S.

    ERIC Educational Resources Information Center

    Arnfield, Edwin A.

    1991-01-01

    Discusses the succession of ecological and geological structures as exhibited at Glacier National Park, Montana. Topics discussed include glaciers, the geological history of Glacier National Park, glaciation of the Rocky Mountains, paleoecology, the vegetational history of the Northwestern United States, and glaciation and the modern vegetation.…

  8. Appraisal of ground-water quality near wastewater-treatment facilities, Glacier National Park, Montana

    USGS Publications Warehouse

    Moreland, Joe A.; Wood, Wayne A.

    1982-01-01

    Water-level and water-quality data were collected from monitoring wells at wastewater-treatment facilities in Glacier National Park. Five additional shallow observation wells were installed at the Glacier Park Headquarters facility to monitor water quality in the shallow ground-water system.Water-level, water-quality, and geologic information indicate that some of the initial monitoring wells are not ideally located to sample ground water most likely to be affected by waste disposal at the sites. Small differences in chemical characteristics between samples from monitor wells indicate that effluent may be affecting ground-water quality but that impacts are not significant.Future monitoring of ground-water quality could be limited to selected wells most likely to be impacted by percolating effluent. Laboratory analyses for common ions could detect future impacts.

  9. Whitebark and limber pine restoration and monitoring in Glacier National Park

    Treesearch

    Jennifer M. Asebrook; Joyce Lapp; Tara. Carolin

    2011-01-01

    Whitebark pine (Pinus albicaulis) and limber pine (Pinus flexilis) are keystone species important to watersheds, grizzly and black bears, squirrels, birds, and other wildlife. Both high elevation five-needled pines have dramatically declined in Glacier National Park primarily due to white pine blister rust (Cronartium ribicola) and fire exclusion, with mountain pine...

  10. Fire history of southeastern Glacier National Park: Missouri River Drainage

    USGS Publications Warehouse

    Barrett, Stephen W.

    1993-01-01

    In 1982, Glacier National Park (GNP) initiated long-term studies to document the fire history of all forested lands in the 410,000 ha. park. To date, studies have been conducted for GNP west of the Continental Divide (Barrett et al. 1991), roughly half of the total park area. These and other fire history studies in the Northern Rockies (Arno 1976, Sneck 1977, Arno 1980, Romme 1982, Romme and Despain 1989, Barrett and Arno 1991, Barrett 1993a, Barrett 1993b) have shown that fire history data can be an integral element of fire management planning, particularly wen natiral fire plans are being developed for parks and wilderness. The value of site specific fire history data is apparent when considering study results for lodgepole pin (Pinus contorta var. latifolia) forests. Lodgepole pine is a major subalpine type in the Northern Rockies and such stands experiences a wide range of presettlement fire patterns. On relatively warm-dry sites at lower elevations, such as in GNP's North Fork drainage (Barrett et al. 1991), short to moderately long interval (25-150 yr) fires occurred in a mixed severity pattern ranging from non-lethal underburns to total stand replacement (Arno 1976, Sneck 1977, Barrett and Arno 1991). Markedly different fire history occurred at high elevation lodgepole pine stands on highly unproductive sites, such as on Yellowstone National Park's (YNP) subalpine plateau. Romme (1982) found that, on some sites, stand replacing fires recurred after very long intervals (300-400 yr), and that non-lethal surface fires were rare. For somewhat more productive sites in the Absaroka Mountains in YNP, Barrett (1993a) estimated a 200 year mean replacement interval, in a pattern similar to that found in steep mountain terrain elsewhere, such as in the Middle Fork Flathead River drainage (Barrett et al. 1991, Sneck 1977). Aside from post-1900 written records (ayres 1900; fire atlas data on file, GNP Archives Div. and GNP Resources Mgt. Div.), little fire history

  11. Retreat and stagnation of Little Ice Age glaciers in Yosemite National Park

    NASA Astrophysics Data System (ADS)

    Stock, G. M.; Anderson, R. S.; Painter, T. H.

    2016-12-01

    The high peaks of Yosemite National Park in the Sierra Nevada, California, retain several small (<1 km2) glaciers formed during the Little Ice Age. The largest of these, the Lyell and Maclure glaciers, occupy the headwaters of the Tuolumne River and have been the subject of detailed scientific study since the late 19th century. We repeated historical photographs, field surveys, and velocity measurements on these glaciers to document their response to climate change. Field surveys and remote sensing data indicate that glacier surface areas have diminished by 67-78% since 1883, with 10% of that loss coinciding with the 2012-2015 California drought. The naturalist John Muir first measured the velocity of the Maclure Glacier in 1872, finding that the glacier moved about 2.6 cm/day during the late summer and early autumn. We reproduced Muir's measurements over the same seasonal period and found the glacier to be moving at the same rate, despite the marked reduction in surface area. Time-averaged velocities measured over a four-year period show strong seasonality, with rates near zero in winter. Much of the present movement of the Maclure Glacier must therefore occur as sliding at the bed, which is apparently enhanced by greater melt. The adjacent Lyell Glacier displayed virtually no movement over the same four-year time period, likely because it has thinned below a critical threshold; both glaciers have thinned by more than 40 m since 1932, with thinning up to 3 m/yr during the 2012-2015 drought. New remote sensing data collected as part of NASA's Airborne Snow Observatory project offer opportunities to measure glacier volume and mass balance changes from 2012 onward. Numerical modeling of glacier mass balance will help to predict the timing of complete glacier loss and to assess the associated hydrological impacts on downstream ecosystems.

  12. Increased Ocean Access to Totten Glacier, East Antarctica

    NASA Astrophysics Data System (ADS)

    Blankenship, D. D.; Greenbaum, J. S.; Young, D. A.; Richter, T. G.; Roberts, J. L.; Aitken, A.; Legresy, B.; Warner, R. C.; van Ommen, T. D.; Siegert, M. J.

    2015-12-01

    The Totten Glacier is the largest ice sheet outlet in East Antarctica, draining 3.5 meters of eustatic sea level potential from the Aurora Subglacial Basin (ASB) into the Sabrina Coast. Recent work has shown that the ASB has drained and filled many times since largescale glaciation began including evidence that it collapsed during the Pliocene. Steady thinning rates observed near Totten Glacier's grounding line since the beginning of the satellite altimetry record are the largest in East Antarctica and the nature of the thinning suggests that it is driven by enhanced basal melting due to ocean processes. Warm Modified Circumpolar Deep Water (MCDW), which has been linked to glacier retreat in West Antarctica, has been observed in summer and winter on the Sabrina Coast continental shelf in the 400-500 m depth range. Using airborne geophysical data acquired over multiple years we delineate seafloor valleys connecting the inner continental shelf to the cavity beneath Totten Glacier that cut through a large sill centered along the ice shelf calving front. The sill shallows to depths of about 300 mbsl and was likely a grounding line pinning point during Holocene retreat, however, the two largest seafloor valleys are deeper than the observed range of thermocline depths. The deeper of the two valleys, a 4 km-wide trough, connects to the ice shelf cavity through an area of the coastline that was previously believed to be grounded but that our analysis demonstrates is floating, revealing a second, deeper entryway to ice shelf cavity. The previous coastline was charted using satellite-based mapping techniques that infer subglacial properties based on surface expression and behavior; the new geophysical analysis techniques we use enable inferences of subglacial characteristics using direct observations of the ice-water interface. The results indicate that Totten Glacier and, by extension, the Aurora Subglacial Basin are vulnerable to MCDW that has been observed on the nearby

  13. The distribution and abundance ofa nuisance native alga, Didymosphenia geminata,in streams of Glacier National Park: Climate drivers and management implications

    USGS Publications Warehouse

    Muhlfeld, Clint C.; Jones, Leslie A.; E. William Schweiger,; Isabel W. Ashton,; Loren L. Bahls,

    2011-01-01

    Didymosphenia geminata (didymo) is a freshwater alga native to North America, including Glacier National Park, Montana. It has long been considered a cold-water species, but has recently spread to lower latitudes and warmer waters, and increasingly forms large blooms that cover streambeds. We used a comprehensive monitoring data set from the National Park Service (NPS) and USGS models of stream temperatures to explore the drivers of didymo abundance in Glacier National Park. We estimate that approximately 64% of the stream length in the park contains didymo, with around 5% in a bloom state. Results suggest that didymo abundance likely increased over the study period (2007–2009), with blooms becoming more common. Our models suggest that didymo abundance is positively related to summer stream temperatures and negatively related to total nitrogen and the distance downstream from lakes. Regional climate model simulations indicate that stream temperatures in the park will likely continue to increase over the coming decades, which may increase the extent and severity of didymo blooms. As a result, didymo may be a useful indicator of thermal and hydrological modification associated with climate warming, especially in a relatively pristine system like Glacier where proximate human-related disturbances are absent or reduced. Glacier National Park plays an important role as a sentinel for climate change and associated education across the Rocky Mountain region.

  14. Mapping tide-water glacier dynamics in east Greenland using landsat data

    USGS Publications Warehouse

    Dwyer, John L.

    1995-01-01

    Landsat multispectral scanner and thematic mapper images were co-registered For the Kangerdlugssuaq Fjord region in East Greenland and were used to map glacier drainage-basin areas, changes in the positions of tide-water glacier termini and to estimate surface velocities of the larger tide-water glaciers. Statistics were compiled to document distance and area changes to glacier termini. The methodologies developed in this study are broadly applicable to the investigation of tide-water glaciers in other areas. The number of images available for consecutive years and the accuracy with which images are co-registered are key factors that influence the degree to which regional glacier dynamics can be characterized using remotely sensed data.Three domains of glacier state were interpreted: net increase in terminus area in the southern part of the study area, net loss of terminus area for glaciers in upper Kangerdlugssuaq Fjord and a slight loss of glacier terminus area northward from Ryberg Fjord. Local increases in the concentrations of drifting icebergs in the fjords coincide with the observed extension of glacier termini positions Ice-surface velocity estimates were derived for several glaciers using automated image cross-correlation techniques The velocity determined for Kangerdlugssuaq Gletscher is approximately 5.0 km a−1 and that for Kong Christian IV Gletscher is 0.9 km a−1. The continuous presence of icebergs and brash ice in front of these glaciers indicates sustained rates of ice-front calving.

  15. Going-to-the-Sun Road rehabilitation plan/final environmental impact statement : Glacier National Park

    DOT National Transportation Integrated Search

    2003-04-01

    Glacier National Park is considering the rehabilitation of the 50-mile (80-kilometer) Going-to-the-Sun Road, a National Historic Landmark. Road rehabilitation is needed to correct structural deficiencies in the deteriorating roadway, improve safety, ...

  16. Preliminary hydrodynamic analysis of landslide-generated waves in Tidal Inlet, Glacier Bay National Park, Alaska

    USGS Publications Warehouse

    Geist, Eric L.; Jakob, Matthias; Wieczoreck, Gerald F.; Dartnell, Peter

    2003-01-01

    A landslide block perched on the northern wall of Tidal Inlet, Glacier Bay National Park (Figure 1), has the potential to generate large waves in Tidal Inlet and the western arm of Glacier Bay if it were to fail catastrophically. Landslide-generated waves are a particular concern for cruise ships transiting through Glacier Bay on a daily basis during the summer months. The objective of this study is to estimate the range of wave amplitudes and periods in the western arm of Glacier Bay from a catastrophic landslide in Tidal Inlet. This study draws upon preliminary findings of a field survey by Wieczorek et al. (2003), and evaluates the effects of variations in landslide source parameters on the wave characteristics.

  17. Mechanisms that Amplify, Attenuate and Deviate Glacier Response to Climate Change in Central East Greenland. (Invited)

    NASA Astrophysics Data System (ADS)

    Jiskoot, H.

    2013-12-01

    A multidecadal review of glacier fluctuations and case-studies of glacier processes and environments in central East Greenland will be used to demonstrate Mechanisms that Amplify, Attenuate and Deviate glacier response to climate forcings (MAAD). The different spatial and temporal scales at which MAAD affect mass balance and ice flow may complicate interpretation and longterm extrapolation of glacier response to climate change. A framework of MAAD characterisation and best-practice for interpreting climate signals while taking into account MAAD will be proposed. Glaciers in the Watkins Bjerge, Geikie Plateau and Stauning Alps regions of central East Greenland (68°-72°N) contain about 50000 km2 of glacierized area peripheral to the Greenland Ice Sheet. Within the region, large north-south and coast-inland climatic gradients, as well as complicated topography and glacier dynamics, result in discrepant glacier behaviour. Average retreat rates have doubled from about 2 to 4 km2 a-1 between the late 20th and early 21st centuries. However, glaciers terminating along the Atlantic coast display two times the retreat, thinning, and acceleration rates compared to glaciers terminating in inland fjords or on land. Despite similar climatic forcing variable glacier behaviour is apparent: individual glacier length change ranges from +57 m a-1 to -428 m a-1, though most retreat -20 to -100 m a-1. Interacting dynamic, mass balance and glacio-morphological mechanisms can amplify, attenuate or deviate glacier response (MAAD) to climate change, thus complicating the climatological interpretation of glacier length, area, and thickness changes. East Greenland MAAD include a range of common positive and negative feedback mechanisms in surface mass balance and terminus and subglacial boundary conditions affecting ice flow, but also mechanisms that have longterm or delayed effects. Certain MAAD may affect glacier change interpretation on multiple timescales: e.g. surging glaciers do not

  18. Glacier Change and Biologic Succession: a new Alaska Summer Research Academy (ASRA) Science Camp Module for Grades 8-12 in Glacier Bay National Park, Alaska

    NASA Astrophysics Data System (ADS)

    Connor, C. L.; Drake, J.; Good, C.; Fatland, R.; Hakala, M.; Woodford, R.; Donohoe, R.; Brenner, R.; Moriarty, T.

    2008-12-01

    During the summer of 2008, university faculty and instructors from southeast Alaska joined the University Alaska Fairbanks(UAF)Alaska Summer Research Academy(ASRA)to initiate a 12-day module on glacier change and biologic succession in Glacier Bay National Park. Nine students from Alaska, Colorado, Massachusetts, and Texas, made field observations and collected data while learning about tidewater glacier dynamics, plant succession, post-glacial uplift, and habitat use of terrestrial and marine vertebrates and invertebrates in this dynamic landscape that was covered by 6,000 km2 of ice just 250 years ago. ASRA students located their study sites using GPS and created maps in GIS and GOOGLE Earth. They deployed salinometers and temperature sensors to collect vertical profiles of seawater characteristics up-bay near active tidewater glacier termini and down-bay in completely deglaciated coves. ASRA student data was then compared with data collected during the same time period by Juneau undergraduates working on the SEAMONSTER project in Mendenhall Lake. ASRA students traversed actively forming, up-bay recessional moraines devoid of vegetation, and the fully reforested Little Ice Age terminal moraine near Park Headquarters in the lower bay region. Students surveyed marine organisms living between supratidal and subtidal zones near glaciers and far from glaciers, and compared up-bay and down-bay communities. Students made observations and logged sightings of bird populations and terrestrial mammals in a linear traverse from the bay's northwestern most fjord near Mt. Fairweather for 120 km to the bay's entrance, south of Park Headquarters at Bartlett Cove. One student constructed an ROV and was able to deploy a video camera and capture changing silt concentrations in the water column as well as marine life on the fjord bottom. Students also observed exhumed Neoglacial spruce forests and visited outcrops of Silurian reef faunas, now fossilized in Alexander terrane

  19. Sculpted by water, elevated by earthquakes—The coastal landscape of Glacier Bay National Park, Alaska

    USGS Publications Warehouse

    Witter, Robert C.; LeWinter, Adam; Bender, Adrian M.; Glennie, Craig; Finnegan, David

    2017-05-22

    Within Glacier Bay National Park in southeastern Alaska, the Fairweather Fault represents the onshore boundary between two of Earth’s constantly moving tectonic plates: the North American Plate and the Yakutat microplate. Satellite measurements indicate that during the past few decades the Yakutat microplate has moved northwest at a rate of nearly 5 centimeters per year relative to the North American Plate. Motion between the tectonic plates results in earthquakes on the Fairweather Fault during time intervals spanning one or more centuries. For example, in 1958, a 260-kilometer section of the Fairweather Fault ruptured during a magnitude 7.8 earthquake, causing permanent horizontal (as much as 6.5 meters) and vertical (as much as 1 meter) displacement of the ground surface across the fault. Thousands to millions of years of tectonic plate motion, including earthquakes like the one in 1958, raised and shifted the ground surface across the Fairweather Fault, while rivers, glaciers, and ocean waves eroded and sculpted the surrounding landscape along the Gulf of Alaska coast in Glacier Bay National Park.

  20. Ground-nesting marine birds and potential for human disturbance in Glacier Bay National Park

    USGS Publications Warehouse

    Arimitsu, Mayumi L.; Romano, Marc D.; Piatt, John F.; Piatt, John F.; Gende, S.M.

    2004-01-01

    Glacier Bay National Park and Preserve contains a diverse assemblage of marine birds that use the area for nesting, foraging and molting. The abundance and diversity of marine bird species in Glacier Bay is unmatched in the region, due in part to the geomorphic and successional characteristics that result in a wide array of habitat types (Robards and others, 2003). The opportunity for proactive management of these species is unique in Glacier Bay National Park because much of the suitable marine bird nesting habitat occurs in areas designated as wilderness. Ground-nesting marine birds are vulnerable to human disturbance wherever visitors can access nest sites during the breeding season. Human disturbance of nest sites can be significant because intense parental care is required for egg and hatchling survival, and repeated disturbance can result in reduced productivity (Leseberg and others, 2000). Temporary nest desertion by breeding birds in disturbed areas can lead to increased predation on eggs and hatchlings by conspecifics or other predators (Bolduc and Guillemette, 2003). Human disturbance of ground-nesting birds may also affect incubation time and adult foraging success, which in turn can alter breeding success (Verhulst and others, 2001). Furthermore, human activity can potentially cause colony failure when disturbance prevents the initiation of nesting (Hatch, 2002). There is management concern about the susceptibility of breeding birds to disturbance from human activities, but little historical data has been collected on the distribution of ground-nesting marine birds in Glacier Bay. This report summarizes results obtained during two years of a three-year study to determine the distribution of ground-nesting marine birds in Glacier Bay, and the potential for human disturbance of those nesting birds.

  1. The distribution and abundance of a nuisance native alga, Didymosphen Didymosphenia geminata, in streams of Glacier National Park: Climate drivers and management implications

    USGS Publications Warehouse

    William, Schweiger E.; Ashton, I.W.; Muhlfeld, C.C.; Jones, L.A.; Bahls, L.L.

    2011-01-01

    Didymosphenia geminata (didymo) is a freshwater alga native to North America, including Glacier National Park, Montana. It has long been considered a cold-water species, but has recently spread to lower latitudes and warmer waters, and increasingly forms large blooms that cover streambeds. We used a comprehensive monitoring data set from the National Park Service (NPS) and USGS models of stream temperatures to explore the drivers of didymo abundance in Glacier National Park. We estimate that approximately 64% of the stream length in the park contains didymo, with around 5% in a bloom state. Results suggest that didymo abundance likely increased over the study period (2007-2009), with blooms becoming more common. Our models suggest that didymo abundance is positively related to summer stream temperatures and negatively related to total nitrogen and the distance downstream from lakes. Regional climate model simulations indicate that stream temperatures in the park will likely continue to increase over the coming decades, which may increase the extent and severity of didymo blooms. As a result, didymo may be a useful indicator of thermal and hydrological modification associated with climate warming, especially in a relatively pristine system like Glacier where proximate human-related disturbances are absent or reduced. Glacier National Park plays an important role as a sentinel for climate change and associated education across the Rocky Mountain region.

  2. Bear-human interactions at Glacier Bay National Park and Preserve: Conflict risk assessment

    USGS Publications Warehouse

    Smith, Tom S.; DeBruyn, Terry D.; Lewis, Tania; Yerxa, Rusty; Partridge, Steven T.

    2003-01-01

    Many bear-human conflicts have occurred in Alaska parks and refuges, resulting in area closures, property damage, human injury, and loss of life. Human activity in bear country has also had negative and substantial consequences for bears: disruption of their natural activity patterns, displacement from important habitats, injury, and death. It is unfortunate for both people and bears when conflicts occur. Fortunately, however, solutions exist for reducing, and in some instances eliminating, bear-human conflict. This article presents ongoing work at Glacier Bay National Park and Preserve by U.S. Geological Survey (USGS) and National Park Service scientists who are committed to finding solutions for the bear-human conflicts that periodically occurs there.

  3. Holocene glacier and deep water dynamics, Adélie Land region, East Antarctica

    NASA Astrophysics Data System (ADS)

    Denis, Delphine; Crosta, Xavier; Schmidt, Sabine; Carson, Damien S.; Ganeshram, Raja S.; Renssen, Hans; Bout-Roumazeilles, Viviane; Zaragosi, Sebastien; Martin, Bernard; Cremer, Michel; Giraudeau, Jacques

    2009-06-01

    This study presents a high-resolution multi-proxy investigation of sediment core MD03-2601 and documents major glacier oscillations and deep water activity during the Holocene in the Adélie Land region, East Antarctica. A comparison with surface ocean conditions reveals synchronous changes of glaciers, sea ice and deep water formation at Milankovitch and sub-Milankovitch time scales. We report (1) a deglaciation of the Adélie Land continental shelf from 11 to 8.5 cal ka BP, which occurred in two phases of effective glacier grounding-line retreat at 10.6 and 9 cal ka BP, associated with active deep water formation; (2) a rapid glacier and sea ice readvance centred around 7.7 cal ka BP; and (3) five rapid expansions of the glacier-sea ice systems, during the Mid to Late Holocene, associated to a long-term increase of deep water formation. At Milankovich time scales, we show that the precessionnal component of insolation at high and low latitudes explains the major trend of the glacier-sea ice-ocean system throughout the Holocene, in the Adélie Land region. In addition, the orbitally-forced seasonality seems to control the coastal deep water formation via the sea ice-ocean coupling, which could lead to opposite patterns between north and south high latitudes during the Mid to Late Holocene. At sub-Milankovitch time scales, there are eight events of glacier-sea ice retreat and expansion that occurred during atmospheric cooling events over East Antarctica. Comparisons of our results with other peri-Antarctic records and model simulations from high southern latitudes may suggest that our interpretation on glacier-sea ice-ocean interactions and their Holocene evolutions reflect a more global Antarctic Holocene pattern.

  4. Morphological evidence and direct estimates of rapid melting beneath Totten Glacier Ice Shelf, East Antarctica

    NASA Astrophysics Data System (ADS)

    Greenbaum, Jamin; Schroeder, Dustin; Grima, Cyril; Habbal, Feras; Dow, Christine; Roberts, Jason; Gwyther, David; van Ommen, Tas; Siegert, Martin; Blankenship, Donald

    2017-04-01

    Totten Glacier drains at least 3.5 meters of eustatic sea level potential from marine-based ice in the Aurora Subglacial Basin (ASB) in East Antarctica, more than the combined total of all glaciers in West Antarctica. Totten Glacier has been the most rapidly thinning glacier in East Antarctica since satellite altimetry time series began and the nature of the thinning suggests that it is driven by enhanced basal melting due to ocean processes. While grounded ice thinning rates have been steady, recent work has shown that Totten's floating ice shelf may not have the same thinning behavior; as a result, it is critical to observe ice shelf and cavity boundary conditions and basal processes to understand this apparent discrepancy. Warm Modified Circumpolar Deep Water (MCDW), which has been linked to glacier retreat in West Antarctica, has been observed in summer and winter on the nearby Sabrina Coast continental shelf and deep depressions in the seafloor provide access for MCDW to reach the ice shelf cavity. Given its northern latitude, numerical ice sheet modeling indicates that Totten Glacier may be prone to retreat caused by hydrofracture in a warming climate, so it is important to understand how intruding MCDW is affecting thinning of Totten Glacier's ice shelf. Here we use post-processed, focused airborne radar observations of the Totten Glacier Ice Shelf to delineate multi-km wide basal channels and flat basal terraces associated with high basal reflectivity and specularity (flatness) anomalies and correspondingly large ice surface depressions that indicate active basal melting. Using a simple temperature-attenuation model, and basal roughness corrections, we present basal melt rates associated with the radar reflection and specularity anomalies and compare them to those derived from numerical ocean circulation modeling and an ice flow divergence calculation. Sub-ice shelf ocean circulation modeling and under-ice robotic observations of Pine Island Glacier Ice

  5. The Glacier National Park: A popular guide to its geology and scenery

    USGS Publications Warehouse

    Campbell, Marius R.

    1914-01-01

    The Glacier National Park includes that part of the Front Range of the Rocky Mountains lying just south of the Canadian line, in Teton and Flathead counties, Mont. It is bounded on the west by Flathead River (locally called North Fork), on the south by the Middle Fork of Flathead River and the Great Northern Railway, and on the east by the Blackfeet Indian Reservation. Although this part of the Rocky Mountains has been known since Lewis and Clark crossed the continent in 1805-6, the region later made a park appears not to have been visited by white men until 1853, when Cut Bank Pass was crossed by A. W. Tinkham, one of the Government engineers engaged in exploring a route for the Pacific railroad. Tinkham, who was encamped in the Bitterroot Valley, was ordered to examine Marias Pass, but in traversing Middle Fork of Flathead River along the line of the present railroad he was evidently misled by the large size of the valley of Nyack Creek and ascended that instead of keeping to the right up the main stream. He reported the pass impracticable for railroad construction, and so this region dropped out of public attention for a long time. The next explorers to enter the region were a group of surveyors who, under the direction of American and British commissioners, established the international boundary line along the forty-ninth parallel from the Pacific coast to the main summit of the Rocky Mountains. This party reached the area now included in the park in the summer of 1861, and the stone monument shown in Plate I, B, which they erected on the Continental Divide west of Waterton Lake, still marks a point on the boundary between the United States and Canada. The land on the west side of the range formed a part of the public domain which, until the erection of the park, was open to settlement, but the land on the east originally belonged to the Blackfeet Indians and the white men had no rights upon it. About 1890 copper ore was found near the heads of Quartz and

  6. Genetic status and conservation of Westslope Cutthroat Trout in Glacier National Park

    USGS Publications Warehouse

    Muhlfeld, Clint C.; D'Angelo, Vincent S.; Downs, Christopher C.; Powell, John D.; Amish, Stephen J.; Luikart, Gordon; Kovach, Ryan; Boyer, Matthew; Kalinowski, Steven T.

    2016-01-01

    Invasive hybridization is one of the greatest threats to the persistence of Westslope Cutthroat Trout Oncorhynchus clarkii lewisi. Large protected areas, where nonhybridized populations are interconnected and express historical life history and genetic diversity, provide some of the last ecological and evolutionary strongholds for conserving this species. Here, we describe the genetic status and distribution of Westslope Cutthroat Trout throughout Glacier National Park, Montana. Admixture between Westslope Cutthroat Trout and introduced Rainbow Trout O. mykiss and Yellowstone Cutthroat Trout O. clarkii bouvieri was estimated by genotyping 1,622 fish collected at 115 sites distributed throughout the Columbia, Missouri, and South Saskatchewan River drainages. Currently, Westslope Cutthroat Trout occupy an estimated 1,465 km of stream habitat and 45 lakes (9,218 ha) in Glacier National Park. There was no evidence of introgression in samples from 32 sites along 587 km of stream length (40% of the stream kilometers currently occupied) and 17 lakes (2,555 ha; 46% of the lake area currently occupied). However, nearly all (97%) of the streams and lakes that were occupied by nonhybridized populations occurred in the Columbia River basin. Based on genetic status (nonnative genetic admixture ≤ 10%), 36 Westslope Cutthroat Trout populations occupying 821 km of stream and 5,482 ha of lakes were identified as “conservation populations.” Most of the conservation populations (N = 27; 736 km of stream habitat) occurred in the Columbia River basin, whereas only a few geographically restricted populations were found in the South Saskatchewan River (N = 7; 55 km) and Missouri River (N = 2; 30 km) basins. Westslope Cutthroat Trout appear to be at imminent risk of genomic extinction in the South Saskatchewan and Missouri River basins, whereas populations in the Columbia River basin are widely distributed and conservation efforts are actively addressing threats from

  7. Velocities along Byrd Glacier, East Antarctica, derived from Automatic Feature Tracking

    NASA Astrophysics Data System (ADS)

    Stearns, L. A.; Hamilton, G. S.

    2003-12-01

    Automatic feature tracking techniques are applied to recently acquired ASTER (Advanced Spaceborne Thermal Emission and Reflection Radiometer) imagery in order to determine the velocity field of Byrd Glacier, East Antarctica. The software IMCORR tracks the displacement of surface features (crevasses, drift mounds) in time sequential images, to produce the velocity field. Due to its high resolution, ASTER imagery is ideally suited for detecting small features changes. The produced result is a dense array of velocity vectors, which allows more thorough characterization of glacier dynamics. Byrd Glacier drains approximately 20.5 km3 of ice into the Ross Ice Shelf every year. Previous studies have determined ice velocities for Byrd Glacier by using photogrammetry, field measurements and manual feature tracking. The most recent velocity data is from 1986 and, as evident in the West Antarctic ice streams, substantial changes in velocity can occur on decadal time scales. The application of ASTER-based velocities fills this time lapse, and increased temporal resolution allows for a more complete analysis of Byrd Glacier. The ASTER-derived ice velocities are used in updating mass balance and force budget calculations to assess the stability of Byrd Glacier. Ice thickness information from BEDMAP, surface slopes from the OSUDEM and a compilation of accumulation rates are used to complete the calculations.

  8. Preliminary assessment of landslide-induced wave hazards, Tidal Inlet, Glacier Bay National Park, Alaska

    USGS Publications Warehouse

    Wieczorek, Gerald F.; Jakob, Matthias; Motyka, Roman J.; Zirnheld, Sandra L.; Craw, Patricia

    2003-01-01

    A large potential rock avalanche above the northern shore of Tidal Inlet, Glacier Bay National Park, Alaska, was investigated to determine hazards and risks of landslide-induced waves to cruise ships and other park visitors. Field and photographic examination revealed that the 5 to 10 million cubic meter landslide moved between AD 1892 and 1919 after the retreat of Little Ice Age glaciers from Tidal Inlet by AD 1890. The timing of landslide movement and the glacial history suggest that glacial debuttressing caused weakening of the slope and that the landslide could have been triggered by large earthquakes of 1899-1900 in Yakutat Bay. Evidence of recent movement includes fresh scarps, back-rotated blocks, and smaller secondary landslide movements. However, until there is evidence of current movement, the mass is classified as a dormant rock slump. An earthquake on the nearby active Fairweather fault system could reactivate the landslide and trigger a massive rock slump and debris avalanche into Tidal Inlet. Preliminary analyses show that waves induced by such a landslide could travel at speeds of 45 to 50 m/s and reach heights up to 76 m with wave runups of 200 m on the opposite shore of Tidal Inlet. Such waves would not only threaten vessels in Tidal Inlet, but would also travel into the western arm of Glacier Bay endangering large cruise ships and their passengers.

  9. Monitoring population status of sea otters (Enhydra lutris) in Glacier Bay National Park and Preserve, Alaska: options and considerations

    USGS Publications Warehouse

    Esslinger, George G.; Esler, Daniel N.; Howlin, S.; Starcevich, L.A.

    2015-06-25

    After many decades of absence from southeast Alaska, sea otters (Enhydra lutris) are recolonizing parts of their former range, including Glacier Bay, Alaska. Sea otters are well known for structuring nearshore ecosystems and causing community-level changes such as increases in kelp abundance and changes in the size and number of other consumers. Monitoring population status of sea otters in Glacier Bay will help park researchers and managers understand and interpret sea otter-induced ecosystem changes relative to other sources of variation, including potential human-induced impacts such as ocean acidification, vessel disturbance, and oil spills. This report was prepared for the National Park Service (NPS), Southeast Alaska Inventory and Monitoring Network following a request for evaluation of options for monitoring sea otter population status in Glacier Bay National Park and Preserve. To meet this request, we provide a detailed consideration of the primary method of assessment of abundance and distribution, aerial surveys, including analyses of power to detect interannual trends and designs to reduce variation around annual abundance estimates. We also describe two alternate techniques for evaluating sea otter population status—(1) quantifying sea otter diets and energy intake rates, and (2) detecting change in ages at death. In addition, we provide a brief section on directed research to identify studies that would further our understanding of sea otter population dynamics and effects on the Glacier Bay ecosystem, and provide context for interpreting results of monitoring activities.

  10. Pan–ice-sheet glacier terminus change in East Antarctica reveals sensitivity of Wilkes Land to sea-ice changes

    PubMed Central

    Miles, Bertie W. J.; Stokes, Chris R.; Jamieson, Stewart S. R.

    2016-01-01

    The dynamics of ocean-terminating outlet glaciers are an important component of ice-sheet mass balance. Using satellite imagery for the past 40 years, we compile an approximately decadal record of outlet-glacier terminus position change around the entire East Antarctic Ice Sheet (EAIS) marine margin. We find that most outlet glaciers retreated during the period 1974–1990, before switching to advance in every drainage basin during the two most recent periods, 1990–2000 and 2000–2012. The only exception to this trend was in Wilkes Land, where the majority of glaciers (74%) retreated between 2000 and 2012. We hypothesize that this anomalous retreat is linked to a reduction in sea ice and associated impacts on ocean stratification, which increases the incursion of warm deep water toward glacier termini. Because Wilkes Land overlies a large marine basin, it raises the possibility of a future sea level contribution from this sector of East Antarctica. PMID:27386519

  11. Pan-ice-sheet glacier terminus change in East Antarctica reveals sensitivity of Wilkes Land to sea-ice changes.

    PubMed

    Miles, Bertie W J; Stokes, Chris R; Jamieson, Stewart S R

    2016-05-01

    The dynamics of ocean-terminating outlet glaciers are an important component of ice-sheet mass balance. Using satellite imagery for the past 40 years, we compile an approximately decadal record of outlet-glacier terminus position change around the entire East Antarctic Ice Sheet (EAIS) marine margin. We find that most outlet glaciers retreated during the period 1974-1990, before switching to advance in every drainage basin during the two most recent periods, 1990-2000 and 2000-2012. The only exception to this trend was in Wilkes Land, where the majority of glaciers (74%) retreated between 2000 and 2012. We hypothesize that this anomalous retreat is linked to a reduction in sea ice and associated impacts on ocean stratification, which increases the incursion of warm deep water toward glacier termini. Because Wilkes Land overlies a large marine basin, it raises the possibility of a future sea level contribution from this sector of East Antarctica.

  12. Black bear density in Glacier National Park, Montana

    USGS Publications Warehouse

    Stetz, Jeff B.; Kendall, Katherine C.; Macleod, Amy C.

    2013-01-01

    We report the first abundance and density estimates for American black bears (Ursus americanus) in Glacier National Park (NP),Montana, USA.We used data from 2 independent and concurrent noninvasive genetic sampling methods—hair traps and bear rubs—collected during 2004 to generate individual black bear encounter histories for use in closed population mark–recapture models. We improved the precision of our abundance estimate by using noninvasive genetic detection events to develop individual-level covariates of sampling effort within the full and one-half mean maximum distance moved (MMDM) from each bear’s estimated activity center to explain capture probability heterogeneity and inform our estimate of the effective sampling area.Models including the one-halfMMDMcovariate received overwhelming Akaike’s Information Criterion support suggesting that buffering our study area by this distance would be more appropriate than no buffer or the full MMDM buffer for estimating the effectively sampled area and thereby density. Our modelaveraged super-population abundance estimate was 603 (95% CI¼522–684) black bears for Glacier NP. Our black bear density estimate (11.4 bears/100 km2, 95% CI¼9.9–13.0) was consistent with published estimates for populations that are sympatric with grizzly bears (U. arctos) and without access to spawning salmonids. Published 2013. This article is a U.S. Government work and is in the public domain in the USA.

  13. Organochlorine compounds and current-use pesticides in snow and lake sediment in Rocky Mountain National Park, Colorado, and Glacier National Park, Montana, 2002-03

    USGS Publications Warehouse

    Mast, M. Alisa; Foreman, William T.; Skaates, Serena V.

    2006-01-01

    Organochlorine compounds and current-use pesticides were measured in snow and lake-sediment samples from Rocky Mountain National Park in Colorado and Glacier National Park in Montana to determine their occurrence and distribution in high-elevation aquatic ecosystems. The U.S. Geological Survey, in cooperation with the National Park Service, collected snow samples at eight sites in Rocky Mountain National Park and at eight sites in Glacier National Park during spring of 2002 and 2003 just prior to the start of snowmelt. Surface sediments were collected from 11 lakes in Rocky Mountain National Park and 10 lakes in Glacier National Park during summer months of 2002 and 2003. Samples were analyzed for organochlorine compounds by gas chromatography with electron-capture detection and current-use pesticides by gas chromatography with electron-impact mass spectrometry. A subset of samples was reanalyzed using a third instrumental technique (gas chromatography with electron-capture negative ion mass spectrometry) to verify detected concentrations in the initial analysis and to investigate the presence of additional compounds. For the snow samples, the pesticides most frequently detected were endosulfan, dacthal, and chlorothalonil, all of which are chlorinated pesticides that currently are registered for use in North America. Concentrations of these pesticides in snow were very low, ranging from 0.07 to 2.36 nanograms per liter. Of the historical-use pesticides, hexachlorobenzene, dieldrin, and trans-nonachlor were detected in snow but only in one sample each. Annual deposition rates of dacthal, endosulfan, and chlorothalonil were estimated at 0.7 to 3.0 micrograms per square meter. These estimates are likely biased low because they do not account for pesticide deposition during summer months. For the lake-sediment samples, DDE (p,p'-dichlorodiphenyldichoroethene) and DDD (p,p'-dichlorodiphenyldichoroethane) were the most frequently detected organochlorine compounds. DDE

  14. Glacier National Park : Going-to-the-Sun Road Corridor management plan - existing conditions of the transportation system

    DOT National Transportation Integrated Search

    2014-06-27

    The Going-to-the-Sun Road (GTSR) Corridor has been undergoing major shifts in use due to ongoing construction, implementation of a shuttle system, and changes in visitor use patterns. Glacier National Park (GLAC) is developing the GTSR Transportation...

  15. Olympic National Park

    NASA Image and Video Library

    2017-12-08

    It has to be one of America’s most diverse national park landscapes. If you walked from west to east across Olympic National Park, you would start at the rocky Pacific shoreline, move into rare temperate rainforests and lush river valleys, ascend glaciers and rugged mountain peaks, and then descend into a comparatively dry rain shadow and alpine forest. From the beach to the top of Mount Olympus, you would rise 7,980 feet (2430 meters) above sea level. Situated on the Olympic Peninsula in northwestern Washington, these lands were first set aside as a national monument in 1909 by Theodore Roosevelt. Twenty-nine years later, his cousin Franklin officially established Olympic National Park. International institutions have also made a case for treasuring this land, as the area was declared an International Biosphere Reserve in 1976 and a World Heritage Site in 1981. The park encompasses nearly 923,000 acres of wild lands, including 60 named glaciers, 73 miles of coast, and 3,000 miles of rivers and streams. Read more: go.nasa.gov/2bRmzSJ Credit: NASA/Landsat8 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. Duplex development and abandonment during evolution of the Lewis thrust system, southern Glacier National Park, Montana

    NASA Astrophysics Data System (ADS)

    Yin, An; Kelty, Thomas K.; Davis, Gregory A.

    1989-09-01

    Geologic mapping in southern Glacier National Park, Montana, reveals the presence of two duplexes sharing the same floor thrust fault, the Lewis thrust. The westernmost duplex (Brave Dog Mountain) includes the low-angle Brave Dog roof fault and Elk Mountain imbricate system, and the easternmost (Rising Wolf Mountain) duplex includes the low-angle Rockwell roof fault and Mt. Henry imbricate system. The geometry of these duplexes suggests that they differ from previously described geometric-kinematic models for duplex development. Their low-angle roof faults were preexisting structures that were locally utilized as roof faults during the formation of the imbricate systems. Crosscutting of the Brave Dog fault by the Mt. Henry imbricate system indicates that the two duplexes formed at different times. The younger Rockwell-Mt. Henry duplex developed 20 km east of the older Brave Dog-Elk Mountain duplex; the roof fault of the former is at a higher structural level. Field relations confirm that the low-angle Rockwell fault existed across the southern Glacier Park area prior to localized formation of the Mt. Henry imbricate thrusts beneath it. These thrusts kinematically link the Rockwell and Lewis faults and may be analogous to P shears that form between two synchronously active faults bounding a simple shear system. The abandonment of one duplex and its replacement by another with a new and higher roof fault may have been caused by (1) warping of the older and lower Brave Dog roof fault during the formation of the imbricate system (Elk Mountain) beneath it, (2) an upward shifting of the highest level of a simple shear system in the Lewis plate to a new decollement level in subhorizontal belt strata (= the Rockwell fault) that lay above inclined strata within the first duplex, and (3) a reinitiation of P-shear development (= Mt. Henry imbricate faults) between the Lewis thrust and the subparallel, synkinematic Rockwell fault.

  17. Antarctic Peninsula Tidewater Glacier Dynamics

    NASA Astrophysics Data System (ADS)

    Pettit, E. C.; Scambos, T. A.; Haran, T. M.; Wellner, J. S.; Domack, E. W.; Vernet, M.

    2015-12-01

    The northern Antarctic Peninsula (nAP, north of 66°S) is a north-south trending mountain range extending transverse across the prevailing westerly winds of the Southern Ocean resulting in an extreme west-to-east precipitation gradient. Snowfall on the west side of the AP is one to two orders of magnitude higher than the east side. This gradient drives short, steep, fast-flowing glaciers into narrow fjords on the west side, while longer lower-sloping glaciers flow down the east side into broader fjord valleys. This pattern in ice dynamics affects ice-ocean interaction on timescales of decades to centuries, and shapes the subglacial topography and submarine bathymetry on timescales of glacial cycles. In our study, we calculate ice flux for the western and eastern nAP using a drainage model that incorporates the modern ice surface topography, the RACMO-2 precipitation estimate, and recent estimates of ice thinning. Our results, coupled with observed rates of ice velocity from InSAR (I. Joughin, personal communication) and Landsat 8 -derived flow rates (this study), provide an estimate of ice thickness and fjord depth in grounded-ice areas for the largest outlet glaciers. East-side glaciers either still terminate in or have recently terminated in ice shelves. Sedimentary evidence from the inner fjords of the western glaciers indicates they had ice shelves during LIA time, and may still have transient floating ice tongues (tabular berg calvings are observed). Although direct oceanographic evidence is limited, the high accumulation rate and rapid ice flux implies cold basal ice for the western nAP glaciers and therefore weak subglacial discharge relative to eastern nAP glaciers and or other tidewater fjord systems such as in Alaska. Finally, despite lower accumulation rates on the east side, the large elongate drainage basins result in a greater ice flux funneled through fewer deeper glaciers. Due to the relation between ice flux and erosion, these east-side glaciers

  18. 36 CFR 13.1109 - Off-road vehicle use in Glacier Bay National Preserve.

    Code of Federal Regulations, 2012 CFR

    2012-07-01

    ... Glacier Bay National Preserve. 13.1109 Section 13.1109 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Administrative Provisions § 13.1109 Off-road vehicle use in Glacier Bay National...

  19. 36 CFR 13.1109 - Off-road vehicle use in Glacier Bay National Preserve.

    Code of Federal Regulations, 2011 CFR

    2011-07-01

    ... Glacier Bay National Preserve. 13.1109 Section 13.1109 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Administrative Provisions § 13.1109 Off-road vehicle use in Glacier Bay National...

  20. 36 CFR 13.1109 - Off-road vehicle use in Glacier Bay National Preserve.

    Code of Federal Regulations, 2013 CFR

    2013-07-01

    ... Glacier Bay National Preserve. 13.1109 Section 13.1109 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Administrative Provisions § 13.1109 Off-road vehicle use in Glacier Bay National...

  1. 36 CFR 13.1109 - Off-road vehicle use in Glacier Bay National Preserve.

    Code of Federal Regulations, 2014 CFR

    2014-07-01

    ... Glacier Bay National Preserve. 13.1109 Section 13.1109 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Administrative Provisions § 13.1109 Off-road vehicle use in Glacier Bay National...

  2. 36 CFR 13.1109 - Off-road vehicle use in Glacier Bay National Preserve.

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... Glacier Bay National Preserve. 13.1109 Section 13.1109 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Administrative Provisions § 13.1109 Off-road vehicle use in Glacier Bay National...

  3. 36 CFR 13.1116 - Do I need a camping permit in Glacier Bay?

    Code of Federal Regulations, 2012 CFR

    2012-07-01

    ... Glacier Bay? 13.1116 Section 13.1116 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve General Provisions § 13.1116 Do I need a camping permit in Glacier Bay? From May 1...

  4. 36 CFR 13.1116 - Do I need a camping permit in Glacier Bay?

    Code of Federal Regulations, 2011 CFR

    2011-07-01

    ... Glacier Bay? 13.1116 Section 13.1116 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve General Provisions § 13.1116 Do I need a camping permit in Glacier Bay? From May 1...

  5. 2001-2010 glacier changes in the Central Karakoram National Park: a contribution to evaluate the magnitude and rate of the "Karakoram anomaly"

    NASA Astrophysics Data System (ADS)

    Minora, U.; Bocchiola, D.; D'Agata, C.; Maragno, D.; Mayer, C.; Lambrecht, A.; Mosconi, B.; Vuillermoz, E.; Senese, A.; Compostella, C.; Smiraglia, C.; Diolaiuti, G.

    2013-06-01

    Karakoram is one of the most glacierized region worldwide, and glaciers therein are the main water resource of Pakistan. The attention paid to this area is increasing, because the evolution of its glaciers recently depicted a situation of general stability, known as "Karakoram Anomaly", in contrast to glacier retreat worldwide. Here we focused our attention upon the glacier evolution within the Central Karakoram National Park (CKNP, a newborn park of this region, ca. 12 162 km2 in area) to assess the magnitude and rate of such anomaly. By means of Remote Sensing data (i.e.: Landsat images), we analyzed a sample of more than 700 glaciers, and we found out their area change between 2001 and 2010 is not significant (+27 km2 ± 42 km2), thus confirming their stationarity. We analyzed climate data, snow coverage from MODIS, and supraglacial debris presence, as well as potential (con-) causes. We found a slight decrease of summer temperatures (down to -1.5 °C during 1980-2009) and an increase of wet days during winter (up +3.3 days yr-1 during 1980-2009), possibly increasing snow cover duration, consistently with MODIS data. We further detected considerable supra-glacial debris coverage (ca. 20% of the glacier area which rose up to 31% considering only the ablation area), which could have reduced buried ice melting during the last decade. These results provide further ground to uphold the existence of the Karakoram Anomaly, and present an useful template for assessment of water availability within the glaciers of the CKNP.

  6. Rapid Holocene thinning of an East Antarctic outlet glacier driven by marine ice sheet instability

    PubMed Central

    Jones, R. S.; Mackintosh, A. N.; Norton, K. P.; Golledge, N. R.; Fogwill, C. J.; Kubik, P. W.; Christl, M.; Greenwood, S. L.

    2015-01-01

    Outlet glaciers grounded on a bed that deepens inland and extends below sea level are potentially vulnerable to ‘marine ice sheet instability'. This instability, which may lead to runaway ice loss, has been simulated in models, but its consequences have not been directly observed in geological records. Here we provide new surface-exposure ages from an outlet of the East Antarctic Ice Sheet that reveal rapid glacier thinning occurred approximately 7,000 years ago, in the absence of large environmental changes. Glacier thinning persisted for more than two and a half centuries, resulting in hundreds of metres of ice loss. Numerical simulations indicate that ice surface drawdown accelerated when the otherwise steadily retreating glacier encountered a bedrock trough. Together, the geological reconstruction and numerical simulations suggest that centennial-scale glacier thinning arose from unstable grounding line retreat. Capturing these instability processes in ice sheet models is important for predicting Antarctica's future contribution to sea level change. PMID:26608558

  7. 36 CFR 13.1150 - Is a permit required for a vessel in Glacier Bay?

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... vessel in Glacier Bay? 13.1150 Section 13.1150 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Vessel Permits § 13.1150 Is a permit required for a vessel in Glacier Bay? A permit from...

  8. 36 CFR 13.1150 - Is a permit required for a vessel in Glacier Bay?

    Code of Federal Regulations, 2011 CFR

    2011-07-01

    ... vessel in Glacier Bay? 13.1150 Section 13.1150 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Vessel Permits § 13.1150 Is a permit required for a vessel in Glacier Bay? A permit from...

  9. 36 CFR 13.1150 - Is a permit required for a vessel in Glacier Bay?

    Code of Federal Regulations, 2012 CFR

    2012-07-01

    ... vessel in Glacier Bay? 13.1150 Section 13.1150 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Vessel Permits § 13.1150 Is a permit required for a vessel in Glacier Bay? A permit from...

  10. 36 CFR 13.1150 - Is a permit required for a vessel in Glacier Bay?

    Code of Federal Regulations, 2014 CFR

    2014-07-01

    ... vessel in Glacier Bay? 13.1150 Section 13.1150 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Vessel Permits § 13.1150 Is a permit required for a vessel in Glacier Bay? A permit from...

  11. Factors influencing the distribution of native bull trout and westslope cutthroat trout in western Glacier National Park, Montana

    USGS Publications Warehouse

    D'Angelo, Vincent S.; Muhlfeld, Clint C.

    2013-01-01

    The widespread declines of native bull trout (Salvelinus confluentus) and westslope cutthroat trout (Oncorhynchus clarkii lewisi) populations prompted researchers to investigate factors influencing their distribution and status in western Glacier National Park, Montana. We evaluated the association of a suite of abiotic factors (stream width, elevation, gradient, large woody debris density, pool density, August mean stream temperature, reach surface area) with the occurrence (presence or absence) of bull trout and westslope cutthroat trout in 79 stream reaches in five sub-drainages containing glacial lakes. We modeled the occurrence of each species using logistic regression and evaluated competing models using an information theoretic approach. Westslope cutthroat trout were widely distributed (47 of 79 reaches), and there appeared to be no restrictions on their distribution other than physical barriers. Westslope cutthroat trout were most commonly found in relatively warm reaches downstream of lakes and in headwater reaches with large amounts of large woody debris and abundant pools. By contrast, bull trout were infrequently detected (10 of 79 reaches), with 7 of the 10 (70%) detections in sub-drainages that have not been compromised by non-native lake trout (S. namaycush). Bull trout were most often found in cold, low-gradient reaches upstream of glacial lakes. Our results indicate that complex stream habitats in sub-drainages free of non-native species are important to the persistence of native salmonids in western Glacier National Park. Results from this study may help managers monitor and protect important habitats and populations, inform conservation and recovery programs, and guide non-native species suppression efforts in Glacier National Park and elsewhere.

  12. View of east elevation of Building No. 23. Parking Area ...

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

    View of east elevation of Building No. 23. Parking Area No. 25 in foreground. Looking west-northwest - Easter Hill Village, Building No. 23, North side of South Twenty-sixth Street, east of Corto Square, Richmond, Contra Costa County, CA

  13. 36 CFR 13.1312 - Climbing and walking on Exit Glacier.

    Code of Federal Regulations, 2014 CFR

    2014-07-01

    ... Glacier. 13.1312 Section 13.1312 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF... General Provisions § 13.1312 Climbing and walking on Exit Glacier. Except for areas designated by the Superintendent, climbing or walking on, in, or under Exit Glacier is prohibited within 1/2 mile of the glacial...

  14. 36 CFR 13.1312 - Climbing and walking on Exit Glacier.

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... Glacier. 13.1312 Section 13.1312 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF... General Provisions § 13.1312 Climbing and walking on Exit Glacier. Except for areas designated by the Superintendent, climbing or walking on, in, or under Exit Glacier is prohibited within 1/2 mile of the glacial...

  15. 36 CFR 13.1312 - Climbing and walking on Exit Glacier.

    Code of Federal Regulations, 2013 CFR

    2013-07-01

    ... Glacier. 13.1312 Section 13.1312 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF... General Provisions § 13.1312 Climbing and walking on Exit Glacier. Except for areas designated by the Superintendent, climbing or walking on, in, or under Exit Glacier is prohibited within 1/2 mile of the glacial...

  16. 36 CFR 13.1312 - Climbing and walking on Exit Glacier.

    Code of Federal Regulations, 2011 CFR

    2011-07-01

    ... Glacier. 13.1312 Section 13.1312 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF... General Provisions § 13.1312 Climbing and walking on Exit Glacier. Except for areas designated by the Superintendent, climbing or walking on, in, or under Exit Glacier is prohibited within 1/2 mile of the glacial...

  17. The Glacier National Park GLORIA Project: A new US Target Region for Alpine Plant Monitoring Installed in the Northern Rocky Mountains, Montana

    NASA Astrophysics Data System (ADS)

    Holzer, K.; Fagre, D.

    2004-12-01

    The Global Observation Research Initiative in Alpine Environments (GLORIA) is an international research network whose purpose is to assess climate change impacts on vegetation in alpine environments worldwide. A standard protocol was developed by the international office in Vienna, Austria, and has specific site requirements and techniques that allow sites to be compared worldwide. This protocol requires four summits to be selected within a target region, covering zonal differences of subalpine to nival, and on each of these summits intensive vegetation plots are set up and monitored on a five year interval. Only three target regions in North America have been completed to date, one in Glacier National Park, Montana, and the other two in the Sierra Nevada and White Mountains, California. The four GLORIA summit plots in Glacier National Park were completed over the summers of 2003 and 2004. Because the Continental Divide bisects Glacier National Park (north to south), we chose summits only East of the divide to stay within a similar climatic pattern. Establishing sites was difficult due to the steep and rocky glaciated terrain and the remoteness of suitable sites that required multi-day approaches. Our highest summit (Seward Mtn. 2717 m) is the northernmost and our lowest summit (Dancing Lady Mtn. 2245 m) is southernmost. Treeline is strongly influenced by terrain and is significantly more variable than in the central Rocky Mountains. This also was true of zonal differences of alpine vegetation. Subalpine and even grassland species were found on the same summits as upper alpine species and areas considered subnival. While different zonal areas often occurred on one summit, they were highly influenced by the aspect and slope of that summit area. Between 51 and 82 vascular plants were documented on each summit. There was a high degree of variability in species diversity and percent cover on each summit that was correlated to directional exposure. The summit morphology

  18. Hydrologic conditions and hazards in the Kennicott River basin, Wrangell-St. Elias National Park Preserve, Alaska

    USGS Publications Warehouse

    Rickman, R.L.; Rosenkrans, D.S.

    1997-01-01

    McCarthy, Alaska, is on the Kennicott River, about 1 mile from the terminus of Kennicott Glacier in the Wrangell-St. Elias National Park and Preserve. Most visitors to McCarthy and the park cross the West Fork Kennicott River using a hand-pulled tram and cross the East Fork Kennicott River on a temporary footbridge. Outburst floods from glacier-dammed lakes result in channel erosion, aggradation, and migration of the Kennicott River, which disrupt transportation links, destroy property, and threaten life. Hidden Creek Lake, the largest of six glacier-dammed lakes in the Kennicott River Basin, has annual outbursts that cause the largest floods on the Kennicott River. Outbursts from Hidden Creek Lake occur from early fall to mid-summer, and lake levels at the onset of the outbursts have declined between 1909 and 1995. Criteria for impending outbursts for Hidden Creek Lake include lake stage near or above 3,000 to 3,020 feet, stationary or declining lake stage, evidence of recent calving of large ice blocks from the ice margin, slush ice and small icebergs stranded on the lakeshore, and fresh fractures in the ice-margin region. The lower Kennicott Glacier has thinned and retreated since about 1860. The East and West Fork Kennicott River channels migrated in response to changes in the lower Kennicott Glacier. The largest channel changes occur during outburst floods from Hidden Creek Lake, whereas channel changes from the other glacier-dammed lake outbursts are small. Each year, the West Fork Kennicott River conveys a larger percentage of the Kennicott Glacier drainage than it did the previous year. Outburst floods on the Kennicott River cause the river stage to rise over a period of several hours. Smaller spike peaks have a very rapid stage rise. Potential flood magnitude was estimated by combining known maximum discharges from Hidden Creek Lake and Lake Erie outburst floods with a theoretical large regional flood. Flood hazard areas at the transportation corridor were

  19. Motion of David Glacier in East Antarctica Observed by COSMO-SkyMed Differential SAR Interferometry

    NASA Astrophysics Data System (ADS)

    Han, H.; Lee, H.

    2011-12-01

    David glacier, located in Victoria Land, East Antarctica (75°20'S, 161°15'E), is an outlet glacier of 13 km width near the grounding line and 50 km long from the source to the grounding line. David glacier flows into Ross Sea forming Drygalski Ice Tongue, 100 km long and 23 km wide. In this study, we extracted a surface displacement map of David by applying differential SAR interferometry (DInSAR) to one-day tandem pairs obtained from COSMO-SkyMed satellites on April 28-29 (descending orbit) and May 5-6 (ascending orbit), 2011, respectively. Terra ASTER global digital elevation model (GDEM) is used to remove the topographic effect from the COSMO-SkyMed interferograms. David glacier showed maximum displacement of 35 cm during April 28-29 and 20 cm during May 5-6 in the direction of radar line of sight. The glacier can be divided into several blocks by the disparities of displacement between the different sliding zone. Surface displacement map contains errors originated from orbit data, atmospheric conditions, DEM error. GDEM is generated from the ASTER optical images acquired from 2000 to 2008. It has the vertical accuracy of about 20 m at 95% confidence with the 30 m of horizontal posting. The accuracy of GDEM reduces when cloud cover is included in the ASTER image. Particularly in the snow and ice area, GDEM is inaccurate due to whiteout effect during stereo matching. The inaccuracy of GDEM could be a reason of the observed glacier motion in the opposite direction of gravity. This problem can be solved by using TanDEM-X DEM. Bistatic acquisition of SAR images from the constellation of TerraSAR-X and TanDEM-X will generate a global DEM with the vertical accuracy better than 2 m and the horizontal posting of 12 m. We plan to perform DInSAR of COSMO-SkyMed one-day tandem pairs again when the high-accuracy TanDEM-X DEM is available in the near future. As a conclusion, we could analyze the displacement of David glacier in East Antarctica. The glacier showed very fast

  20. 36 CFR 13.1116 - Do I need a camping permit in Glacier Bay?

    Code of Federal Regulations, 2014 CFR

    2014-07-01

    ... 36 Parks, Forests, and Public Property 1 2014-07-01 2014-07-01 false Do I need a camping permit in Glacier Bay? 13.1116 Section 13.1116 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park...

  1. 36 CFR 13.1116 - Do I need a camping permit in Glacier Bay?

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... 36 Parks, Forests, and Public Property 1 2010-07-01 2010-07-01 false Do I need a camping permit in Glacier Bay? 13.1116 Section 13.1116 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park...

  2. 36 CFR 13.1116 - Do I need a camping permit in Glacier Bay?

    Code of Federal Regulations, 2013 CFR

    2013-07-01

    ... 36 Parks, Forests, and Public Property 1 2013-07-01 2013-07-01 false Do I need a camping permit in Glacier Bay? 13.1116 Section 13.1116 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park...

  3. 36 CFR 13.1134 - Who is eligible for a Glacier Bay commercial fishing lifetime access permit?

    Code of Federal Regulations, 2013 CFR

    2013-07-01

    ... Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1134 Who is eligible for a Glacier... 36 Parks, Forests, and Public Property 1 2013-07-01 2013-07-01 false Who is eligible for a Glacier...

  4. 36 CFR 13.1132 - What types of commercial fishing are authorized in Glacier Bay?

    Code of Federal Regulations, 2014 CFR

    2014-07-01

    ... fishing are authorized in Glacier Bay? 13.1132 Section 13.1132 Parks, Forests, and Public Property...-Glacier Bay National Park and Preserve Commercial Fishing § 13.1132 What types of commercial fishing are authorized in Glacier Bay? Three types of commercial fishing are authorized in Glacier Bay non-wilderness...

  5. 36 CFR 13.1132 - What types of commercial fishing are authorized in Glacier Bay?

    Code of Federal Regulations, 2011 CFR

    2011-07-01

    ... fishing are authorized in Glacier Bay? 13.1132 Section 13.1132 Parks, Forests, and Public Property...-Glacier Bay National Park and Preserve Commercial Fishing § 13.1132 What types of commercial fishing are authorized in Glacier Bay? Three types of commercial fishing are authorized in Glacier Bay non-wilderness...

  6. 36 CFR 13.1132 - What types of commercial fishing are authorized in Glacier Bay?

    Code of Federal Regulations, 2012 CFR

    2012-07-01

    ... fishing are authorized in Glacier Bay? 13.1132 Section 13.1132 Parks, Forests, and Public Property...-Glacier Bay National Park and Preserve Commercial Fishing § 13.1132 What types of commercial fishing are authorized in Glacier Bay? Three types of commercial fishing are authorized in Glacier Bay non-wilderness...

  7. Bacterial succession in Antarctic soils of two glacier forefields on Larsemann Hills, East Antarctica.

    PubMed

    Bajerski, Felizitas; Wagner, Dirk

    2013-07-01

    Antarctic glacier forefields are extreme environments and pioneer sites for ecological succession. Increasing temperatures due to global warming lead to enhanced deglaciation processes in cold-affected habitats, and new terrain is becoming exposed to soil formation and microbial colonization. However, only little is known about the impact of environmental changes on microbial communities and how they develop in connection to shifting habitat characteristics. In this study, using a combination of molecular and geochemical analysis, we determine the structure and development of bacterial communities depending on soil parameters in two different glacier forefields on Larsemann Hills, East Antarctica. Our results demonstrate that deglaciation-dependent habitat formation, resulting in a gradient in soil moisture, pH and conductivity, leads to an orderly bacterial succession for some groups, for example Cyanobacteria, Bacteroidetes and Deltaproteobacteria in a transect representing 'classical' glacier forefields. A variable bacterial distribution and different composed communities were revealed according to soil heterogeneity in a slightly 'matured' glacier forefield transect, where Gemmatimonadetes, Flavobacteria, Gamma- and Deltaproteobacteria occur depending on water availability and soil depth. Actinobacteria are dominant in both sites with dominance connected to certain trace elements in the glacier forefields. © 2013 Federation of European Microbiological Societies. Published by Blackwell Publishing Ltd. All rights reserved.

  8. Rock falls from Glacier Point above Camp Curry, Yosemite National Park, California

    USGS Publications Warehouse

    Wieczorek, Gerald F.; Snyder, James B.

    1999-01-01

    A series of rock falls from the north face of Glacier Point above Camp Curry, Yosemite National Park, California, have caused reexamination of the rock-fall hazard because beginning in June, 1999 a system of cracks propagated through a nearby rock mass outlining a future potential rock fall. If the estimated volume of the potential rock fall fails as a single piece, there could be a risk from rock-fall impact and airborne rock debris to cabins in Camp Curry. The role of joint plane orientation and groundwater pressure in the fractured rock mass are discussed in light of the pattern of developing cracks and potential modes of failure.

  9. Hazard assessment of the Tidal Inlet landslide and potential subsequent tsunami, Glacier Bay National Park, Alaska

    USGS Publications Warehouse

    Wieczorek, G.F.; Geist, E.L.; Motyka, R.J.; Jakob, M.

    2007-01-01

    An unstable rock slump, estimated at 5 to 10????????10 6 m3, lies perched above the northern shore of Tidal Inlet in Glacier Bay National Park, Alaska. This landslide mass has the potential to rapidly move into Tidal Inlet and generate large, long-period-impulse tsunami waves. Field and photographic examination revealed that the landslide moved between 1892 and 1919 after the retreat of the Little Ice Age glaciers from Tidal Inlet in 1890. Global positioning system measurements over a 2-year period show that the perched mass is presently moving at 3-4 cm annually indicating the landslide remains unstable. Numerical simulations of landslide-generated waves suggest that in the western arm of Glacier Bay, wave amplitudes would be greatest near the mouth of Tidal Inlet and slightly decrease with water depth according to Green's law. As a function of time, wave amplitude would be greatest within approximately 40 min of the landslide entering water, with significant wave activity continuing for potentially several hours. ?? 2007 Springer-Verlag.

  10. Post-breeding habitat use by adult boreal toads (Bufo boreas boreas) after wildfire in Glacier National Park, USA

    Treesearch

    C. Gregory Guscio; Blake R. Hossack; Lisa A. Eby; Paul Stephen Corn

    2008-01-01

    Effects of wildfire on amphibians are complex, and some species may benefit from the severe disturbance of stand-replacing fire. Boreal Toads (Bufo boreas boreas) in Glacier National Park, Montana, USA increased in occurrence after fires in 2001 and 2003. We used radio telemetry to track adult B. boreas in a mosaic of terrestrial...

  11. Tropical Glaciers

    NASA Astrophysics Data System (ADS)

    Fountain, Andrew

    The term "tropical glacier" calls to mind balmy nights and palm trees on one hand and cold, blue ice on the other. Certainly author Gabriel Garcia Marqez exploited this contrast in One Hundred Years of Solitude. We know that tropical fish live in warm, Sun-kissed waters and tropical plants provide lush, dense foliage populated by colorful tropical birds. So how do tropical glaciers fit into this scene? Like glaciers everywhere, tropical glaciers form where mass accumulation—usually winter snow—exceeds mass loss, which is generally summer melt. Thus, tropical glaciers exist at high elevations where precipitation can occur as snowfall exceeds melt and sublimation losses, such as the Rwenzori Mountains in east Africa and the Maoke Range of Irian Jaya.

  12. 36 CFR 13.1150 - Is a permit required for a vessel in Glacier Bay?

    Code of Federal Regulations, 2013 CFR

    2013-07-01

    ... 36 Parks, Forests, and Public Property 1 2013-07-01 2013-07-01 false Is a permit required for a vessel in Glacier Bay? 13.1150 Section 13.1150 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park...

  13. Decadal-scale climate drivers for glacial dynamics in Glacier National Park, Montana, USA

    USGS Publications Warehouse

    Pederson, G.T.; Fagre, D.B.; Gray, S.T.; Graumlich, L.J.

    2004-01-01

    Little Ice Age (14th-19th centuries A.D.) glacial maxima and 20th century retreat have been well documented in Glacier National Park, Montana, USA. However, the influence of regional and Pacific Basin driven climate variability on these events is poorly understood. We use tree-ring reconstructions of North Pacific surface temperature anomalies and summer drought as proxies for winter glacial accumulation and summer ablation, respectively, over the past three centuries. These records show that the 1850's glacial maximum was likely produced by ???70 yrs of cool/wet summers coupled with high snowpack. Post 1850, glacial retreat coincides with an extended period (>50 yr) of summer drought and low snowpack culminating in the exceptional events of 1917 to 1941 when retreat rates for some glaciers exceeded 100 m/yr. This research highlights potential local and ocean-based drivers of glacial dynamics, and difficulties in separating the effects of global climate change from regional expressions of decadal-scale climate variability. Copyright 2004 by the American Geophysical Union.

  14. East façade, Burton Park Club House, with Amphitheater in foreground, ...

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

    East façade, Burton Park Club House, with Amphitheater in foreground, view to north from Amphitheater stage (90 mm lens). - Burton Park, Club House & Amphitheater, Adjacent ot south end of Chestnut Avenue, San Carlos, San Mateo County, CA

  15. Inventory of marine and estuarine fishes in southeast and central Alaska National Parks

    USGS Publications Warehouse

    Arimitsu, Mayumi L.; Litzow, Michael A.; Piatt, John F.; Robards, Martin D.; Abookire, Alisa A.; Drew, Gary S.

    2003-01-01

    As part of a national inventory program funded by the National Park Service, we conducted an inventory of marine and estuarine fishes in Glacier Bay National Park and Preserve, Wrangell-St. Elias National Park and Preserve, Sitka National Historical Park, and Klondike Gold Rush National Historical Park in 2001 and 2002. In addition, marine fish data from a previous project that focused on forage fishes and marine predators during 1999 and 2000 in Glacier Bay proper were compiled for this study. Sampling was conducted with modified herring and Isaacs-Kidd midwater trawls, a plumb staff beam trawl, and beach seines. Species lists of relative abundance were generated for nearshore fishes in all parks, and for demersal and pelagic fishes in Glacier Bay National Park and Preserve and Wrangell-St. Elias National Park and Preserve. With a total sampling effort of 531 sets, we captured 100 species in Glacier Bay National Park and Preserve, 31 species in Wrangell-St. Elias National Park and Preserve, 23 species in Sitka National Historical Park, and 11 species in Klondike Gold Rush National Historical Park. We estimated that between 59 and 85 percent of the total marine fish species present were sampled by us in the various habitat-park units. We also combined these data with historical records and prepared an annotated species list of 160 marine and estuarine fishes known to occur in Glacier Bay National Park and Preserve. Shannon-Wiener diversity index and catch per unit effort were used to assess the effects of depth and latitude (distance from tidewater glaciers) on marine fish community ecology in Glacier Bay proper. Our findings suggest that demersal fishes are more abundant and diverse with increased distance from tidewater glaciers, and that pelagic fishes sampled deeper than 50 m are more abundant in areas closer to tidewater glaciers.

  16. Cartographic modeling of snow avalanche path location within Glacier National Park, Montana

    NASA Technical Reports Server (NTRS)

    Walsh, Stephen J.; Brown, Daniel G.; Bian, Ling; Butler, David R.

    1990-01-01

    Geographic information system (GIS) techniques were applied to the study of snow-avalanche path location within Glacier National Park, Montana. Aerial photointerpretation and field surveys confirmed the location of 121 avalanche paths within the selected study area. Spatial and nonspatial information on each path were integrated using the ARC/INFO GIS. Lithologic, structural, hydrographic, topographic, and land-cover impacts on path location were analyzed. All path frequencies within variable classes were normalized by the area of class occurrence relative to the total area of the study area and were added to the morphometric information contained within INFO tables. The normalized values for each GIS coverage were used to cartographically model, by means of composite factor weightings, avalanche path locations.

  17. Proceedings of the Fourth Glacier Bay Science Symposium

    USGS Publications Warehouse

    Piatt, John F.; Gende, Scott M.

    2007-01-01

    Foreword Glacier Bay was established as a National Monument in 1925, in part to protect its unique character and natural beauty, but also to create a natural laboratory to examine evolution of the glacial landscape. Today, Glacier Bay National Park and Preserve is still a place of profound natural beauty and dynamic landscapes. It also remains a focal point for scientific research and includes continuing observations begun decades ago of glacial processes and terrestrial ecosystems. In recent years, research has focused on glacial-marine interactions and ecosystem processes that occur below the surface of the bay. In October 2004, Glacier Bay National Park convened the fourth in a series of science symposiums to provide an opportunity for researchers, managers, interpreters, educators, students and the general public to share knowledge about Glacier Bay. The Fourth Glacier Bay Science Symposium was held in Juneau, Alaska, rather than at the Park, reflecting a desire to maximize attendance and communication among a growing and diverse number of stakeholders interested in science in the park. More than 400 people attended the symposium. Participants provided 46 oral presentations and 41 posters covering a wide array of disciplines including geology, glaciology, oceanography, wildlife and fisheries biology, terrestrial and marine ecology, socio-cultural research and management issues. A panel discussion focused on the importance of connectivity in Glacier Bay research, and keynote speakers (Gary Davis and Terry Chapin) spoke of long-term monitoring and ecological processes. These proceedings include 56 papers from the symposium. A summary of the Glacier Bay Science Plan-itself a subject of a meeting during the symposium and the result of ongoing discussions between scientists and resource managers-also is provided. We hope these proceedings illustrate the diversity of completed and ongoing scientific studies, conducted within the Park. To this end, we invited all

  18. The Bay in Place of a Glacier.

    ERIC Educational Resources Information Center

    Howell, Wayne

    1997-01-01

    The cultural resource specialist at Glacier Bay National Park (Alaska) explains the collaborative efforts of park staff and the Hoonah Tlingit to overcome language and cultural barriers in documenting park place names and clan oral history and traditions. The new park-community relationship, which follows decades of conflict, includes training…

  19. Grizzly bear density in Glacier National Park, Montana

    USGS Publications Warehouse

    Kendall, K.C.; Stetz, J.B.; Roon, David A.; Waits, L.P.; Boulanger, J.B.; Paetkau, David

    2008-01-01

    We present the first rigorous estimate of grizzly bear (Ursus arctos) population density and distribution in and around Glacier National Park (GNP), Montana, USA. We used genetic analysis to identify individual bears from hair samples collected via 2 concurrent sampling methods: 1) systematically distributed, baited, barbed-wire hair traps and 2) unbaited bear rub trees found along trails. We used Huggins closed mixture models in Program MARK to estimate total population size and developed a method to account for heterogeneity caused by unequal access to rub trees. We corrected our estimate for lack of geographic closure using a new method that utilizes information from radiocollared bears and the distribution of bears captured with DNA sampling. Adjusted for closure, the average number of grizzly bears in our study area was 240.7 (95% CI = 202–303) in 1998 and 240.6 (95% CI = 205–304) in 2000. Average grizzly bear density was 30 bears/1,000 km2, with 2.4 times more bears detected per hair trap inside than outside GNP. We provide baseline information important for managing one of the few remaining populations of grizzlies in the contiguous United States.

  20. 36 CFR 13.1134 - Who is eligible for a Glacier Bay commercial fishing lifetime access permit?

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... 36 Parks, Forests, and Public Property 1 2010-07-01 2010-07-01 false Who is eligible for a Glacier... Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1134 Who is eligible for a Glacier Bay commercial fishing lifetime access permit? A Glacier Bay commercial fishing lifetime access permit...

  1. 36 CFR 13.1134 - Who is eligible for a Glacier Bay commercial fishing lifetime access permit?

    Code of Federal Regulations, 2011 CFR

    2011-07-01

    ... 36 Parks, Forests, and Public Property 1 2011-07-01 2011-07-01 false Who is eligible for a Glacier... Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1134 Who is eligible for a Glacier Bay commercial fishing lifetime access permit? A Glacier Bay commercial fishing lifetime access permit...

  2. 36 CFR 13.1134 - Who is eligible for a Glacier Bay commercial fishing lifetime access permit?

    Code of Federal Regulations, 2014 CFR

    2014-07-01

    ... 36 Parks, Forests, and Public Property 1 2014-07-01 2014-07-01 false Who is eligible for a Glacier... Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1134 Who is eligible for a Glacier Bay commercial fishing lifetime access permit? A Glacier Bay commercial fishing lifetime access permit...

  3. 36 CFR 13.1134 - Who is eligible for a Glacier Bay commercial fishing lifetime access permit?

    Code of Federal Regulations, 2012 CFR

    2012-07-01

    ... 36 Parks, Forests, and Public Property 1 2012-07-01 2012-07-01 false Who is eligible for a Glacier... Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1134 Who is eligible for a Glacier Bay commercial fishing lifetime access permit? A Glacier Bay commercial fishing lifetime access permit...

  4. The length of the glaciers in the world - a straightforward method for the automated calculation of glacier center lines

    NASA Astrophysics Data System (ADS)

    Machguth, H.; Huss, M.

    2014-05-01

    Glacier length is an important measure of glacier geometry but global glacier inventories are mostly lacking length data. Only recently semi-automated approaches to measure glacier length have been developed and applied regionally. Here we present a first global assessment of glacier length using a fully automated method based on glacier surface slope, distance to the glacier margins and a set of trade-off functions. The method is developed for East Greenland, evaluated for the same area as well as for Alaska, and eventually applied to all ∼200 000 glaciers around the globe. The evaluation highlights accurately calculated glacier length where DEM quality is good (East Greenland) and limited precision on low quality DEMs (parts of Alaska). Measured length of very small glaciers is subject to a certain level of ambiguity. The global calculation shows that only about 1.5% of all glaciers are longer than 10 km with Bering Glacier (Alaska/Canada) being the longest glacier in the world at a length of 196 km. Based on model output we derive global and regional area-length scaling laws. Differences among regional scaling parameters appear to be related to characteristics of topography and glacier mass balance. The present study adds glacier length as a central parameter to global glacier inventories. Global and regional scaling laws might proof beneficial in conceptual glacier models.

  5. A cave in Glacier Grey in Torres del Paine National Park, seen during NASA's AirSAR 2004 campaign in Chile

    NASA Image and Video Library

    2004-03-11

    A cave in Glacier Grey in Torres del Paine National Park, seen during NASA's AirSAR 2004 campaign in Chile. AirSAR 2004 is a three-week expedition in Central and South America by an international team of scientists that is using an all-weather imaging tool, called the Airborne Synthetic Aperture Radar (AirSAR), located onboard NASA's DC-8 airborne laboratory. Scientists from many parts of the world are combining ground research with NASA's AirSAR technology to improve and expand on the quality of research they are able to conduct. Founded in 1959, Torres del Paine National Park encompasses 450,000 acres in the Patagonia region of Chile. This region is being studied by NASA using a DC-8 equipped with an Airborne Synthetic Aperture Radar (AirSAR) developed by scientists from NASA’s Jet Propulsion Laboratory. This is a very sensitive region that is important to scientists because the temperature has been consistently rising causing a subsequent melting of the region’s glaciers. AirSAR will provide a baseline model and unprecedented mapping of the region. This data will make it possible to determine whether the warming trend is slowing, continuing or accelerating. AirSAR will also provide reliable information on ice shelf thickness to measure the contribution of the glaciers to sea level.

  6. 1. EAST SIDE SHOWING LOW CINDERBLOCK WALL AND ASPHALTPAVED PARKING ...

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

    1. EAST SIDE SHOWING LOW CINDER-BLOCK WALL AND ASPHALT-PAVED PARKING LOT FOR NEW CONTROL BUILDING. VIEW TO NORTHWEST. - Bishop Creek Hydroelectric System, Control Station, Hydrographer's Office, Bishop Creek, Bishop, Inyo County, CA

  7. Source levels and call parameters of harbor seal breeding vocalizations near a terrestrial haulout site in Glacier Bay National Park and Preserve.

    PubMed

    Matthews, Leanna P; Parks, Susan E; Fournet, Michelle E H; Gabriele, Christine M; Womble, Jamie N; Klinck, Holger

    2017-03-01

    Source levels of harbor seal breeding vocalizations were estimated using a three-element planar hydrophone array near the Beardslee Islands in Glacier Bay National Park and Preserve, Alaska. The average source level for these calls was 144 dB RMS re 1 μPa at 1 m in the 40-500 Hz frequency band. Source level estimates ranged from 129 to 149 dB RMS re 1 μPa. Four call parameters, including minimum frequency, peak frequency, total duration, and pulse duration, were also measured. These measurements indicated that breeding vocalizations of harbor seals near the Beardslee Islands of Glacier Bay National Park are similar in duration (average total duration: 4.8 s, average pulse duration: 3.0 s) to previously reported values from other populations, but are 170-220 Hz lower in average minimum frequency (78 Hz).

  8. 90-year-old firn air from Styx glacier, East Antarctica

    NASA Astrophysics Data System (ADS)

    Jang, Y.; Ahn, J.; Buizert, C.; Lee, H. G.; Hong, S.; Han, Y.; Jun, S. J.; Hur, S. D.

    2017-12-01

    Firn is the upper part of the glacier that has not yet been completely changed to the ice. In this layer, firn air can move through the open pores and be pumped for sampling. We obtained firn air and ice cores from Styx glacier (73°51'95″ S, 163°41'217″ E, 1623m asl.), East Antarctica during 2014-2015. The Styx glacier is located near coast, and has an accumulation rate of 0.13 Mgm-2y-1 with a mean annual temperature of -31.7 °. We found that the lock-in depth (depth where gas diffusion starts to stop, "LID") is 52.4 m and bubble close-off depth (the depth to the snow-ice transition perfectly, "COD") is 65.1 m. Therefore lock-in zone (between LID and COD, "LIZ") is 52.4 - 65.1 m. Concentrations of greenhouse gases (CO2, CH4, n=13) in the firn air were analyzed at US National Oceanic and Atmospheric Administration (NOAA) and 15N of N2 was measured at the Scripps Institution of Oceanography (SIO). We find that the firn air ages are up to about 90 years, the oldest firn air ages observed among coastal glaciers. In order to better understand physical properties and chemical composition, methane concentration and total air content of the closed bubbles in the LIZ (3 cm resolution, n=124) were analyzed by a wet extraction method at Seoul National University. The CH4 concentration and total air content show large variations in cm-scale depth intervals, and they are anti-correlated with each other. The CH4 concentration changes in a few cm corresponds to up to 40 years in CH4 age. We also applied Centre for Ice and Climate (CIC) 1-dimensional diffusion model and simulated greenhouse gas concentration profiles to quantitatively understand how the air moves in the Styx firn column. We hypothesize that density variations in the firn may increase thickness of LIZ and consequently increase of firn gas ages.

  9. Climate Change and Glacier Retreat: Scientific Fact and Artistic Opportunity

    NASA Astrophysics Data System (ADS)

    Fagre, D. B.

    2008-12-01

    Mountain glaciers continue to retreat rapidly over most of the globe. In North America, at Glacier National Park, Montana, recent research results from Sperry Glacier (2005-2007) indicate negative mass balances are now 3-4 times greater than in the 1950s. A geospatial model of glacier retreat in the Blackfoot-Jackson basin suggested all glaciers would be gone by 2030 but has proved too conservative. Accelerated glacier shrinkage since the model was developed has mirrored an increase in actual annual temperature that is almost twice the rate used in the model. The glaciers in Glacier National Park are likely to be gone well before 2030. A variety of media, curricula, and educational strategies have been employed to communicate the disappearance of the glaciers as a consequence of global warming. These have included everything from print media and television coverage to podcasts and wayside exhibits along roads in the park. However, a new thrust is to partner with artists to communicate climate change issues to new audiences and through different channels. A scientist-artist retreat was convened to explore the tension between keeping artistic products grounded in factually-based reality while providing for freedom to express artistic creativity. Individual artists and scientists have worked to create aesthetic and emotional images, using painting, poetry, music and photography, to convey core messages from research on mountain ecosystems. Finally, a traveling art exhibit was developed to highlight the photography that systematically documents glacier change through time. The aim was to select photographs that provide the most compelling visual experience for an art-oriented viewer and also accurately reflect the research on glacier retreat. The exhibit opens on January 11, 2009

  10. Parks In Partnership.

    ERIC Educational Resources Information Center

    Bowman, Sally-Jo

    1998-01-01

    More than 50 National Park Service (NPS) sites interpret Native cultures or early Native contact with Europeans. In about 30 of those, American Indians, Alaska Natives, or Native Hawaiians, in partnership with the NPS, present their own heritage and issues. Describes Native-run aspects of Sitka National Historical Park, Glacier National Park, and…

  11. Sea otter studies in Glacier Bay National Park and Preserve: annual report 2002

    USGS Publications Warehouse

    Bodkin, James L.; Kloecker, Kimberly A.; Esslinger, George G.; Monson, Daniel H.; Coletti, Heather A.; Doherty, Janet

    2003-01-01

    will also trigger ecosystem level changes, as prey for other predators, such as octopus, sea stars, fishes, birds and mammals are modified. Sea otters will also modify benthic habitats through excavation of sediments required to extract burrowing infauna such as clams. Effects of sediment disturbance by foraging sea otters are not understood. Glacier Bay also supports large populations of other preferred sea otter prey, such as king (Paralithodes sp.), tanner (Chionoecetes sp.) and dungeness (Cancer magister) crabs and green sea urchins (S. droebachiensis). As the colonization of Park waters by sea otters continues, it is also likely that dramatic changes will occur in the species composition, abundance, and size class distribution of many components of the nearshore marine ecosystem. Many of the changes will occur as a direct result of predation by sea otters. Others will result from indirect or cascading effects of sea otter foraging, such as increased kelp production and modified prey availability for other nearshore predators. Without recognizing and quantifying the extent of change initiated by the colonization of Glacier Bay by sea otters, management of nearshore resources will be severely constrained for many decades.

  12. Analysis of a 24-Year photographic record of Nisqually glacier, Mount Rainier National Park, Washington

    USGS Publications Warehouse

    Veatch, Fred M.

    1969-01-01

    A systematic coverage of Nisqually Glacier by photographs taken from a network of stations on the ground was begun in 1942 to explore the value and limitations of such photographs as an aid in glacier study. Principles developed may be of value elsewhere, especially for the program 'Measurement of Glacier Variations on a World-Wide Basis' of the International Hydrological Decade. Nisqually Glacier in Mount Rainier National Park, Wash., covers 2.5 square miles (6.5 square kilometers) (1961) and extends from an altitude of about 14,300 feet (4,400 meters) near the top of Mount Rainier down to 4,700 feet (1,400 meters), in a horizontal distance of 4.1 miles (6.6 kilometers). Analyses were made of the annual photographs taken by the writer for 24 years from about 20 stations. A number of pictures taken sporadically from 1884 to 1941 by others were also available for use in the study. Where possible, the results obtained from photographs were compared with those from the available engineering surveys. Such detailed analysis of an extensive photographic coverage of a single glacier may be unique. Photographs illustrating the retreat and advance of the glacier's west ice margin in a reach extending for about a mile (1.6 kilometers) downstream from Wilson Glacier show that, by 1965, most of the ice thickness lost in that area between 1890 and 1944 had been recovered. Withering of the stagnant valley tongue down glacier from the nunatak is portrayed, as is its spectacular reactivation in the 1960's by a vigorous advance of fresh ice. Some of the visible characteristics of advancing and receding termini are noted. Annual values of the glacier's surface slope (5 to 10 degrees) at a cross profile were measured on photographs with respect to a projected vertical line identifiable in each picture. The results were found to average about 2 degrees less than those obtained from the 5-year topographic maps, but they are thought to be a little more accurate owing to lack of a

  13. 8. PARK AVENUE EAST OF CEDAR STREET (400 Block). THE ...

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

    8. PARK AVENUE EAST OF CEDAR STREET (400 Block). THE MARCHION HARDWARE BUILDING WAS DESIGNED BY W.W. HISLOP, AND BUILT IN 1895. THE GROUND FLOOR WAS RENOVATED SOME TIME IN THE 1930s. IN THE CENTER IS THE IMPERIAL BLOCK (ca. 1920), AND THE FULLER DRUG COMPANY (1918-1932).THE FULLER SITE WAS OCCUPIED BY THE HIGHLAND THEATER FROM 1932 TO 1972, AND RETAINS MUCH OF THE INTERIOR DECORATION FROM THAT PERIOD - Anaconda Historic District, Park & Commercial Streets, Main Street vicinity, Anaconda, Deer Lodge County, MT

  14. 36 CFR 13.1312 - Climbing and walking on Exit Glacier.

    Code of Federal Regulations, 2012 CFR

    2012-07-01

    ... 36 Parks, Forests, and Public Property 1 2012-07-01 2012-07-01 false Climbing and walking on Exit Glacier. 13.1312 Section 13.1312 Parks, Forests, and Public Property NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National Park...

  15. Khurdopin Glacier, Pakistan

    NASA Image and Video Library

    2018-03-26

    In October 2016, the Khurdopin Glacier in Pakistan began a rapid surge after 20 years of little movement. By March, 2017, a large lake had formed in the Shimshal River, where the glacier had formed a dam. Fortunately, the river carved an outlet through the glacier before the lake could empty catastrophically. In this pair of ASTER images, acquired August 20, 2015 and May 21, 2017, the advance of the Khurdopin Glacier (dark gray and white "river" in lower right quarter of image) is obvious by comparing the before and after images. The images cover an area of 25 by 27.8 km, and are located at 36.3 degrees north, 75.5 degrees east. https://photojournal.jpl.nasa.gov/catalog/PIA22304

  16. New Zealand Glaciers

    NASA Image and Video Library

    2017-03-09

    New Zealand contains over 3,000 glaciers, most of which are in the Southern Alps on the South Island. Since 1890, the glaciers have been retreating, with short periods of small advances, as shown in this image from NASA Terra spacecraft. The image cover an area of 39 by 46 km, and are located at 43.7 degrees south, 170 degrees east. http://photojournal.jpl.nasa.gov/catalog/PIA21509

  17. 4. DETAIL VIEW OF TRIPLE TONGUE AND GROOVE CRIBBING USED ...

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

    4. DETAIL VIEW OF TRIPLE TONGUE AND GROOVE CRIBBING USED IN DAM CONSTRUCTION, NORTH EAST OF EAST DAM, LOOKING NORTH - Three Bears Lake & Dams, East Dam, North of Marias Pass, East Glacier Park, Glacier County, MT

  18. Glacier-derived August runoff in northwest Montana

    USGS Publications Warehouse

    Clark, Adam; Harper, Joel T.; Fagre, Daniel B.

    2015-01-01

    The second largest concentration of glaciers in the U.S. Rocky Mountains is located in Glacier National Park (GNP), Montana. The total glacier-covered area in this region decreased by ∼35% over the past 50 years, which has raised substantial concern about the loss of the water derived from glaciers during the summer. We used an innovative weather station design to collect in situ measurements on five remote glaciers, which are used to parameterize a regional glacier melt model. This model offered a first-order estimate of the summer meltwater production by glaciers. We find, during the normally dry month of August, glaciers in the region produce approximately 25 × 106 m3 of potential runoff. We then estimated the glacier runoff component in five gaged streams sourced from GNP basins containing glaciers. Glacier-melt contributions range from 5% in a basin only 0.12% glacierized to >90% in a basin 28.5% glacierized. Glacier loss would likely lead to lower discharges and warmer temperatures in streams draining basins >20% glacier-covered. Lower flows could even be expected in streams draining basins as little as 1.4% glacierized if glaciers were to disappear.

  19. Nearshore distribution and abundance of Dungeness crabs in Glacier Bay National Park, Alaska

    USGS Publications Warehouse

    O'Clair, Charles E.; Freese, J. Lincoln; Stone, Robert P.; Shirley, Thomas C.; Leder, Erica H.; Taggart, S. James; Kruse, Gordon H.; Engstrom, Daniel R.

    1995-01-01

    As part of an ongoing, multi-agency study to determine the effects of closure of the commercial fishery for Dungeness crabs, Cancer magister, on crab population structure we examined patterns of distribution and abundance of crabs in nearshore habitats at five locations in and near Glacier Bay National Park. Sampling was conducted in April and September 1992 and April 1993 prior to the anticipated closure of the fishery in the park. Divers censused crabs by sex and reproductive state (ovigerous/nonovigerous females) along belt transects (2m x 100m) laid perpendicular to shore in the depth range 0 m (mean lower low water) to 18 m.Preliminary results from the first three sampling periods revealed that the average densities of Dungeness crabs at the five locations ranged from 78 to 2012 crabs/ha. Crab densities differed between populations depending on sex, reproductive state of females and sampling period. Male crabs showed reduced densities at Gustavus Flats in April 1992 (P<0.01) and 1993 (P<0.001). Ovigerous females had greater density at Bartlett Cove in April 1993 (P<0.001). Sex ratios were frequently skewed toward females. At Bartlett Cove and Gustavus Flats females outnumbered males in April 1992 and 1993 (P<0.001). Most of the females at Bartlett Cove and Gustavus Flats in April 1992 and 1993 were ovigerous (P-0.001). Males tended to occupy greater depths than females in April 1992 (P<0.05) but not April 1993 (P-005). The mean depth of males shifted from deeper to shallower water between April and September 1992 (P<0.001). The depth distribution of ovigerous crabs did not differ from that of nonovigerous female crabs. Future research prior to the anticipated closure of the commercial Dungeness crab fishery in Glacier Bay will include a tagging study to determine the extent of crab movement and further study of the temporal as well as the spatial variability observed in the structure of these populations.

  20. Stromatolites of the Belt Series in Glacier National Park and Vicinity, Montana

    USGS Publications Warehouse

    Rezak, Richard

    1957-01-01

    Eight zones of Precambrian stromatolites that are useful for local correlation are recognized in the Belt series of the Glacier National Park region, Montana. The zones vary in composition, thickness, and areal extent. Some are widespread and extend into neighboring regions, and others occur only in small areas. Their names are taken from the dominant species that occurs in each zone. The zones are, from youngest to oldest - Conophyton zone 2 Missoula group Collenia symmetrica zone 2 Collenia undosa zone 2 Collenia multiflabella zone Piegan group Conophyton zone 1 Collenia symmetrica zone 1 Collenia undosa zone 1 Ravalli group Collenia frequens zone Only the Conophyton zones have been mapped in the park area. The present study uses a classification based upon the three criteria of (1) mode of growth, (2) gross form of the colony, and (3) nature and orientation of the laminae. The scheme of classification also seems applicable to Paleozoic and later stromatolites. Possibly a consistent pattern of form-genera and form-species may be developed. Four form-genera and seven form-species are recognized in the Belt series of the park region. These are Cryptozoon occidentale Dawson, Collenia undosa Walcott, C. frequens Walcott, C. symmetrica Fenton and Fenton, Newlandia sp., and Conophyton inclinatum n. sp. It is realized that these structures should not be classified according to biological nomenclature. However, biological names are here applied to the structures until a suitable system of classification can be devised. Comparisons of the stromatolites of the Belt series with modern stromatolites on Andros Island, Bahama Islands, and Pleistocene stromatolites from Lake Lahonton, Nev., reveal similarities in structure that appear to be significant as to physical mode of origin.

  1. Marine benthic habitat mapping of the West Arm, Glacier Bay National Park and Preserve, Alaska

    USGS Publications Warehouse

    Hodson, Timothy O.; Cochrane, Guy R.; Powell, Ross D.

    2013-01-01

    Seafloor geology and potential benthic habitats were mapped in West Arm, Glacier Bay National Park and Preserve, Alaska, using multibeam sonar, groundtruthed observations, and geological interpretations. The West Arm of Glacier Bay is a recently deglaciated fjord system under the influence of glacial and paraglacial marine processes. High glacially derived sediment and meltwater fluxes, slope instabilities, and variable bathymetry result in a highly dynamic estuarine environment and benthic ecosystem. We characterize the fjord seafloor and potential benthic habitats using the recently developed Coastal and Marine Ecological Classification Standard (CMECS) by the National Oceanic and Atmospheric Administration (NOAA) and NatureServe. Due to the high flux of glacially sourced fines, mud is the dominant substrate within the West Arm. Water-column characteristics are addressed using a combination of CTD and circulation model results. We also present sediment accumulation data derived from differential bathymetry. These data show the West Arm is divided into two contrasting environments: a dynamic upper fjord and a relatively static lower fjord. The results of these analyses serve as a test of the CMECS classification scheme and as a baseline for ongoing and future mapping efforts and correlations between seafloor substrate, benthic habitats, and glacimarine processes.

  2. Diverse recreation experiences at Denali National Park and Preserve

    Treesearch

    Katie Knotek; Alan Watson; Neal Christensen

    2007-01-01

    Qualitative interviews were conducted at Denali National Park and Preserve in the 2004 summer use season to improve understanding of recreation visitor experiences in the remote southern portion of the park, including Mount McKinley and the surrounding mountains and glaciers. Descriptions of the experiences of visitors to the mountains and glaciers included elements of...

  3. Antarctic glacier-tongue velocities from Landsat images: First results

    USGS Publications Warehouse

    Lucchitta, Baerbel K.; Mullins, K.F.; Allison, A.L.; Ferrigno, Jane G.

    1993-01-01

    We measured the velocities of six glacier tongues and a few tongues within ice shelves distributed around the Antarctic coastline by determining the displacement of crevasse patterns seen on sequential Landsat images. The velocities range from less than 0.2 km a−1 for East Antarctic ice-shelf tongues to more than 2.5 km a−1 for the Thwaites Glacier Tongue. All glacier tongues show increases in velocity toward their distal margins. In general, the tongues of glaciers draining the West Antarctic ice sheet have moved significantly faster than those in East Antarctica. This observation may be significant in light of the hypothesized possible disintegration of the West Antarctic ice sheet.

  4. Changes in Greenland's peripheral glaciers linked to the North Atlantic Oscillation

    NASA Astrophysics Data System (ADS)

    Bjørk, A. A.; Aagaard, S.; Lütt, A.; Khan, S. A.; Box, J. E.; Kjeldsen, K. K.; Larsen, N. K.; Korsgaard, N. J.; Cappelen, J.; Colgan, W. T.; Machguth, H.; Andresen, C. S.; Peings, Y.; Kjær, K. H.

    2018-01-01

    Glaciers and ice caps peripheral to the main Greenland Ice Sheet contribute markedly to sea-level rise1-3. Their changes and variability, however, have been difficult to quantify on multi-decadal timescales due to an absence of long-term data4. Here, using historical aerial surveys, expedition photographs, spy satellite imagery and new remote-sensing products, we map glacier length fluctuations of approximately 350 peripheral glaciers and ice caps in East and West Greenland since 1890. Peripheral glaciers are found to have recently undergone a widespread and significant retreat at rates of 12.2 m per year and 16.6 m per year in East and West Greenland, respectively; these changes are exceeded in severity only by the early twentieth century post-Little-Ice-Age retreat. Regional changes in ice volume, as reflected by glacier length, are further shown to be related to changes in precipitation associated with the North Atlantic Oscillation (NAO), with a distinct east-west asymmetry; positive phases of the NAO increase accumulation, and thereby glacier growth, in the eastern periphery, whereas opposite effects are observed in the western periphery. Thus, with projected trends towards positive NAO in the future5,6, eastern peripheral glaciers may remain relatively stable, while western peripheral glaciers will continue to diminish.

  5. An Australian contribution to CryoSat-II cal/val in East Antarctica including the Totten glacier region

    NASA Astrophysics Data System (ADS)

    Watson, C. S.; Burgette, R. J.; Tregoning, P.; Coleman, R.; Roberts, J.; Lieser, J. L.; Fricker, H. A.; Legresy, B.

    2010-12-01

    The Australian TOT-Cal project seeks to provide a contribution to the calibration and validation of the CryoSat-II mission over two adjacent important regions in East Antarctica. The first focuses on the Totten glacier, arguably one of the most important outlet glaciers in the East Antarctic, known to be undergoing significant surface lowering. The second includes the coastal slope regions behind Casey station and up on the plateau areas near Law Dome where significant spatial variation in annual accumulation is known to occur. The 2010/11 austral summer is the first field season for this project, with fieldwork to be underway at the time of the AGU FM10. In this poster, we present our current field activities and forward plans for the 2011/12 season. Our field campaign includes three components. A total of six in-situ GPS sites will be deployed over the summer period throughout the Law Dome / Totten Glacier region. These sites will facilitate the computation of the integrated water vapour content of the atmosphere, enabling an assessment against the ECMWF product used in the CyroSat-II data stream. The GPS sites also serve to provide reference stations for the AWI Polar-5 aircraft that will fly over the study area equipped with a scanning LiDAR and the ESA ASIRAS instrument. Finally, a series of kinematic GPS transects, corner cube reflector placements and surface density measurements will be undertaken from our field camp on the western flank of Law Dome to provide high resolution ground measurements for cal/val activities. In a separate project, Antarctic sea ice freeboard measurements will also contribute to the calibration and validation efforts by the Australian Antarctic program. In November 2010, the first set of such measurements will be carried out in the East Antarctic sea ice zone between 77 and 90 degrees East. The primary measurement tools for this campaign will include helicopter mounted scanning LiDAR and aerial photography, combined with in

  6. Sustaining Change: The Struggle to Maintain Identity at Central Park East Secondary School

    ERIC Educational Resources Information Center

    Suiter, Diane

    2009-01-01

    Central Park East Secondary School (CPESS) in East Harlem was one of the most highly acclaimed and successful schools to come out of the period of school reform in the 1980s from which the Coalition of Essential Schools emerged. Noted progressive educator Deborah Meier founded CPESS in 1985 not as a reform model, but as a continuation of the…

  7. Geochemistry of the Johnson River, Lake Clark National Park and Preserve, Alaska

    USGS Publications Warehouse

    Brabets, Timothy P.; Riehle, James R.

    2003-01-01

    The Johnson River Basin, located in Lake Clark National Park and Preserve, drains an area of 96 square miles. A private inholding in the upper part of the basin contains a gold deposit that may be developed in the future. To establish a natural baseline to compare potential effects on water quality if development were to occur, the upper part of the Johnson River Basin was studied from 1999 to 2001 as part of a cooperative study with the National Park Service. Two basic rock types occur within the drainage basin of the study: the Jurassic Talkeetna Formation of interbedded volcanic and volcaniclastic rocks, and the slightly younger plutonic rocks of the Aleutian-Alaska Ranges batholith. The Johnson River gold prospect reflects widespread, secondary mineralization and alteration of the Talkeetna Formation. Metals found at the prospect proper are: arsenic, cadmium, copper, gold, iron, lead, mercury, molybdenum, selenium, silver, and zinc. The Johnson River prospect is located in the East Fork Ore Creek Basin, a 0.5 square mile watershed that is a tributary to the Johnson River. Water quality data from this stream reflect the mineralization of the basin and the highest concentrations of several trace elements and major ions of the water column were found in this stream. Presently, pH in this stream is normal, indicating that there is sufficient buffering capacity. At the Johnson River streamgage, which drains approximately 25 mi2 including the East Fork Ore Creek, concentrations of these constituents are significantly lower, reflecting the runoff from Johnson Glacier and Double Glacier, which account for approximately 75 percent of the total discharge. Streambed concentrations of cadmium, lead, and zinc from East Fork Ore Creek and its receiving stream, Ore Creek, typically exceed concentrations where sediment dwelling organisms would be affected. Similar to the water column chemistry, concentrations of these elements are lower at the Johnson River streamgage

  8. Sea otter studies in Glacier Bay National Park and Preserve: Aerial surveys, foraging observations, and intertidal clam sampling

    USGS Publications Warehouse

    Bodkin, James L.; Kloecker, Kimberly A.; Esslinger, George G.; Monson, Daniel H.; DeGroot, J.D.

    2001-01-01

    Althorp and 2 sites in Dundas Bay. There is no direct evidence of otter foraging at any of our clam sampling sites except at Port Althorp where sea otters have been present for >20 years and regularly forage intertidally. There is some indication of intertidal foraging in Idaho Inlet, based on reduced mean size of preferred clam species. Sea otters have been present in Idaho Inlet for at least 12 years. We sampled 48 systematically selected sites to allow inference throughout Glacier Bay intertidal areas and 12 preferred habitat intertidal sites to estimate maximum clam densities in the Bay. We also sampled 14 and 12 random sites in Idaho Inlet and Port Althorp, respectively, to provide contrast between sites with and without sea otters. Densities and biomass of intertidal clams were greater in the Lower Bay than either the East or West Arms. Mean densities (#/0.25m2) of all species of clams > 10.0 mm total length were 96.5 at preferred sites, 32.8 in the Lower Bay, 12.2 in the East Arm, 6.6 in the West Arm, 11.32 at Port Althorp and 27.1 at Idaho Inlet. Clam densities were lower in the Upper Arms of Glacier Bay, compared to the Lower Bay and were similar to densities at Port Althorp. In the Lower Bay, clam densities were nearly twice as high at preferred clam sites compared to those systematically sampled. Species of Macoma were the numerically dominant intertidal clam at most sites in Glacier Bay, while Protothaca staminea was dominant at Idaho Inlet and Port Althorp. Biomas (g/0.25m2) was higher in the Lower Bay (23.5) than either Arm (2.1 and .91) and higher at preferred sites (73.4) than systematically selected sites in Glacier Bay. Biomass estimates at Port Althorp were 5.2 and 9.7 at Idaho Inlet. Biomass estimates were dominated by species of Saxidomus, Protothaca and Mya in Glacier Bay and by Protothaca and Saxidomus at Idaho Inlet and Port Althrop. We suspect differences in density and biomass relate to habitat differences between areas within Glacier Bay

  9. 78 FR 43226 - Going-to-the-Sun Road Corridor Management Plan, Environmental Impact Statement, Glacier National...

    Federal Register 2010, 2011, 2012, 2013, 2014

    2013-07-19

    ...] Going-to-the-Sun Road Corridor Management Plan, Environmental Impact Statement, Glacier National Park... Statement for the Going-to-the-Sun Road Corridor Management Plan for Glacier National Park, Montana. This effort will result in an integrated visitor and transportation management plan for the Going-to-the Sun...

  10. 2. DETAIL VIEW SHOWING WOODEN CRIBBING WITH LOWERED LAKE LEVEL, ...

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

    2. DETAIL VIEW SHOWING WOODEN CRIBBING WITH LOWERED LAKE LEVEL, EAST DAM, LOOKING NORTHEAST (View is middle of the perimeter showing in MT-88-A-1 above.) - Three Bears Lake & Dams, East Dam, North of Marias Pass, East Glacier Park, Glacier County, MT

  11. Plant Functional Traits on Green Risers and Brown Treads of Periglacial Patterned Ground at Glacier National Park, Montana

    NASA Astrophysics Data System (ADS)

    Apple, M. E.; Ricketts, M. K.

    2016-12-01

    On the stair-stepped solifluction terraces of the periglacial patterned ground at Glacier National Park, Montana, the clearly visible striped pattern of green alternating with brown is formed by contrasts in the percent cover of plants with different functional traits. The sloping green risers dominated by the mat-forming dwarf shrubs, Dryas octopetela (Mountain Dryad) and Salix arctica (Arctic Willow) alternate with the relatively flat, sparsely covered brown rocky treads which are inhabitated by herbaceous, and often taprooted plants. Eleven species were restricted to the brown treads, including the rare arctic-alpine species Papaver pygmaeum (Pygmy Poppy), Aqiulegia jonesii (Jones' Columbine), Draba macounii, and Erigeron lanatus. Of these, the first three arise from taproots or branched rootcrowns. They are restricted to the brown rocky treads while E. lanatus arises from a caudex and grows on the treads and risers. The relative abundance of rare plants was significantly higher on the brown treads and no rare species were restricted to the green risers. The community weighted trait means were significantly higher for Raunkiaer cryptophytes and hemicryptophytes, graminoid, herbaceous and rosetted forms, and stolons, Underground traits varied significantly as well, since taproots, caudices, and other substantial roots had higher incidences on the brown treads than on the green risers. The brown, rocky treads are relatively flat with low percent plant cover and likely a water-stressed environment, hence the substantial investment in underground structures. In contrast, the sloped green risers are essentially covered by the mat-forming dwarf shrubs, D. octopetela and S. arctica, which augment their woody roots with the anchorage of adventitious roots and which provide shade and water retention for other plants, including seedlings of Abies lasiocarpa (Subalpine fir) and Pinus albicaulus (Whitebark Pine). Water from summer thunderstorms and seasonal melting supplies

  12. An elevational gradient in snowpack chemical loading at Glacier National Park, Montana: implications for ecosystem processes

    USGS Publications Warehouse

    Fagre, Daniel; Tonnessen, Kathy; Morris, Kristi; Ingersoll, George; McKeon, Lisa; Holzer, Karen

    2000-01-01

    The accumulation and melting of mountain snowpacks are major drivers of ecosystem processes in the Rocky Mountains. These include the influence of snow water equivalent (SWE) timing and amount of release on soil moisture for annual tree growth, and alpine stream discharge and temperature that control aquatic biota life histories. Snowfall also brings with it atmospheric deposition. Snowpacks will hold as much as 8 months of atmospheric deposition for release into mountain ecosystems during the spring melt. These pulses of chemicals influence soil microbiota and biogeochemical processes affecting mountain vegetation growth. Increased atmospheric nitrogen inputs recently have been documented in remote parts of Colorado's mountain systems but no baseline data exist for the Northern Rockies. We examined patterns of SWE and snow chemistry in an elevational gradient stretching from west to east over the continental divide in Glacier National Park in March 1999 and 2000. Sites ranged from 1080m to 2192m at Swiftcurrent Pass. At each site, two vertically-integrated columns of snow were sampled from snowpits up to 600cm deep and analyzed for major cations and anions. Minor differences in snow chemistry, on a volumetric basis, existed over the elvational gradient. Snowpack chemical loading estimates were calculated for NH4, SO4 and NO3 and closely followed elevational increases in SWE. NO3 (in microequivalents/square meter) ranged from 1,000 ueq/m2 at low elevation sites to 8,000+ ueq/m2 for high elevation sites. Western slopes received greater amounts of SWE and chemical loads for all tested compounds.

  13. Modeling and measuring snow for assessing climate change impacts in Glacier National Park, Montana

    USGS Publications Warehouse

    Fagre, Daniel B.; Selkowitz, David J.; Reardon, Blase; Holzer, Karen; Mckeon, Lisa L.

    2002-01-01

    A 12-year program of global change research at Glacier National Park by the U.S. Geological Survey and numerous collaborators has made progress in quantifying the role of snow as a driver of mountain ecosystem processes. Spatially extensive snow surveys during the annual accumulation/ablation cycle covered two mountain watersheds and approximately 1,000 km2 . Over 7,000 snow depth and snow water equivalent (SWE) measurements have been made through spring 2002. These augment two SNOTEL sites, 9 NRCS snow courses, and approximately 150 snow pit analyses. Snow data were used to establish spatially-explicit interannual variability in snowpack SWE. East of the Continental Divide, snowpack SWE was lower but also less variable than west of the Divide. Analysis of snowpacks suggest downward trends in SWE, a reduction in snow cover duration, and earlier melt-out dates during the past 52 years. Concurrently, high elevation forests and treelines have responded with increased growth. However, the 80 year record of snow from 3 NRCS snow courses reflects a strong influence from the Pacific Decadal Oscillation, resulting in 20-30 year phases of greater or lesser mean SWE. Coupled with the fine-resolution spatial snow data from the two watersheds, the ecological consequences of changes in snowpack can be empirically assessed at a habitat patch scale. This will be required because snow distribution models have had varied success in simulating snowpack accumulation/ablation dynamics in these mountain watersheds, ranging from R2=0.38 for individual south-facing forested snow survey routes to R2=0.95 when aggregated to the watershed scale. Key ecological responses to snowpack changes occur below the watershed scale, such as snow-mediated expansion of forest into subalpine meadows, making continued spatially-explicit snow surveys a necessity. 

  14. Glacier fluctuations in the Kenai Fjords, Alaska, U.S.A.: An evaluation of controls on Iceberg-calving glaciers

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

    Wiles, G.C.; Calkin, P.E.; Post, A.

    The histories of four iceberg-calving outlet-glacier systems in the Kenai Fjords National Park underscore the importance of fiord depth, sediment supply, and fiord geometry on glacier stability. These parameters, in turn, limit the reliability of calving glacier chronologies as records of climatic change. Tree-ring analysis together with radiocarbon dating show that the Northwestern and McCarty glaciers, with large drainage basins, were advancing in concert with nearby land-terminating glaciers about A.D. 600. After an interval of retreat and possible nonclimatically induced extension during the Medieval Warm Period, these ice margins advanced again through the Little Ice Age and then retreated synchronouslymore » with the surrounding land-terminating glaciers about A.D. 1900. In contrast, Holgate and Aialik glaciers, with deeper fiords and smaller basins, retreated about 300 yr earlier. Reconstructions of Little Ice Age glaciers suggest that equilibrium-line altitudes of Northwestern and McCarty glaciers were, respectively, 270 and 500 m lower than now. Furthermore, the reconstructions show that these two glaciers were climatically sensitive when at their terminal moranies. However, with ice margins at their present recessional positions and accumulation area ratios between 0.8 and 0.9, only McCarty Glacier shows evidence of advance. Aialik and Holgate glaciers were climatically insensitive during the Little Ice Age maxima and remain insensitive to climate. 40 refs., 7 figs., 2 tabs.« less

  15. The length of the world's glaciers - a new approach for the global calculation of center lines

    NASA Astrophysics Data System (ADS)

    Machguth, H.; Huss, M.

    2014-09-01

    Glacier length is an important measure of glacier geometry. Nevertheless, global glacier inventories are mostly lacking length data. Only recently semi-automated approaches to measure glacier length have been developed and applied regionally. Here we present a first global assessment of glacier length using an automated method that relies on glacier surface slope, distance to the glacier margins and a set of trade-off functions. The method is developed for East Greenland, evaluated for East Greenland as well as for Alaska and eventually applied to all ~ 200 000 glaciers around the globe. The evaluation highlights accurately calculated glacier length where digital elevation model (DEM) quality is high (East Greenland) and limited accuracy on low-quality DEMs (parts of Alaska). Measured length of very small glaciers is subject to a certain level of ambiguity. The global calculation shows that only about 1.5% of all glaciers are longer than 10 km, with Bering Glacier (Alaska/Canada) being the longest glacier in the world at a length of 196 km. Based on the output of our algorithm we derive global and regional area-length scaling laws. Differences among regional scaling parameters appear to be related to characteristics of topography and glacier mass balance. The present study adds glacier length as a key parameter to global glacier inventories. Global and regional scaling laws might prove beneficial in conceptual glacier models.

  16. Effects of prescribed burning on vegetation and fuel loading in three east Texas state parks

    Treesearch

    Sandra Rideout; Brian P. Oswald

    2002-01-01

    This study was conducted to evaluate the initial effectiveness of prescribed burning in the ecological restoration of forests within selected parks in east Texas. Twenty-four permanent plots were installed to monitor fuel loads, overstory, sapling, seedling, shrub and herbaceous layers within burn and control units of Mission Tejas, Tyler and Village Creek state parks...

  17. Combination of UAV and terrestrial photogrammetry to assess rapid glacier evolution and map glacier hazards

    NASA Astrophysics Data System (ADS)

    Fugazza, Davide; Scaioni, Marco; Corti, Manuel; D'Agata, Carlo; Azzoni, Roberto Sergio; Cernuschi, Massimo; Smiraglia, Claudio; Diolaiuti, Guglielmina Adele

    2018-04-01

    Tourists and hikers visiting glaciers all year round face hazards such as sudden terminus collapses, typical of such a dynamically evolving environment. In this study, we analyzed the potential of different survey techniques to analyze hazards of the Forni Glacier, an important geosite located in Stelvio Park (Italian Alps). We carried out surveys in the 2016 ablation season and compared point clouds generated from an unmanned aerial vehicle (UAV) survey, close-range photogrammetry and terrestrial laser scanning (TLS). To investigate the evolution of glacier hazards and evaluate the glacier thinning rate, we also used UAV data collected in 2014 and a digital elevation model (DEM) created from an aerial photogrammetric survey of 2007. We found that the integration between terrestrial and UAV photogrammetry is ideal for mapping hazards related to the glacier collapse, while TLS is affected by occlusions and is logistically complex in glacial terrain. Photogrammetric techniques can therefore replace TLS for glacier studies and UAV-based DEMs hold potential for becoming a standard tool in the investigation of glacier thickness changes. Based on our data sets, an increase in the size of collapses was found over the study period, and the glacier thinning rates went from 4.55 ± 0.24 m a-1 between 2007 and 2014 to 5.20 ± 1.11 m a-1 between 2014 and 2016.

  18. Glaciers of North America - Glaciers of Alaska

    USGS Publications Warehouse

    Molnia, Bruce F.

    2008-01-01

    Glaciers cover about 75,000 km2 of Alaska, about 5 percent of the State. The glaciers are situated on 11 mountain ranges, 1 large island, an island chain, and 1 archipelago and range in elevation from more than 6,000 m to below sea level. Alaska's glaciers extend geographically from the far southeast at lat 55 deg 19'N., long 130 deg 05'W., about 100 kilometers east of Ketchikan, to the far southwest at Kiska Island at lat 52 deg 05'N., long 177 deg 35'E., in the Aleutian Islands, and as far north as lat 69 deg 20'N., long 143 deg 45'W., in the Brooks Range. During the 'Little Ice Age', Alaska's glaciers expanded significantly. The total area and volume of glaciers in Alaska continue to decrease, as they have been doing since the 18th century. Of the 153 1:250,000-scale topographic maps that cover the State of Alaska, 63 sheets show glaciers. Although the number of extant glaciers has never been systematically counted and is thus unknown, the total probably is greater than 100,000. Only about 600 glaciers (about 1 percent) have been officially named by the U.S. Board on Geographic Names (BGN). There are about 60 active and former tidewater glaciers in Alaska. Within the glacierized mountain ranges of southeastern Alaska and western Canada, 205 glaciers (75 percent in Alaska) have a history of surging. In the same region, at least 53 present and 7 former large ice-dammed lakes have produced jokulhlaups (glacier-outburst floods). Ice-capped volcanoes on mainland Alaska and in the Aleutian Islands have a potential for jokulhlaups caused by subglacier volcanic and geothermal activity. Because of the size of the area covered by glaciers and the lack of large-scale maps of the glacierized areas, satellite imagery and other satellite remote-sensing data are the only practical means of monitoring regional changes in the area and volume of Alaska's glaciers in response to short- and long-term changes in the maritime and continental climates of the State. A review of the

  19. Malaspina Glacier, Alaska

    NASA Image and Video Library

    2017-12-08

    The ice of a piedmont glacier spills from a steep valley onto a relatively flat plain, where it spreads out unconstrained like pancake batter. Elephant Foot Glacier in northeastern Greenland is an excellent example; it is particularly noted for its symmetry. But the largest piedmont glacier in North America (and possibly the world) is Malaspina in southeastern Alaska. On September 24, 2014, the Operational Land Imager (OLI) on Landsat 8 acquired this image of Malaspina Glacier. The main source of ice comes from Seward Glacier, located at the top-center of this image. The Agassiz and Libbey glaciers are visible on the left side, and the Hayden and Marvine glaciers are on the right. The brown lines on the ice are moraines—areas where soil, rock, and other debris have been scraped up by the glacier and deposited at its sides. Where two glaciers flow together, the moraines merge to form a medial moraine. Glaciers that flow at a steady speed tend to have moraines that are relatively straight. But what causes the dizzying pattern of curves, zigzags, and loops of Malaspina’s moraines? Glaciers in this area of Alaska periodically “surge,”meaning they lurch forward quickly for one to several years. As a result of this irregular flow, the moraines at the edges and between glaciers can become folded, compressed, and sheared to form the characteristic loops seen on Malaspina. For instance, a surge in 1986 displaced moraines on the east side of Malaspina by as much as 5 kilometers (3 miles). NASA Earth Observatory image by Jesse Allen, using Landsat data from the U.S. Geological Survey. Caption by Kathryn Hansen. Credit: NASA Earth Observatory 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

  20. Runoff generation from neighboring headwater basins with differing glacier coverage using the distributed hydrological model WaSiM, Eklutna, Alaska

    NASA Astrophysics Data System (ADS)

    Ostman, J. S.; Loso, M.; Liljedahl, A. K.; Gaedeke, A.; Geck, J. E.

    2017-12-01

    Many Alaska glaciers are thinning and retreating, and glacier wastage is projected to affect runoff processes from glacierized basins. Accordingly, effective resource management in glacierized watersheds requires quantification of a glacier's role on streamflow generation. The Eklutna catchment (311 km2) supplies water and electricity for Anchorage, Alaska (pop. 300,000) via Eklutna Lake. The Eklutna headwaters include the West Fork (64 km2, 46% glacier), and the East Fork (101 km2, 12% glacier). Total average annual discharge (2009-2015) is similar from the West (42,100 m3) and East (42,200 m3) forks, while specific annual runoff from the West Fork (2940 mm) exceeds that of the East Fork (1500 mm). To better understand what controls runoff, we are simulating the Eklutna annual water budget using a distributed watershed-level hydrological model. We force the Water Flow and Balance Simulation Model (WaSiM) using continuous air temperature, precipitation, wind speed, shortwave incoming radiation, and relative humidity primarily measured in the West Fork basin. We use Eklutna Glacier snow accumulation and ablation to calibrate the snowmelt and glacier sub-modules. Melt season discharge from the West and East forks is used for runoff comparison. Preliminary results show 2013-2015 simulated glacier point balances (accumulation and melt) are within 15% of glacier stake observations. Runoff was effectively modeled in the West Fork (NSE=0.80), while being over-predicted in the East Fork , which we attribute to a lack of forcing data in the less-glacierized basin. The simulations suggest that 78% of West Fork total runoff is from glacier melt, compared with <40% in the East Fork where glacier runoff contribution is higher during low-snow years.

  1. Changes in glacier dynamics in the northern Antarctic Peninsula since 1985

    NASA Astrophysics Data System (ADS)

    Seehaus, Thorsten; Cook, Alison J.; Silva, Aline B.; Braun, Matthias

    2018-02-01

    The climatic conditions along the northern Antarctic Peninsula have shown significant changes within the last 50 years. Here we present a comprehensive analysis of temporally and spatially detailed observations of the changes in ice dynamics along both the east and west coastlines of the northern Antarctic Peninsula. Temporal evolutions of glacier area (1985-2015) and ice surface velocity (1992-2014) are derived from a broad multi-mission remote sensing database for 74 glacier basins on the northern Antarctic Peninsula ( < 65° S along the west coast and north of the Seal Nunataks on the east coast). A recession of the glaciers by 238.81 km2 is found for the period 1985-2015, of which the glaciers affected by ice shelf disintegration showed the largest retreat by 208.59 km2. Glaciers on the east coast north of the former Prince Gustav Ice Shelf extent in 1986 receded by only 21.07 km2 (1985-2015) and decelerated by about 58 % on average (1992-2014). A dramatic acceleration after ice shelf disintegration with a subsequent deceleration is observed at most former ice shelf tributaries on the east coast, combined with a significant frontal retreat. In 2014, the flow speed of the former ice shelf tributaries was 26 % higher than before 1996. Along the west coast the average flow speeds of the glaciers increased by 41 %. However, the glaciers on the western Antarctic Peninsula revealed a strong spatial variability of the changes in ice dynamics. By applying a hierarchical cluster analysis, we show that this is associated with the geometric parameters of the individual glacier basins (hypsometric indexes, maximum surface elevation of the basin, flux gate to catchment size ratio). The heterogeneous spatial pattern of ice dynamic evolutions at the northern Antarctic Peninsula shows that temporally and spatially detailed observations as well as further monitoring are necessary to fully understand glacier change in regions with such strong topographic and climatic variances.

  2. Spatial Pattern Analysis of Cruise Ship-Humpback Whale Interactions in and Near Glacier Bay National Park, Alaska

    NASA Astrophysics Data System (ADS)

    Harris, Karin; Gende, Scott M.; Logsdon, Miles G.; Klinger, Terrie

    2012-01-01

    Understanding interactions between large ships and large whales is important to estimate risks posed to whales by ships. The coastal waters of Alaska are a summer feeding area for humpback whales ( Megaptera novaeangliae) as well as a prominent destination for large cruise ships. Lethal collisions between cruise ships and humpback whales have occurred throughout Alaska, including in Glacier Bay National Park (GBNP). Although the National Park Service (NPS) establishes quotas and operating requirements for cruise ships within GBNP in part to minimize ship-whale collisions, no study has quantified ship-whale interactions in the park or in state waters where ship traffic is unregulated. In 2008 and 2009, an observer was placed on ships during 49 different cruises that included entry into GBNP to record distance and bearing of whales that surfaced within 1 km of the ship's bow. A relative coordinate system was developed in ArcGIS to model the frequency of whale surface events using kernel density. A total of 514 whale surface events were recorded. Although ship-whale interactions were common within GBNP, whales frequently surfaced in front of the bow in waters immediately adjacent to the park (west Icy Strait) where cruise ship traffic is not regulated by the NPS. When ships transited at speeds >13 knots, whales frequently surfaced closer to the ship's midline and ship's bow in contrast to speeds slower than 13 knots. Our findings confirm that ship speed is an effective mitigation measure for protecting whales and should be applied to other areas where ship-whale interactions are common.

  3. Spatial pattern analysis of cruise ship-humpback whale interactions in and near Glacier Bay National Park, Alaska.

    PubMed

    Harris, Karin; Gende, Scott M; Logsdon, Miles G; Klinger, Terrie

    2012-01-01

    Understanding interactions between large ships and large whales is important to estimate risks posed to whales by ships. The coastal waters of Alaska are a summer feeding area for humpback whales (Megaptera novaeangliae) as well as a prominent destination for large cruise ships. Lethal collisions between cruise ships and humpback whales have occurred throughout Alaska, including in Glacier Bay National Park (GBNP). Although the National Park Service (NPS) establishes quotas and operating requirements for cruise ships within GBNP in part to minimize ship-whale collisions, no study has quantified ship-whale interactions in the park or in state waters where ship traffic is unregulated. In 2008 and 2009, an observer was placed on ships during 49 different cruises that included entry into GBNP to record distance and bearing of whales that surfaced within 1 km of the ship's bow. A relative coordinate system was developed in ArcGIS to model the frequency of whale surface events using kernel density. A total of 514 whale surface events were recorded. Although ship-whale interactions were common within GBNP, whales frequently surfaced in front of the bow in waters immediately adjacent to the park (west Icy Strait) where cruise ship traffic is not regulated by the NPS. When ships transited at speeds >13 knots, whales frequently surfaced closer to the ship's midline and ship's bow in contrast to speeds slower than 13 knots. Our findings confirm that ship speed is an effective mitigation measure for protecting whales and should be applied to other areas where ship-whale interactions are common.

  4. 36 CFR 13.1132 - What types of commercial fishing are authorized in Glacier Bay?

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1132 What types of commercial fishing are... 36 Parks, Forests, and Public Property 1 2010-07-01 2010-07-01 false What types of commercial...

  5. 36 CFR 13.1132 - What types of commercial fishing are authorized in Glacier Bay?

    Code of Federal Regulations, 2013 CFR

    2013-07-01

    ... NATIONAL PARK SERVICE, DEPARTMENT OF THE INTERIOR NATIONAL PARK SYSTEM UNITS IN ALASKA Special Regulations-Glacier Bay National Park and Preserve Commercial Fishing § 13.1132 What types of commercial fishing are... 36 Parks, Forests, and Public Property 1 2013-07-01 2013-07-01 false What types of commercial...

  6. Sea otter studies in Glacier Bay National Park and Preserve

    USGS Publications Warehouse

    Bodkin, James L.; Kloecker, Kimberly A.; Esslinger, George G.; Monson, Daniel H.; DeGroot, J.D.; Doherty, J.

    2002-01-01

    Following translocations to the outer coast of Southeast Alaska in 1965, sea otters have been expanding their range and increasing in abundance. We began conducting surveys for sea otters in Cross Sound, Icy Strait, and Glacier Bay, Alaska in 1994, following initial reports (in 1993) of their presence in Glacier Bay. Since 1995, the number of sea otters in Glacier Bay proper has increased from around 5 to more than 1500. Between 1993 and 1997 sea otters were apparently only occasional visitors to Glacier Bay, but in 1998 long-term residence was established as indicated by the presence of adult females and their dependent pups. Sea otter distribution is limited to the Lower Bay, south of Sandy Cove, and is not continuous within that area. Concentrations occur in the vicinity of Sita Reef and Boulder Island and between Pt. Carolus and Rush Pt. on the west side of the Bay (Figure 1). We describe the diet of sea otters during 2001 in Glacier Bay based on visual observations of prey during 456 successful forage dives. In Glacier Bay, diet consisted of 62% clam, 15% mussel, 9% crab, 7% unidentified, 4& urchins, and 4% other. Most prey recovered by sea otters are commercially, socially, or ecologically important species. Species of clam include Saxidomus gigantea, Protothaca staminea, and Mya truncata. Urchins are primarily Strongylocentrotus droebachiensis and the mussel is Modiolus modiolus. Crabs include species of three genera: Cancer, Chinoecetes, and Telmessus. Although we characterize diet at broad geographic scales, we found diet to vary between sites separated by as little as several hundred meters. Dietary variation among and within sites can reflect differences in prey availability and individual specialization. We estimated species composition, density, biomass, and sizes of subtidal clams, urchins, and mussels at 9 sites in lower Glacier Bay. All sites were selected based on the presence of abundant clam siphons. Sites were not selected to allow inference to

  7. 13. VIEW SHOWING MOST OF THE PERIMETER FROM SPILLWAY BOX ...

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

    13. VIEW SHOWING MOST OF THE PERIMETER FROM SPILLWAY BOX TO END OF EAST DAM. FOREGROUND VIEW SHOWS TRIPLE WALL CONSTRUCTION OF TONGUE AND GROOVE PLANKING USED IN CRIBBING - Three Bears Lake & Dams, North of Marias Pass, East Glacier Park, Glacier County, MT

  8. 36 CFR 7.3 - Glacier National Park.

    Code of Federal Regulations, 2010 CFR

    2010-07-01

    ... regulations, based on management objectives described in the park's Resource Management Plan, are established... governing the Park. 2. Any building or structure used for the purpose of conducting the business herein... the business herein permitted as required by the Superintendent. 4. Permittee, his agents, and...

  9. Inferring Past Climate in Equatorial East Africa using Glacier Models

    NASA Astrophysics Data System (ADS)

    Doughty, A. M.; Kelly, M. A.; Anderson, B.; Russell, J. M.; Jackson, M. S.

    2016-12-01

    Mountain glaciers in the northern and southern middle latitudes advanced nearly synchronously during the Last Glacial Maximum (LGM), but the timing and magnitude of cooling is less certain for the tropics. Knowing the degree of cooling in high altitude, low latitude regions advances our understanding of the cryosphere in understudied areas and contributes to our understanding of what causes ice ages. Here we use a 2-D ice flow and mass balance model to simulate glacier extents in the Rwenzori Mountains of Uganda and the Democratic Republic of the Congo during the Last Glacial Maximum. In particular, we model steady-state ice extent that matches the dated moraines in the Rwenzori Mountains to infer past climate. Steady-state simulations of LGM glacier extents, which match moraines dated to 20,000 years ago, can be obtained with a 20% reduction in precipitation and a 7°C cooling to match the associated moraines. A 0-50% reduction in precipitation combined with a 5-8°C cooling, respectively, agrees well with paleoclimate estimates from independent proxy records. As expected in a high precipitation environment, these glaciers are very sensitive to decreases in temperature, converting large volumes of precipitation from rain to snow as well as decreasing melting. Glaciers in equatorial Africa appear to have been waxing and waning synchronously and by the same magnitude as glaciers in the middle latitudes, suggesting a common, global forcing mechanism.

  10. Changes in water properties and flow regime on the continental shelf off the Adélie/George V Land coast, East Antarctica, after glacier tongue calving

    NASA Astrophysics Data System (ADS)

    Aoki, S.; Kobayashi, R.; Rintoul, S. R.; Tamura, T.; Kusahara, K.

    2017-08-01

    Oceanic changes before and after the relocation of iceberg B9B and calving of the Mertz Glacier Tongue (MGT) in February 2010 are examined on the continental shelf off the Adélie Land/George V Land coast, East Antarctica. Summer hydrographic observations, including stable oxygen isotope ratio (δ18O), in 2001/2008 and 2011/2015 and results of a numerical model are used. Along the western flank of the MGT, temperature decreased between 2001 and 2015 for most of the water column in the Adélie Depression. δ18O generally decreased, especially at the MGT draft depths on the northern side. West of the MGT, temperature, salinity, and δ18O decreased in the intermediate layer. East of the MGT, in contrast, temperature increased between 2001 and 2011 at intermediate depths, salinity increased in the intermediate and deep layers, and δ18O slightly decreased in the deep layer but did not change much around 300 dbar. The numerical experiment exhibits a change in ocean circulation, revealing an increase in modified Circumpolar Deep Water (mCDW) inflow in the east and a decrease in the west. The contrasting changes in mCDW intrusion are consistent between the observations and numerical model, and are indicative of the effect of removal of the ice barriers. The contrast is overlain by overall decreases in salinity and δ18O, which suggests an increase in the continental meltwater fraction of 5-20% and might reveal a wide-ranging influence from West Antarctica. The oxygen isotope ratio is, hence, effective in monitoring the increase in continental melt over the Antarctic shelf.Plain Language SummaryAntarctic <span class="hlt">glaciers</span>, icebergs, and ice sheet have significant impact on the surrounding ocean, and, in turn, are affected by the ocean. The Mertz <span class="hlt">Glacier</span>, <span class="hlt">East</span> Antarctica, had been melted from below by the oceanic heat. The seaward extension of the <span class="hlt">glacier</span> of about 500 m tall obstructed sea ice drift from the <span class="hlt">east</span> and enabled a large</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70192252','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70192252"><span>Thermal tolerance of meltwater stonefly Lednia tumana nymphs from an alpine stream in Waterton–<span class="hlt">Glacier</span> International Peace <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Billman, Hilary G.; Giersch, J. Joseph; Kappenman, K.M.; Muhlfeld, Clint C.; Webb, Molly A. H.</p> <p>2013-01-01</p> <p>Global climate change threatens to affect negatively the structure, function, and diversity of aquatic ecosystems worldwide. In alpine systems, the thermal tolerances of stream invertebrates can be assessed to understand better the potential effects of rising ambient temperatures and continued loss of <span class="hlt">glaciers</span> and snowpack on alpine stream ecosystems. We measured the critical thermal maximum (CTM) and lethal temperature maximum (LTM) of the meltwater stonefly (Lednia tumana), a species limited to glacial and snowmelt-driven alpine streams in the Waterton–<span class="hlt">Glacier</span> International Peace <span class="hlt">Park</span> area and a candidate for listing under the US Endangered Species Act. We collected L. tumana nymphs from Lunch Creek in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana (USA) and transported them to a laboratory at the University of Montana Flathead Lake Biological Station, Polson, Montana. We placed nymphs in a controlled water bath at 1 of 2 acclimation temperatures, 8.5 and 15°C. We increased water temperature at a constant rate of 0.3°C/min. We calculated the average CTM and LTM (± SD) for each acclimation temperature and compared them with Student’s t-tests. Predicted chronic temperature maxima were determined using the ⅓ rule. Mean LTMs were 32.3 ± 0.28°C and 31.05 ± 0.78°C in the 8.5 and 15°C acclimation treatments, respectively. CTM and LTM metrics were lower in the 15 than in the 8.5°C acclimation treatment, but these differences were not statistically significant (p > 0.05). The predicted chronic temperature maxima were 20.6 and 20.2°C for the 8.5 and 15°C acclimation treatments, respectively. More research is needed on the effects of chronic exposures to rising stream temperatures, but our results can be used to assess the potential effects of warming water temperatures on L. tumana and other aquatic macroinvertebrates in alpine ecosystems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mt0305.photos.345042p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mt0305.photos.345042p/"><span>9. DETAIL VIEW OF 2' TONGUE AND GROOVE PLANKING IN ...</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. DETAIL VIEW OF 2' TONGUE AND GROOVE PLANKING IN WATER CONTROL BOX. THIS SAME PLANKING IS USED AS CRIBBING FOR BOTH <span class="hlt">EAST</span> DAM AND WEST DAM - Three Bears Lake & Dams, Water Control Box, North of Marias Pass, <span class="hlt">East</span> <span class="hlt">Glacier</span> <span class="hlt">Park</span>, <span class="hlt">Glacier</span> County, MT</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/p1386f/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/p1386f/"><span><span class="hlt">Glaciers</span> of Asia</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Williams, Richard S.; Ferrigno, Jane G.</p> <p>2010-01-01</p> <p> systems of the world including the Himalaya, Karakorum, Tien Shan and Altay mountain ranges. The <span class="hlt">glaciers</span> are widely scattered and cover an area of about 59,425 km2. The mountain <span class="hlt">glaciers</span> may be classified as maritime, subcontinental or extreme continental. In Afghanistan, more than 3,000 small <span class="hlt">glaciers</span> occur in the Hindu Kush and Pamir mountains. Most <span class="hlt">glaciers</span> occur on north-facing slopes shaded by mountain peaks and on <span class="hlt">east</span> and southeast slopes that are shaded by monsoon clouds. The <span class="hlt">glaciers</span> provide vital water resources to the region and cover an area of about 2,700 km2. <span class="hlt">Glaciers</span> of northern Pakistan are some of the largest and longest mid-latitude <span class="hlt">glaciers</span> on Earth. They are located in the Hindu Kush, Himalaya, and Karakoram mountains and cover an area of about 15,000 km2. <span class="hlt">Glaciers</span> here are important for their role in providing water resources and their hazard potential. The <span class="hlt">glaciers</span> in India are located in the Himalaya and cover about 8,500 km2. The Himalaya contains one of the largest reservoirs of snow and ice outside the polar regions. The <span class="hlt">glaciers</span> are a major source of fresh water and supply meltwater to all the rivers in northern India, thereby affecting the quality of life of millions of people. In Nepal, the <span class="hlt">glaciers</span> are located in the Himalaya as individual <span class="hlt">glaciers</span>; the <span class="hlt">glacierized</span> area covers about 5,324 km2. The region is the highest mountainous region on Earth and includes the Mt. Everest region. <span class="hlt">Glaciers</span> in the Bhutan Himalaya have a total area of about 1,317 km2. Many recent <span class="hlt">glacier</span> studies are focused on <span class="hlt">glacier</span> lakes that have the potential of generating dangerous <span class="hlt">glacier</span> lake outburst floods. Research on the <span class="hlt">glaciers</span> of the middle-latitude, high-mountain <span class="hlt">glaciers</span> of Asia has also focused on the information contained in the ice cores from the <span class="hlt">glaciers</span>. This information helps in the reconstruction of paleoclimatic records, and the computer modeling of global climate change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.C11E..08R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.C11E..08R"><span>Himalayan <span class="hlt">glaciers</span>: understanding contrasting patterns of <span class="hlt">glacier</span> behavior using multi-temporal satellite imagery</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Racoviteanu, A.</p> <p>2014-12-01</p> <p>High rates of <span class="hlt">glacier</span> retreat for the last decades are often reported, and believed to be induced by 20th century climate changes. However, regional <span class="hlt">glacier</span> fluctuations are complex, and depend on a combination of climate and local topography. Furthermore, in ares such as the Hindu-Kush Himalaya, there are concerns about warming, decreasing monsoon precipitation and their impact on local <span class="hlt">glacier</span> regimes. Currently, the challenge is in understanding the magnitude of feedbacks between large-scale climate forcing and small-scale <span class="hlt">glacier</span> behavior. Spatio-temporal patterns of <span class="hlt">glacier</span> distribution are still llimited in some areas of the high Hindu-Kush Himalaya, but multi-temporal satellite imagery has helped fill spatial and temporal gaps in regional <span class="hlt">glacier</span> parameters in the last decade. Here I present a synopsis of the behavior of <span class="hlt">glaciers</span> across the Himalaya, following a west to <span class="hlt">east</span> gradient. In particular, I focus on spatial patterns of <span class="hlt">glacier</span> parameters in the eastern Himalaya, which I investigate at multi-spatial scales using remote sensing data from declassified Corona, ASTER, Landsat ETM+, Quickbird and Worldview2 sensors. I also present the use of high-resolution imagery, including texture and thermal analysis for mapping <span class="hlt">glacier</span> features at small scale, which are particularly useful in understanding surface trends of debris-covered <span class="hlt">glaciers</span>, which are prevalent in the Himalaya. I compare and contrast spatial patterns of <span class="hlt">glacier</span> area and élévation changes in the monsoon-influenced eastern Himalaya (the Everest region in the Nepal Himalaya and Sikkim in the Indian Himalaya) with other observations from the dry western Indian Himalaya (Ladakh and Lahul-Spiti), both field measurements and remote sensing-based. In the eastern Himalaya, results point to <span class="hlt">glacier</span> area change of -0.24 % ± 0.08% per year from the 1960's to the 2006's, with a higher rate of retreat in the last decade (-0.43% /yr). Debris-covered <span class="hlt">glacier</span> tongues show thinning trends of -30.8 m± 39 m</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70177786','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70177786"><span>Suppression of invasive lake trout in an isolated backcountry lake in <span class="hlt">Glacier</span> National <span class="hlt">Park</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>Fredenberg, C. R.; Muhlfeld, Clint C.; Guy, Christopher S.; D'Angelo, Vincent S.; Downs, Christopher C.; Syslo, John M.</p> <p>2017-01-01</p> <p>Fisheries managers have implemented suppression programmes to control non-native lake trout, Salvelinus namaycush (Walbaum), in several lakes throughout the western United States. This study determined the feasibility of experimentally suppressing lake trout using gillnets in an isolated backcountry lake in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA, for the conservation of threatened bull trout, Salvelinus confluentus (Suckley). The demographics of the lake trout population during suppression (2009–2013) were described, and those data were used to assess the effects of suppression scenarios on population growth rate (λ) using an age-structured population model. Model simulations indicated that the population was growing exponentially (λ = 1.23, 95% CI: 1.16–1.28) prior to suppression. However, suppression resulted in declining λ(0.61–0.79) for lake trout, which was concomitant with stable bull trout adult abundances. Continued suppression at or above observed exploitation levels is needed to ensure continued population declines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16340962','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16340962"><span>Planetary science: are there active <span class="hlt">glaciers</span> on Mars?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gillespie, Alan R; Montgomery, David R; Mushkin, Amit</p> <p>2005-12-08</p> <p>Head et al. interpret spectacular images from the Mars Express high-resolution stereo camera as evidence of geologically recent rock <span class="hlt">glaciers</span> in Tharsis and of a piedmont ('hourglass') <span class="hlt">glacier</span> at the base of a 3-km-high massif <span class="hlt">east</span> of Hellas. They attribute growth of the low-latitude <span class="hlt">glaciers</span> to snowfall during periods of increased spin-axis obliquity. The age of the hourglass <span class="hlt">glacier</span>, considered to be inactive and slowly shrinking beneath a debris cover in the absence of modern snowfall, is estimated to be more than 40 Myr. Although we agree that the maximum <span class="hlt">glacier</span> extent was climatically controlled, we find evidence in the images to support local augmentation of accumulation from snowfall through a mechanism that does not require climate change on Mars.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020846','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020846"><span>Assessing simulated ecosystem processes for climate variability research at <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>White, J.D.; Running, S.W.; Thornton, P.E.; Keane, R.E.; Ryan, K.C.; Fagre, D.B.; Key, C.H.</p> <p>1998-01-01</p> <p><span class="hlt">Glacier</span> National <span class="hlt">Park</span> served as a test site for ecosystem analyses than involved a suite of integrated models embedded within a geographic information system. The goal of the exercise was to provide managers with maps that could illustrate probable shifts in vegetation, net primary production (NPP), and hydrologic responses associated with two selected climatic scenarios. The climatic scenarios were (a) a recent 12-yr record of weather data, and (b) a reconstituted set that sequentially introduced in repeated 3-yr intervals wetter-cooler, drier-warmer, and typical conditions. To extrapolate the implications of changes in ecosystem processes and resulting growth and distribution of vegetation and snowpack, the model incorporated geographic data. With underlying digital elevation maps, soil depth and texture, extrapolated climate, and current information on vegetation types and satellite-derived estimates of a leaf area indices, simulations were extended to envision how the <span class="hlt">park</span> might look after 120 yr. The predictions of change included underlying processes affecting the availability of water and nitrogen. Considerable field data were acquired to compare with model predictions under current climatic conditions. In general, the integrated landscape models of ecosystem processes had good agreement with measured NPP, snowpack, and streamflow, but the exercise revealed the difficulty and necessity of averaging point measurements across landscapes to achieve comparable results with modeled values. Under the extremely variable climate scenario significant changes in vegetation composition and growth as well as hydrologic responses were predicted across the <span class="hlt">park</span>. In particular, a general rise in both the upper and lower limits of treeline was predicted. These shifts would probably occur along with a variety of disturbances (fire, insect, and disease outbreaks) as predictions of physiological stress (water, nutrients, light) altered competitive relations and hydrologic</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017QSRv..162...60A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017QSRv..162...60A"><span>Dust composition changes from Taylor <span class="hlt">Glacier</span> (<span class="hlt">East</span> Antarctica) during the last glacial-interglacial transition: A multi-proxy approach</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aarons, Sarah M.; Aciego, Sarah M.; Arendt, Carli A.; Blakowski, Molly A.; Steigmeyer, August; Gabrielli, Paolo; Sierra-Hernández, M. Roxana; Beaudon, Emilie; Delmonte, Barbara; Baccolo, Giovanni; May, Nathaniel W.; Pratt, Kerri A.</p> <p>2017-04-01</p> <p>Mineral dust is transported in the atmosphere and deposited in oceans, ice sheets and the terrestrial biosphere. Temporal changes in locations of dust source areas and transport pathways have implications for global climate and biogeochemical cycles. The chemical and physical characterization of the dust record preserved in ice cores is useful for identifying of dust source regions, dust transport, dominant wind direction and storm trajectories. Here, we present a 50,000-year geochemical characterization of mineral dust entrapped in a horizontal ice core from the Taylor <span class="hlt">Glacier</span> in <span class="hlt">East</span> Antarctica. Strontium (Sr) and neodymium (Nd) isotopes, grain size distribution, trace and rare earth element (REE) concentrations, and inorganic ion (Cl- and Na+) concentrations were measured in 38 samples, corresponding to a time interval from 46 kyr before present (BP) to present. The Sr and Nd isotope compositions of insoluble dust in the Taylor <span class="hlt">Glacier</span> ice shows distinct changes between the Last Glacial Period (LGP in this study ranging from ∼46.7-15.3 kyr BP) the early Holocene (in this study ranging from ∼14.5-8.7 kyr BP), and zero-age samples. The 87Sr/86Sr isotopic composition of dust in the Taylor <span class="hlt">Glacier</span> ice ranged from 0.708 to 0.711 during the LGP, while the variability during the early Holocene is higher ranging from 0.707 to 0.714. The εNd composition ranges from 0.1 to -3.9 during the LGP, and is more variable from 1.9 to -8.2 during the early Holocene. The increased isotopic variability during the early Holocene suggests a shift in dust provenance coinciding with the major climate transition from the LGP to the Holocene. The isotopic composition and multiple physical and chemical constraints support previous work attributing Southern South America (SSA) as the main dust source to <span class="hlt">East</span> Antarctica during the LGP, and a combination of both local Ross Sea Sector dust sources and SSA after the transition into the Holocene. This study provides the first high time</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EnMan..50.1125G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EnMan..50.1125G"><span>Evaluating Tourist Perception of Environmental Changes as a Contribution to Managing Natural Resources in <span class="hlt">Glacierized</span> Areas: A Case Study of the Forni <span class="hlt">Glacier</span> (Stelvio National <span class="hlt">Park</span>, Italian Alps)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Garavaglia, Valentina; Diolaiuti, Guglielmina; Smiraglia, Claudio; Pasquale, Vera; Pelfini, Manuela</p> <p>2012-12-01</p> <p>Climate change effects are noticeably evident above the timberline where <span class="hlt">glacier</span> and permafrost processes and mass movements drive the surface evolution. In particular, the cryosphere shrinkage is deeply changing the features and characteristics of several <span class="hlt">glacierized</span> mountain areas of the world, and these modifications can also affect the landscape perception of tourists and mountaineers. On the one hand <span class="hlt">glacier</span> retreat is increasing the interest of tourists and visitors in areas witnessing clear climate change impacts; on the other hand cryosphere shrinkage can impact the touristic appeal of mountain territories which, diminishing their ice and snow coverage, are also losing part of their aesthetic value. Then, to promote <span class="hlt">glacierized</span> areas in a changing climate and to prepare exhaustive and actual proposals for sustainable tourism, it is important to deepen our knowledge about landscape perception of tourists and mountaineers and their awareness of the ongoing environmental modifications. Here we present the results from a pilot study we performed in summer 2009 on a representative <span class="hlt">glacierized</span> area of the Alps, the Forni Valley (Stelvio National <span class="hlt">Park</span>, Lombardy, Italy), a valley shaped by Forni, the largest Italian valley <span class="hlt">glacier</span>. During the 2009 summer season we asked tourists visiting the Forni Valley to complete a questionnaire. This study was aimed at both describing the features and characteristics of tourists and mountaineers visiting this Alpine zone in summer and evaluating their landscape perception and their ability to recognize climate change impacts and evidence. Our results suggest that the dissemination strategies in a natural protected area have to take into account not only the main landscape features but also the sites where the information will be given. In particular considering the peculiarities of the huts located in the area, such as their different accessibility and the fact that they are included or not in a mountaineering network like that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/50023','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/50023"><span>Development of a management plan for coast live oak forests affected by sudden oak death in <span class="hlt">East</span> Bay Regional <span class="hlt">Parks</span></span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Brice A. McPherson; Joshua O’Neill; Gregory Biging; Maggi Kelly; David L. Wood</p> <p>2015-01-01</p> <p>The <span class="hlt">East</span> Bay Regional <span class="hlt">Park</span> District maintains the largest urban <span class="hlt">park</span> system in the United States, comprising over 45 000 ha, and more than 1900 km of trails, with extensive forests bordering residential areas. Sudden oak death (SOD), caused by the introduced oomycete Phytophthora ramorum, was first detected in a district <span class="hlt">park</span> in 2001. Both...</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://images.nasa.gov/#/details-GSFC_20171208_Archive_e001640.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001640.html"><span>Torres del Paine National <span class="hlt">Park</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Grinding <span class="hlt">glaciers</span> and granite peaks mingle in Chile’s Torres del Paine National <span class="hlt">Park</span>. The Advanced Land Imager (ALI) on NASA’s Earth Observing-1 (EO-1) satellite captured this summertime image of the <span class="hlt">park</span> on January 21, 2013. This image shows just a portion of the <span class="hlt">park</span>, including Grey <span class="hlt">Glacier</span> and the mountain range of Cordillera del Paine. The rivers of glacial ice in Torres del Paine National <span class="hlt">Park</span> grind over bedrock, turning some of that rock to dust. Many of the <span class="hlt">glaciers</span> terminate in freshwater lakes, which are rich with glacial flour that colors them brown to turquoise. Skinny rivers connect some of the lakes to each other (image upper and lower right). Cordillera del Paine rises between some of the wide glacial valleys. The compact mountain range is a combination of soaring peaks and small <span class="hlt">glaciers</span>, most notably the Torres del Paine (Towers of Paine), three closely spaced peaks emblematic of the mountain range and the larger <span class="hlt">park</span>. By human standards, the mountains of Cordillera del Paine are quite old. But compared to the Rocky Mountains (70 million years old), and the Appalachians (about 480 million years), the Cordillera del Paine are very young—only about 12 million years old. A study published in 2008 described how scientists used zircon crystals to estimate the age of Cordillera del Paine. The authors concluded that the mountain range was built in three pulses, creating a granite laccolith, or dome-shaped feature, more than 2,000 meters (7,000 feet) thick. NASA Earth Observatory image created by Jesse Allen and Robert Simmon, using Advanced Land Imager data from the NASA EO-1 team. Caption by Michon Scott. Instrument: EO-1 - ALI View more info: earthobservatory.nasa.gov/IOTD/view.php?id=80266 Credit: NASA Earth Observatory 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GPC...165..137V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GPC...165..137V"><span>Early 21st century spatially detailed elevation changes of Jammu and Kashmir <span class="hlt">glaciers</span> (Karakoram-Himalaya)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vijay, Saurabh; Braun, Matthias</p> <p>2018-06-01</p> <p>Although a number of studies indicate the regional heterogeneity of the <span class="hlt">glacier</span> elevation and mass changes in high-mountain Asia in the early 21st century, little is known about these changes with high spatial detail for some of the regions. In this study we present respective <span class="hlt">glacier</span> elevation and mass change estimates in the Indian state of Jammu and Kashmir (JK) for the period 2000-2012. Our estimates are based on the interferometric analysis of SRTM DEM and the bistatic TanDEM-X data. On an average the JK <span class="hlt">East</span> (Karakoram) <span class="hlt">glaciers</span> showed less negative elevation changes (- 0.19 ± 0.22 m yr-1) compared to the JK West (Himalaya) <span class="hlt">glaciers</span> (- 0.50 ± 0.28 m yr-1). This agrees very well with previous studies that show a transition from larger changes in the western Himalaya to a steady-state situation in the Karakoram. We observe distinct elevation change patterns on a <span class="hlt">glacier</span> scale that is most likely linked to debris insulation and the enhanced ice melting due to supraglacial lakes, ponds and ice cliffs. We also found 16 surge-type <span class="hlt">glaciers</span> in the JK <span class="hlt">East</span> which were not documented before. In total, 25 <span class="hlt">glaciers</span> surged and 4 others appeared to be in a quiescent phase in the observation period. Our results also reveal that the <span class="hlt">glacier</span>-averaged elevation change rates of surge-type and non surge-type <span class="hlt">glaciers</span> in the JK <span class="hlt">East</span> region are not significantly different.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001938.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001938.html"><span>Matusevich <span class="hlt">Glacier</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>NASA image acquired September 6, 2010 The Matusevich <span class="hlt">Glacier</span> flows toward the coast of <span class="hlt">East</span> Antarctica, pushing through a channel between the Lazarev Mountains and the northwestern tip of the Wilson Hills. Constrained by surrounding rocks, the river of ice holds together. But stresses resulting from the glacier’s movement make deep crevasses, or cracks, in the ice. After passing through the channel, the <span class="hlt">glacier</span> has room to spread out as it floats on the ocean. The expanded area and the jostling of ocean waves prompts the ice to break apart, which it often does along existing crevasses. On September 6, 2010, the Advanced Land Imager (ALI) on NASA’s Earth Observing-1 (EO-1) satellite captured this natural-color image of the margin of Matusevich <span class="hlt">Glacier</span>. Shown here just past the rock-lined channel, the <span class="hlt">glacier</span> is calving large icebergs. Low-angled sunlight illuminates north-facing surfaces and casts long shadows to the south. Fast ice anchored to the shore surrounds both the <span class="hlt">glacier</span> tongue and the icebergs it has calved. Compared to the <span class="hlt">glacier</span> and icebergs, the fast ice is thinner with a smoother surface. Out to sea (image left), the sea ice is even thinner and moves with winds and currents. Matusevich <span class="hlt">Glacier</span> does not drain a significant amount of ice off of the Antarctic continent, so the glacier’s advances and retreats lack global significance. Like other Antarctic <span class="hlt">glaciers</span>, however, Matusevich helps glaciologists form a larger picture of Antarctica’s glacial health and ice sheet volume. NASA Earth Observatory image created by Jesse Allen and Robert Simmon, using EO-1 ALI data provided courtesy of the NASA EO-1 team. Caption by Michon Scott based on image interpretation by Robert Bindschadler, NASA Goddard Space Flight Center, and Walt Meier, National Snow and Ice Data Center. Instrument: EO-1 - ALI Credit: NASA Earth Observatory NASA Goddard Space Flight Center enables NASA’s mission through four scientific endeavors: Earth Science, Heliophysics, Solar</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012TCry....6.1031D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012TCry....6.1031D"><span>Variable <span class="hlt">glacier</span> response to atmospheric warming, northern Antarctic Peninsula, 1988-2009</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Davies, B. J.; Carrivick, J. L.; Glasser, N. F.; Hambrey, M. J.; Smellie, J. L.</p> <p>2012-09-01</p> <p>The northern Antarctic Peninsula has recently exhibited ice-shelf disintegration, <span class="hlt">glacier</span> recession and acceleration. However, the dynamic response of land-terminating, ice-shelf tributary and tidewater <span class="hlt">glaciers</span> has not yet been quantified or assessed for variability, and there are sparse data for <span class="hlt">glacier</span> classification, morphology, area, length or altitude. This paper firstly classifies the area, length, altitude, slope, aspect, geomorphology, type and hypsometry of 194 <span class="hlt">glaciers</span> on Trinity Peninsula, Vega Island and James Ross Island in 2009 AD. Secondly, this paper documents <span class="hlt">glacier</span> change 1988-2009. In 2009, the glacierised area was 8140±262 km2. From 1988-2001, 90% of <span class="hlt">glaciers</span> receded, and from 2001-2009, 79% receded. This equates to an area change of -4.4% for Trinity Peninsula eastern coast <span class="hlt">glaciers</span>, -0.6% for western coast <span class="hlt">glaciers</span>, and -35.0% for ice-shelf tributary <span class="hlt">glaciers</span> from 1988-2001. Tidewater <span class="hlt">glaciers</span> on the drier, cooler eastern Trinity Peninsula experienced fastest shrinkage from 1988-2001, with limited frontal change after 2001. <span class="hlt">Glaciers</span> on the western Trinity Peninsula shrank less than those on the <span class="hlt">east</span>. Land-terminating <span class="hlt">glaciers</span> on James Ross Island shrank fastest in the period 1988-2001. This <span class="hlt">east</span>-west difference is largely a result of orographic temperature and precipitation gradients across the Antarctic Peninsula, with warming temperatures affecting the precipitation-starved <span class="hlt">glaciers</span> on the eastern coast more than on the western coast. Reduced shrinkage on the western Peninsula may be a result of higher snowfall, perhaps in conjunction with the fact that these <span class="hlt">glaciers</span> are mostly grounded. Rates of area loss on the eastern side of Trinity Peninsula are slowing, which we attribute to the floating ice tongues receding into the fjords and reaching a new dynamic equilibrium. The rapid shrinkage of tidewater <span class="hlt">glaciers</span> on James Ross Island is likely to continue because of their low elevations and flat profiles. In contrast, the higher and steeper</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/1776/e/pdf/pp1776E.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/1776/e/pdf/pp1776E.pdf"><span>Geochronology of plutonic rocks and their tectonic terranes in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, southeast Alaska: Chapter E in Studies by the U.S. Geological Survey in Alaska, 2008-2009</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Brew, David A.; Tellier, Kathleen E.; Lanphere, Marvin A.; Nielsen, Diane C.; Smith, James G.; Sonnevil, Ronald A.</p> <p>2014-01-01</p> <p>We have identified six major belts and two nonbelt occurrences of plutonic rocks in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve and characterized them on the basis of geologic mapping, igneous petrology, geochemistry, and isotopic dating. The six plutonic belts and two other occurrences are, from oldest to youngest: (1) Jurassic (201.6–145.5 Ma) diorite and gabbro of the Lituya belt; (2) Late Jurassic (161.0–145.5 Ma) leucotonalite in Johns Hopkins Inlet; (3) Early Cretaceous (145.5–99.6 Ma) granodiorite and tonalite of the Muir-Chichagof belt; (4) Paleocene tonalite in Johns Hopkins Inlet (65.5–55.8 Ma); (5) Eocene granodiorite of the Sanak-Baranof belt; (6) Eocene and Oligocene (55.8–23.0 Ma) granodiorite, quartz diorite, and granite of the Muir-Fairweather felsic-intermediate belt; (7) Eocene and Oligocene (55.8–23.0 Ma) layered gabbros of the Crillon-La Perouse mafic belt; and (8) Oligocene (33.9–23.0 Ma) quartz monzonite and quartz syenite of the Tkope belt. The rocks are further classified into 17 different combination age-compositional units; some younger belts are superimposed on older ones. Almost all these plutonic rocks are related to Cretaceous and Tertiary subduction events. The six major plutonic belts intrude the three southeast Alaska geographic subregions in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, from west to <span class="hlt">east</span>: (1) the Coastal Islands, (2) the Tarr Inlet Suture Zone (which contains the Border Ranges Fault Zone), and (3) the Central Alexander Archipelago. Each subregion includes rocks assigned to one or more tectonic terranes. The various plutonic belts intrude different terranes in different subregions. In general, the Early Cretaceous plutons intrude rocks of the Alexander and Wrangellia terranes in the Central Alexander Archipelago subregion, and the Paleogene plutons intrude rocks of the Chugach, Alexander, and Wrangellia terranes in the Coastal Islands, Tarr Inlet Suture Zone, and Central Alexander Archipelago subregions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMGC31C1007B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMGC31C1007B"><span>Changes of <span class="hlt">glacier</span>, <span class="hlt">glacier</span>-fed rivers and lakes in Altai Tavan Bogd National <span class="hlt">Park</span>, Western Mongolia, based on multispectral satellite data from 1990 to 2017</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Batsaikhan, B.; Lkhamjav, O.; Batsaikhan, N.</p> <p>2017-12-01</p> <p>Impacts on <span class="hlt">glaciers</span> and water resource management have been altering through climate changes in Mongolia territory characterized by dry and semi-arid climate with low precipitation. Melting <span class="hlt">glaciers</span> are early indicators of climate change unlike the response of the forests which is slower and takes place over a long period of time. Mountain <span class="hlt">glaciers</span> are important environmental components of local, regional, and global hydrological cycles. The study calculates an overview of changes for <span class="hlt">glacier</span>, <span class="hlt">glacier</span>-fed rivers and lakes in Altai Tavan Bogd mountain, the Western Mongolia, based on the indexes of multispectral data and the methods typically applied in <span class="hlt">glacier</span> studies. Were utilized an integrated approach of Normalized Difference Snow Index (NDSI) and Normalized Difference Water Index (NDWI) to combine Landsat, MODIS imagery and digital elevation model, to identify <span class="hlt">glacier</span> cover are and quantify water storage change in lakes, and compared that with and climate parameters including precipitation, land surface temperature, evaporation, moisture. Our results show that melts of <span class="hlt">glacier</span> at the study area has contributed to significantly increase of water storage of lakes in valley of The Altai Tavan Bogd mountain. There is hydrologic connection that lake basin is directly fed by <span class="hlt">glacier</span> meltwater.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170003150','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170003150"><span>An Intensive Observation of Calving at Helheim <span class="hlt">Glacier</span>, <span class="hlt">East</span> Greenland</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Holland, David M.; Voytenko, Denis; Christianson, Knut; Dixon, Timothy H.; Mei, M. Jeffrey; Parizek, Byron R.; Vankova, Irena; Walker, Ryan T.; Walter, Jacob I.; Nicholls, Keith; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170003150'); toggleEditAbsImage('author_20170003150_show'); toggleEditAbsImage('author_20170003150_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170003150_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170003150_hide"></p> <p>2016-01-01</p> <p>Calving of glacial ice into the ocean from the Greenland Ice Sheet is an important component of global sea-level rise. The calving process itself is relatively poorly observed, understood, and modeled; as such, it represents a bottleneck in improving future global sea-level estimates in climate models. We organized a pilot project to observe the calving process at Helheim <span class="hlt">Glacier</span> in <span class="hlt">east</span> Greenland in an effort to better understand it. During an intensive one-week survey, we deployed a suite of instrumentation, including a terrestrial radar interferometer, global positioning system (GPS) receivers, seismometers, tsunameters, and an automated weather station. We were fortunate to capture a calving process and to measure various glaciological, oceanographic, and atmospheric parameters before, during, and after the event. One outcome of our observations is evidence that the calving process actually consists of a number of discrete events, spread out over time, in this instance over at least two days. This time span has implications for models of the process. Realistic projections of future global sea level will depend on an accurate parametrization of calving, and we argue that more sustained observations will be required to reach this objective.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AnGla..40..225P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AnGla..40..225P"><span>Basal and thermal control mechanisms of the Ragnhild <span class="hlt">glaciers</span>, <span class="hlt">East</span> Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pattyn, Frank; de Brabander, Sang; Huyghe, Ann</p> <p></p> <p>The Ragnhild <span class="hlt">glaciers</span> are three enhanced-flow features situated between the Sør Rondane and Yamato Mountains in eastern Dronning Maud Land, Antarctica. We investigate the glaciological mechanisms controlling their existence and behavior, using a three-dimensional numerical thermomechanical ice-sheet model including higher-order stress gradients. This model is further extended with a steady-state model of subglacial water flow, based on the hydraulic potential gradient. Both static and dynamic simulations are capable of reproducing the enhanced ice-flow features. Although basal topography is responsible for the existence of the flow pattern, thermomechanical effects and basal sliding seem to locally soften and lubricate the ice in the main trunks. Lateral drag is a contributing factor in balancing the driving stress, as shear margins can be traced over a distance of hundreds of kilometers along west Ragnhild <span class="hlt">glacier</span>. Different basal sliding scenarios show that central Ragnhild <span class="hlt">glacier</span> stagnates as west Ragnhild <span class="hlt">glacier</span> accelerates and progressively drains the whole catchment area by ice and water piracy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C13G..03W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C13G..03W"><span>The sleeping giant wakes its neighbors?: Observations of unexpected <span class="hlt">glacier</span> change around Law Dome, <span class="hlt">East</span> Antarctica in response to the changing ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Walker, C. C.; Gardner, A. S.; Nilsson, J.</p> <p>2017-12-01</p> <p>The <span class="hlt">East</span> Antarctic has started to show signs of measureable change, albeit slow. Of particular note is the Totten <span class="hlt">Glacier</span> that drains the Aurora Subglacial Basin (ASB), one of two vast basins in <span class="hlt">East</span> Antarctica with the potential to significantly alter sea levels. Totten experienced flow acceleration between 2001-2007 linked to warm waters reaching its subglacial cavity (Rintoul et al., 2016). Directly to the west of Totten are the Vincennes Bay (VB) <span class="hlt">glaciers</span>. They also drain the ASB, and are grounded considerably below sea level. They have not been identified as changing - until now. Recent mapping of Antarctic-wide velocity via Landsat image pairs (Gardner et al., 2017) confirmed earlier findings (Li et al. 2016) that Totten's 2001-2007 velocity increase has since stabilized. At the same time, we have detected evidence of increased flow acceleration in VB, between 2008-2015. Here, we characterize these recent changes in the VB <span class="hlt">glacier</span> system to determine (1) the mechanisms driving change; (2) if the changes signify long- or short-term response; (3) if the temporal offset in response between Totten ( 2001-2007) and VB <span class="hlt">glaciers</span> ( 2008-2015) is representative of a scavenging relationship between the Law Dome neighbors. We use several datasets in addition to Landsat-derived velocities. We use space-borne altimetry measurements from the ICESat (2003-2009) and CryoSat-2 (2011-present) missions, and airborne laser altimetry (Operation IceBridge) where available, to show that after separating the dynamic signal from the detected elevation signal, sections of VB <span class="hlt">glaciers</span> lowered by 0.25 m/yr between satellite epochs. Seemingly small, this means they have approximately doubled their rate of lowering since 2009. We use ice penetrating radar (HICARS) measurements (2009-2012) to determine changes in subsurface ice structure, layering, and bed topography. We identify possible links to changes in ocean conditions using shipboard CTD measurements and those collected via</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70018406','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70018406"><span>The Border Ranges fault system in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, Alaska: Evidence for major early Cenozoic dextral strike-slip motion</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Smart, K.J.; Pavlis, T.L.; Sisson, V.B.; Roeske, S.M.; Snee, L.W.</p> <p>1996-01-01</p> <p>The Border Ranges fault system of southern Alaska, the fundamental break between the arc basement and the forearc accretionary complex, is the boundary between the Peninsular-Alexander-Wrangellia terrane and the Chugach terrane. The fault system separates crystalline rocks of the Alexander terrane from metamorphic rocks of the Chugach terrane in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>. Mylonitic rocks in the zone record abundant evidence for dextral strike-slip motion along north-northwest-striking subvertical surfaces. Geochronologic data together with regional correlations of Chugach terrane rocks involved in the deformation constrain this movement between latest Cretaceous and Early Eocene (???50 Ma). These findings are in agreement with studies to the northwest and southeast along the Border Ranges fault system which show dextral strike-slip motion occurring between 58 and 50 Ma. Correlations between <span class="hlt">Glacier</span> Bay plutons and rocks of similar ages elsewhere along the Border Ranges fault system suggest that as much as 700 km of dextral motion may have been accommodated by this structure. These observations are consistent with oblique convergence of the Kula plate during early Cenozoic and forearc slivering above an ancient subduction zone following late Mesozoic accretion of the Peninsular-Alexander-Wrangellia terrane to North America.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344965p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344965p/"><span>3. AERIAL VIEW OF THREE BEARS LAKE, SHOWING OUTLET STREAM, ...</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. AERIAL VIEW OF THREE BEARS LAKE, SHOWING OUTLET STREAM, BURLINGTON NORTHERN TRACKS, AND U.S. HIGHWAY 2, LOOKING NORTHEAST - Three Bears Lake & Dams, North of Marias Pass, <span class="hlt">East</span> <span class="hlt">Glacier</span> <span class="hlt">Park</span>, <span class="hlt">Glacier</span> County, MT</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344974p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344974p/"><span>12. CLOSEUP VIEW OF CROSS SECTION OF SPILLWAY FIFTY FEET ...</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. CLOSE-UP VIEW OF CROSS SECTION OF SPILLWAY FIFTY FEET FROM LAKESHORE, SHOWING REMAINS OF SPILLWAY TIMBERS, LOOKING WEST - Three Bears Lake & Dams, North of Marias Pass, <span class="hlt">East</span> <span class="hlt">Glacier</span> <span class="hlt">Park</span>, <span class="hlt">Glacier</span> County, MT</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mt0305.photos.345039p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mt0305.photos.345039p/"><span>6. VIEW OF SPILLWAY TIMBERS AND WATER CONTROL BOX, 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>6. VIEW OF SPILLWAY TIMBERS AND WATER CONTROL BOX, SHOWING WATER CONTROL BOX WITH LOWERED LAKE LEVEL - Three Bears Lake & Dams, Water Control Box, North of Marias Pass, <span class="hlt">East</span> <span class="hlt">Glacier</span> <span class="hlt">Park</span>, <span class="hlt">Glacier</span> County, MT</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://rosap.ntl.bts.gov/view/dot/9550','DOTNTL'); return false;" href="https://rosap.ntl.bts.gov/view/dot/9550"><span>Kenai Fjords National <span class="hlt">Park</span> Over-the-Snow Transportation Feasibility Study.</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>2012-01-31</p> <p>Kenai Fjords National <span class="hlt">Park</span> seeks to expand winter access to the Exit <span class="hlt">Glacier</span> Area. Year-round access would better enable the <span class="hlt">park</span> to accomplish its mission related to visitor experience, education, and research. The road to the area is inaccessible t...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PolSc..10..132A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PolSc..10..132A"><span>Observations of vertical tidal motions of a floating iceberg in front of Shirase <span class="hlt">Glacier</span>, <span class="hlt">East</span> Antarctica, using a geodetic-mode GPS buoy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aoyama, Yuichi; Kim, Tae-Hee; Doi, Koichiro; Hayakawa, Hideaki; Higashi, Toshihiro; Ohsono, Shingo; Shibuya, Kazuo</p> <p>2016-06-01</p> <p>A dual-frequency GPS receiver was deployed on a floating iceberg downstream of the calving front of Shirase <span class="hlt">Glacier</span>, <span class="hlt">East</span> Antarctica, on 28 December 2011 for utilizing as floating buoy. The three-dimensional position of the buoy was obtained by GPS every 30 s with a 4-5-cm precision for ca. 25 days. The height uncertainty of the 1-h averaged vertical position was ∼0.5 cm, even considering the uncertainties of un-modeled ocean loading effects. The daily evolution of north-south (NS), <span class="hlt">east</span>-west (EW), and up-down (UD) motions shows periodic UD variations sometimes attaining an amplitude of 1 m. Observed amplitudes of tidal harmonics of major constituents were 88%-93% (O1) and 85%-88% (M2) of values observed in the global ocean tide models FES2004 and TPXO-8 Atlas. The basal melting rate of the iceberg is estimated to be ∼0.6 m/day, based on a firn densification model and using a quasi-linear sinking rate of the iceberg surface. The 30-s sampling frequency geodetic-mode GPS buoy helps to reveal ice-ocean dynamics around the calving front of Antarctic <span class="hlt">glaciers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/EJ1091718.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/EJ1091718.pdf"><span>Eco-Tourism Development Strategy Balurannational <span class="hlt">Park</span> in the Regency of Situbondo, <span class="hlt">East</span> Java, Indonesia</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>Siswanto, Adil; Moeljadi</p> <p>2015-01-01</p> <p>Baluran National <span class="hlt">Park</span> in the regency of Situbondo, <span class="hlt">East</span> Java-Indonesia, highly prospective for development of sustainable tourism that can improve the welfare of local people. The suitable tourism type is eco-tourism with local people involvement. The purposes of this study are: 1) To know the local people involvement in eco-tourism development;…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002EGSGA..27.3570H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002EGSGA..27.3570H"><span>Assessing <span class="hlt">Glacier</span> Hazards At Ghiacciaio Del Belvedere, Macugnaga, Italian Alps</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Haeberli, W.; Chiarle, M.; Mortara, G.; Mazza, A.</p> <p></p> <p>The uppermost section of the Valle Anzasca behind and above the community of Macugnaga in the Italian Alps is one of the most spectacular high-mountain land- scapes in Europe, with gigantic rock walls and numerous steep hanging <span class="hlt">glaciers</span>. Its main <span class="hlt">glacier</span>, Ghiacciaio del Belvedere at the foot of the huge Monte Rosa <span class="hlt">east</span> face, is a heavily debris-covered <span class="hlt">glacier</span> flowing on a thick sediment bed. Problems with floods, avalanches and debris flows from this ice body have been known for extended time periods. Most recently, however, the evolution of this highly dynamic environ- ment has become more dramatic. An outburst of Lago delle Locce, an ice-dammed lake at the confluenec of the tributary Ghiacciaio delle Locce with Ghiacciaio del Belvedere, caused heavy damage in 1979 and necessitated site investigation and con- struction work to be done for flood protection. The intermittent <span class="hlt">glacier</span> growth ten- dency in the 1970es induced strong bulging of the <span class="hlt">glacier</span> surface and, in places, caused the <span class="hlt">glacier</span> tongue to override historical morains and to destroy newly-grown forest stands. A surge-type flow acceleration started in the lower parts of the Monte- Rosa <span class="hlt">east</span> face during summer 2000, leading to strong crevassing and deformation of Ghiacciaio del Belvedere and extreme bulging of its orographic right margin. High water pressure and accelerated movement lasted into winter 2001/2002: the ice now started overriding the LIA moraine near Rifugio Zamboni of the CAI. In addition but rather independently, a most active detachment zone for rock falls and debris flows developed for several years now in the <span class="hlt">east</span> face of Monte Rosa, somewhat more to the south of the accelerated <span class="hlt">glacier</span> movement and at an altitude where relatively warm permafrost must be expected. Besides the scientific interest in these phenomena, the growing hazard potential to the local infrastructure must be considered seriously. Es- pecially potentials for the destabilization of large rock and ice masses in the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.H51D0833S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.H51D0833S"><span><span class="hlt">Glacier</span> Dynamics Within a Small Alpine Cirque</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sanders, J. W.; Cuffey, K. M.; MacGregor, K. R.; Kavanaugh, J. L.; Dow, C. F.</p> <p>2008-12-01</p> <p>Cirques, with their steep walls and overdeepened basins, have captivated the imagination of scientists since the mid-1800s. <span class="hlt">Glaciers</span> in cirques, by generating these spectacular amphitheater-shaped landforms, contribute significantly to erosion in the core of mountain ranges and are one of the principal agents responsible for the relief structure at high elevations. Yet comprehensive studies of the dynamics of cirque <span class="hlt">glaciers</span>, and their link to erosional processes, have never been undertaken. To this end, we acquired an extensive new set of measurements at the West Washmawapta <span class="hlt">Glacier</span>, which sits in a cirque on the <span class="hlt">east</span> side of Helmet Mountain in the Vermillion Range of the Canadian Rockies. Ice thickness surveys with ground penetrating radar revealed that the <span class="hlt">glacier</span> occupies a classic bowl-shaped depression complete with a nearly continuous riegel. Using GPS-derived surface velocities of a <span class="hlt">glacier</span>-wide grid network and the tilt of one borehole, we calculated the complete force balance of the <span class="hlt">glacier</span>. This analysis also produced a map of basal sliding velocity and a value for the viscosity of temperate ice. We will discuss the implications of these findings for the problem of how cirques are formed by glacial erosion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70035114','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70035114"><span>The Neoglacial landscape and human history of <span class="hlt">Glacier</span> Bay, <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, southeast Alaska, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Connor, C.; Streveler, G.; Post, A.; Monteith, D.; Howell, W.</p> <p>2009-01-01</p> <p>The Neoglacial landscape of the Huna Tlingit homeland in <span class="hlt">Glacier</span> Bay is recreated through new interpretations of the lower Bay's fjordal geomorphology, late Quaternary geology and its ethnographic landscape. Geological interpretation is enhanced by 38 radiocarbon dates compiled from published and unpublished sources, as well as 15 newly dated samples. Neoglacial changes in ice positions, outwash and lake extents are reconstructed for c. 5500?????"200 cal. yr ago, and portrayed as a set of three landscapes at 1600?????"1000, 500?????"300 and 300?????"200 cal. yr ago. This history reveals episodic ice advance towards the Bay mouth, transforming it from a fjordal seascape into a terrestrial environment dominated by <span class="hlt">glacier</span> outwash sediments and ice-marginal lake features. This extensive outwash plain was building in lower <span class="hlt">Glacier</span> Bay by at least 1600 cal. yr ago, and had filled the lower bay by 500 cal. yr ago. The geologic landscape evokes the human-described landscape found in the ethnographic literature. Neoglacial climate and landscape dynamism created difficult but endurable environmental conditions for the Huna Tlingit people living there. Choosing to cope with environmental hardship was perhaps preferable to the more severely deteriorating conditions outside of the Bay as well as conflicts with competing groups. The central portion of the outwash plain persisted until it was overridden by ice moving into Icy Strait between AD 1724?????"1794. This final ice advance was very abrupt after a prolonged still-stand, evicting the Huna Tlingit from their <span class="hlt">Glacier</span> Bay homeland. ?? 2009 SAGE Publications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344968p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344968p/"><span>6. VIEW OF THREE BEARS LAKE, SHOWING WASHED UP 12' ...</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>6. VIEW OF THREE BEARS LAKE, SHOWING WASHED UP 12' x 12' DAM SUPPORT TIMBERS, LOOKING NORTHEAST FROM SOUTH SIDE OF LAKE - Three Bears Lake & Dams, North of Marias Pass, <span class="hlt">East</span> <span class="hlt">Glacier</span> <span class="hlt">Park</span>, <span class="hlt">Glacier</span> County, MT</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://eric.ed.gov/?q=Why+AND+dream&pg=7&id=ED465852','ERIC'); return false;" href="https://eric.ed.gov/?q=Why+AND+dream&pg=7&id=ED465852"><span>Central <span class="hlt">Park</span> <span class="hlt">East</span> and Its Graduates: "Learning by Heart." The Series on School Reform.</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>Bensman, David</p> <p></p> <p>This book describes New York City's Central <span class="hlt">Park</span> <span class="hlt">East</span> (CPE) Elementary School, which provides inner city children with the highest quality educators and pedagogy and is considered one of the most academically enriching U.S. schools. The book gives voice to young adults who emerged from poverty as a result of powerful experiences within CPE.…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFMPP43A1219R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFMPP43A1219R"><span>Do <span class="hlt">Glaciers</span> on Cascade Volcanoes Behave Differently Than Other <span class="hlt">Glaciers</span> in the Region?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Riedel, J. L.; Ryane, C.; Osborn, J.; Davis, T.; Menounos, B.; Clague, J. J.; Koch, J.; Scott, K. M.; Reasoner, M.</p> <p>2006-12-01</p> <p>It has been suggested that <span class="hlt">glaciers</span> on two stratovolcanoes in the Cascade Range of Washington state, Mt. Baker and <span class="hlt">Glacier</span> Peak, achieved their maximum extent of the past 10,000 years during the early Holocene. These findings differ from most evidence in western North America, which indicates that Little Ice Age moraines represent the most extensive <span class="hlt">glacier</span> advances of the Holocene. Significant early Holocene advances are difficult to reconcile with the documented warm, dry conditions at this time in western North America. Our data indicate that <span class="hlt">glaciers</span> on these volcanoes responded similarly to Holocene climatic events as <span class="hlt">glaciers</span> in other areas in Washington and British Columbia. Heavy winter accumulation and favorable hypsometry have been proposed as the explanations for the unusual behavior of <span class="hlt">glaciers</span> on volcanoes compared to similar-sized <span class="hlt">glaciers</span> elsewhere in the Cascade Range. However, <span class="hlt">glacier</span> mass balance on the volcanoes is controlled by not only these factors, but also by <span class="hlt">glacier</span> geometry, snow erosion and ablation. Accumulation zones of <span class="hlt">glaciers</span> on isolated Cascade stratovolcanoes are high, but are narrow at the top. For example, the accumulation zone of Deming <span class="hlt">Glacier</span> on the southwest side of Mt. Baker extends above 3000 m asl, but due to its wedge shape lies largely below 2500 m asl. Furthermore, <span class="hlt">glaciers</span> on Mt. Baker and other symmetrical volcanoes have high ablation rates because they are not shaded, and south-southwest aspects are subject to erosion of snow by prevailing southwesterly winds. Modern <span class="hlt">glacier</span> observations in the North Cascades quantify the important influence of aspect and snow erosion on <span class="hlt">glacier</span> mass balance. For example, average equilibrium line altitude (ELA) of Easton <span class="hlt">Glacier</span> on the south flank of Mt. Baker is 2160 m, whereas the ELA of a north-facing cirque <span class="hlt">glacier</span> 25km to the <span class="hlt">east</span> is 2040m. Our research at Mt. Baker contradicts the claim of extensive early Holocene advances on the south flank of the volcano. Tephra set SC, which</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sim/2945/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sim/2945/"><span>Map Showing Limits of Tahoe Glaciation in Sequoia and Kings Canyon National <span class="hlt">Parks</span>, 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>Moore, James Gregory; Mack, Gregory S.</p> <p>2008-01-01</p> <p>The latest periods of extensive ice cover in the Sierra Nevada include the Tahoe glaciation followed by the Tioga glaciation, and evidence for these ice ages is widespread in the Sequoia and Kings Canyon National <span class="hlt">Parks</span> area. However, the timing of the advances and retreats of the <span class="hlt">glaciers</span> during the periods of glaciation continues to be a matter of debate. A compilation of existing work (Clark and others, 2003) defines the Tioga glaciation at 14-25 thousand years ago and splits the Tahoe glaciation into two stages that range from 42-50 and 140-200 thousand years ago. The extent of the Tahoe ice mass shown in the map area is considered to represent the younger Tahoe stage, 42-50 thousand years ago. Evidence of glaciations older than the Tahoe is limited in the southern Sierra Nevada. After the Tioga glaciation, only minor events with considerably less ice cover occurred. The Tioga glaciation was slightly less extensive than the Tahoe glaciation, and each covered about half of the area of Sequoia and Kings Canyon National <span class="hlt">Parks</span>. The Tahoe <span class="hlt">glaciers</span> extended 500-1,000 ft lower and 0.5-1.2 mi farther down valleys. Evidence for the Tahoe glacial limits is not as robust as that for Tioga, but the extent of the Tahoe ice is mapped because it covered a larger area and the ice did leave prominent moraines (piles of sediment and boulders deposited by <span class="hlt">glaciers</span> as they melted at their margins) lower on the <span class="hlt">east</span> front of the range. Current Sierra redwood (Sequoiadendron giganteum) groves occur in a belt on the west side of the Sierra Nevada, generally west of the area of Tahoe glaciation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344969p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mt0273.photos.344969p/"><span>7. CLOSEUP VIEW OF WASHED UP 12' x 12' DAM ...</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. CLOSE-UP VIEW OF WASHED UP 12' x 12' DAM SUPPORT TIMBERS, THREE BEARS LAKE, LOOKING NORTHEAST FROM SOUTH SIDE OF LAKE - Three Bears Lake & Dams, North of Marias Pass, <span class="hlt">East</span> <span class="hlt">Glacier</span> <span class="hlt">Park</span>, <span class="hlt">Glacier</span> County, MT</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/1008609','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/1008609"><span>Distribution of boreal toad populations in relation to estimated UV-B dose in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hossack, B.R.; Diamond, S.A.; Corn, P.S.</p> <p>2006-01-01</p> <p>A recent increase in ultraviolet B radiation is one hypothesis advanced to explain suspected or documented declines of the boreal toad (Bufo boreas Baird and Girard, 1852) across much of the western USA, where some experiments have shown ambient UV-B can reduce embryo survival. We examined B. boreas occupancy relative to daily UV-B dose at 172 potential breeding sites in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, to assess whether UV-B limits the distribution of toads. Dose estimates were based on ground-level UV-B data and the effects of elevation, local topographic and vegetative features, and attenuation in the water column. We also examined temporal trends in surface UV-B and spring snowpack to determine whether populations are likely to have experienced increased UV-B exposure in recent decades. We found no support for the hypothesis that UV-B limits the distribution of populations in the <span class="hlt">park</span>, even when we analyzed high-elevation ponds separately. Instead, toads were more likely to breed in water bodies with higher estimated UV-B doses. The lack of a detectable trend in surface UV-B since 1979, combined with earlier snow melt in the region and increasing forest density at high elevations, suggests B. boreas embryos and larvae likely have not experienced increased UV-B.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018TCry...12..103W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018TCry...12..103W"><span>Recent <span class="hlt">glacier</span> mass balance and area changes in the Kangri Karpo Mountains from DEMs and <span class="hlt">glacier</span> inventories</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Kunpeng; Liu, Shiyin; Jiang, Zongli; Xu, Junli; Wei, Junfeng; Guo, Wanqin</p> <p>2018-01-01</p> <p>Due to the influence of the Indian monsoon, the Kangri Karpo Mountains in the south-<span class="hlt">east</span> of the Tibetan Plateau is in the most humid and one of the most important and concentrated regions containing maritime (temperate) <span class="hlt">glaciers</span>. <span class="hlt">Glacier</span> mass loss in the Kangri Karpo is an important contributor to global mean sea level rise, and changes run-off distribution, increasing the risk of glacial-lake outburst floods (GLOFs). Because of its inaccessibility and high labour costs, information about the Kangri Karpo <span class="hlt">glaciers</span> is still limited. Using geodetic methods based on digital elevation models (DEMs) derived from 1980 topographic maps from the Shuttle Radar Topography Mission (SRTM) (2000) and from TerraSAR-X/TanDEM-X (2014), this study has determined <span class="hlt">glacier</span> elevation changes. <span class="hlt">Glacier</span> area and length changes between 1980 and 2015 were derived from topographical maps and Landsat TM/ETM+/OLI images. Results show that the Kangri Karpo contained 1166 <span class="hlt">glaciers</span> with an area of 2048.50 ± 48.65 km2 in 2015. Ice cover diminished by 679.51 ± 59.49 km2 (24.9 ± 2.2 %) or 0.71 ± 0.06 % a-1 from 1980 to 2015, although nine <span class="hlt">glaciers</span> advanced. A <span class="hlt">glacierized</span> area of 788.28 km2, derived from DEM differencing, experienced a mean mass loss of 0.46 ± 0.08 m w.e. a-1 from 1980 to 2014. Shrinkage and mass loss accelerated significantly from 2000 to 2015 compared to 1980-2000, consistent with a warming climate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.C21A0728W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C21A0728W"><span>Spatio-temporal Variation in <span class="hlt">Glacier</span> Ice as Habitat for Harbor Seals in an Alaskan Tidewater <span class="hlt">Glacier</span> Fjord</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Womble, J. N.; McNabb, R. W.; Gens, R.; Prakash, A.</p> <p>2015-12-01</p> <p>Some of the largest aggregations of harbor seals (Phoca vitulina richardii) in Alaska occur in tidewater <span class="hlt">glacier</span> fjords where seals rest upon icebergs that are calved from tidewater <span class="hlt">glaciers</span> into the marine environment. The distribution, amount, and size of floating ice in fjords are likely important factors influencing the spatial distribution and abundance of harbor seals; however, fine-scale characteristics of ice habitat that are used by seals have not been quantified using automated methods. We quantified the seasonal changes in ice habitat for harbor seals in Johns Hopkins Inlet, a tidewater <span class="hlt">glacier</span> fjord in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, Alaska, using aerial photography, object-based image analysis, and spatial models. Aerial photographic surveys (n = 53) were conducted of seals and ice during the whelping (June) and molting (August) seasons from 2007-2014. Surveys were flown along a grid of 12 transects and high-resolution digital photos were taken directly under the plane using a vertically aimed camera. Seal abundance and spatial distribution was consistently higher during June (range: 1,672-4,340) than August (range: 1,075-2,582) and corresponded to the spatial distribution and amount of ice. Preliminary analyses from 2007 suggest that the average percent of icebergs (ice ≥ than 1.6m2) and brash ice (ice < 1.6m2) per scene were greater in June (icebergs: 1.8% ± 1.6%; brash ice: 43.8% ± 38.9%) than August (icebergs: 0.2% ± 0.7%; brash ice; 15.8% ± 26.4%). Iceberg angularity (an index of iceberg shape) was also greater in June (1.7 ± 0.9) than August (0.9 ± 0.9). Potential factors that may influence the spatio-temporal variation in ice habitat for harbor seals in tidewater <span class="hlt">glacier</span> fjords include frontal ablation rates of <span class="hlt">glaciers</span>, fjord circulation, and local winds. Harbor seals exhibit high seasonal fidelity to tidewater <span class="hlt">glacier</span> fjords, thus understanding the relationships between <span class="hlt">glacier</span> dynamics and harbor seal distribution will be critical for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70197916','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70197916"><span>Using stereo satellite imagery to account for ablation, entrainment, and compaction in volume calculations for rock avalanches on <span class="hlt">Glaciers</span>: Application to the 2016 Lamplugh Rock Avalanche in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Bessette-Kirton, Erin; Coe, Jeffrey A.; Zhou, Wendy</p> <p>2018-01-01</p> <p>The use of preevent and postevent digital elevation models (DEMs) to estimate the volume of rock avalanches on <span class="hlt">glaciers</span> is complicated by ablation of ice before and after the rock avalanche, scour of material during rock avalanche emplacement, and postevent ablation and compaction of the rock avalanche deposit. We present a model to account for these processes in volume estimates of rock avalanches on <span class="hlt">glaciers</span>. We applied our model by calculating the volume of the 28 June 2016 Lamplugh rock avalanche in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, Alaska. We derived preevent and postevent 2‐m resolution DEMs from WorldView satellite stereo imagery. Using data from DEM differencing, we reconstructed the rock avalanche and adjacent surfaces at the time of occurrence by accounting for elevation changes due to ablation and scour of the ice surface, and postevent deposit changes. We accounted for uncertainties in our DEMs through precise coregistration and an assessment of relative elevation accuracy in bedrock control areas. The rock avalanche initially displaced 51.7 ± 1.5 Mm3 of intact rock and then scoured and entrained 13.2 ± 2.2 Mm3 of snow and ice during emplacement. We calculated the total deposit volume to be 69.9 ± 7.9 Mm3. Volume estimates that did not account for topographic changes due to ablation, scour, and compaction underestimated the deposit volume by 31.0–46.8 Mm3. Our model provides an improved framework for estimating uncertainties affecting rock avalanche volume measurements in glacial environments. These improvements can contribute to advances in the understanding of rock avalanche hazards and dynamics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012cosp...39...44A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012cosp...39...44A"><span>Remote Sensing of Cryosphere: Estimation of Mass Balance Change in Himalayan <span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ambinakudige, Shrinidhi; Joshi, Kabindra</p> <p>2012-07-01</p> <p>Glacial changes are an important indicator of climate change. Our understanding mass balance change in Himalayan <span class="hlt">glaciers</span> is limited. This study estimates mass balance of some major <span class="hlt">glaciers</span> in the Sagarmatha National <span class="hlt">Park</span> (SNP) in Nepal using remote sensing applications. Remote sensing technique to measure mass balance of <span class="hlt">glaciers</span> is an important methodological advance in the highly rugged Himalayan terrain. This study uses ASTER VNIR, 3N (nadir view) and 3B (backward view) bands to generate Digital Elevation Models (DEMs) for the SNP area for the years 2002, 2003, 2004 and 2005. <span class="hlt">Glacier</span> boundaries were delineated using combination of boundaries available in the Global land ice measurement (GLIMS) database and various band ratios derived from ASTER images. Elevation differences, glacial area, and ice densities were used to estimate the change in mass balance. The results indicated that the rate of <span class="hlt">glacier</span> mass balance change was not uniform across <span class="hlt">glaciers</span>. While there was a decrease in mass balance of some <span class="hlt">glaciers</span>, some showed increase. This paper discusses how each <span class="hlt">glacier</span> in the SNP area varied in its annual mass balance measurement during the study period.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMGC34C..05T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMGC34C..05T"><span>Role of sub-regional variations on melting Response of Indian-Himalayan <span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tayal, S.; Hasnain, S. I.</p> <p>2010-12-01</p> <p><span class="hlt">Glaciers</span> play a crucial role in maintaining ecosystem stability as they act as buffers and regulate the runoff water supply from high mountains to the plains during both dry and wet spells. Retreat of Hindu Kush-Himalaya-Tibetan <span class="hlt">glaciers</span> is one of the major environmental problems facing the south Asian and south-<span class="hlt">east</span> Asian region. The Himalayan mountain range spans 2500 km <span class="hlt">east</span> to west and includes diverse cultures of five countries (Afghanistan, Pakistan, India, Tibet (China), Nepal, Bhutan) and a range of weather patterns, which has been strongly affected by regional climate change. The <span class="hlt">glaciers</span> of Indian Himalayan ranges covers an area of 19000 km2 contains over 9500 <span class="hlt">glaciers</span> and feed major perennial river systems like Indus, Ganges, Brahmaputra, and sustain the livelihood of over 0.5 billion south Asians. <span class="hlt">Glaciers</span> are melting fast but their response time varies from westerly nourished Kashmir Himalaya <span class="hlt">glaciers</span> to south-west monsoon nourished Sikkim Himalaya <span class="hlt">glaciers</span> based on regional climatic variations. Changes in mass balance of a <span class="hlt">glacier</span> are considered as the most direct representative of the impacts of meteorological parameters on the <span class="hlt">glacier</span> dynamic responses. A comparative study of mass balance, based on field measurements techniques is being conducted on two benchmark <span class="hlt">glaciers</span> in the Indian Himalaya. The <span class="hlt">glaciers</span> currently being monitored are Kolahoi <span class="hlt">glacier</span> (340 07 - 340 12 N: 750 16 - 750 23E), Kashmir Himalaya and E.Rathong <span class="hlt">glacier</span> (270 33 - 480 36 N: 880 06 - 880 08 E), Sikkim Himalaya. One year mass balance results (2008-2009) for both the benchmark <span class="hlt">glaciers</span> are now available and are being presented. Mass balance for Kolahoi <span class="hlt">glacier</span> located in sub-tropical to temperate setting and nourished by westerly system show range from -2.0 m.w.e. to -3.5 m.w.e. per annum. Whereas, the E. Rathong <span class="hlt">glacier</span> located in tropical climatic settings and nourished by SW monsoon system show range from -2.0 m.w.e. to -5.0 m.w.e. per annum. The (2009/2010) mass balance</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1146.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1146.pdf"><span>36 CFR 13.1146 - What other closures and restrictions apply to commercial fishermen and commercial fishing vessels?</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>2014-07-01</p> <p>... IN ALASKA Special Regulations-<span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve Commercial Fishing § 13.1146... Inlets. (b) Commercial fishing in the waters of the west arm of <span class="hlt">Glacier</span> Bay north of 58° 50.0′ N latitude... fishing regulations. (c) Commercial fishing in the <span class="hlt">east</span> arm of <span class="hlt">Glacier</span> Bay, north of an imaginary line...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title36-vol1/pdf/CFR-2012-title36-vol1-sec13-1146.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title36-vol1/pdf/CFR-2012-title36-vol1-sec13-1146.pdf"><span>36 CFR 13.1146 - What other closures and restrictions apply to commercial fishermen and commercial fishing vessels?</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>... IN ALASKA Special Regulations-<span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve Commercial Fishing § 13.1146... Inlets. (b) Commercial fishing in the waters of the west arm of <span class="hlt">Glacier</span> Bay north of 58° 50.0′ N latitude... fishing regulations. (c) Commercial fishing in the <span class="hlt">east</span> arm of <span class="hlt">Glacier</span> Bay, north of an imaginary line...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca2206.photos.182655p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca2206.photos.182655p/"><span>2. SOUTH SIDE, FROM <span class="hlt">PARK</span> ACROSS <span class="hlt">PARKING</span> LOT/F STREET, 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>2. SOUTH SIDE, FROM <span class="hlt">PARK</span> ACROSS <span class="hlt">PARKING</span> LOT/F STREET, LOOKING NORTH. - Oakland Naval Supply Center, Administration Building-Dental Annex-Dispensary, Between E & F Streets, <span class="hlt">East</span> of Third Street, Oakland, Alameda County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://rosap.ntl.bts.gov/view/dot/9555','DOTNTL'); return false;" href="https://rosap.ntl.bts.gov/view/dot/9555"><span><span class="hlt">Glacier</span> Going to the Sun Road Rehabilitation Mitigation Shuttle Bus Evaluation.</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>2008-03-31</p> <p>As a mitigation measure during reconstruction of the Going to the Sun Road, <span class="hlt">Glacier</span> National <span class="hlt">Park</span> operated a shuttle bus system along three routes during the 2007 season. This report presents a multi-dimensional evaluation of the transportation servi...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6734309-national-parks-latest-citations-from-ntis-database-published-search','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6734309-national-parks-latest-citations-from-ntis-database-published-search"><span>National <span class="hlt">parks</span>. (Latest citations from the NTIS database). Published Search</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></p> <p>The bibliography contains citations concerning U.S. and foreign national <span class="hlt">parks</span>. Citations discuss terrestrial biology, botany, coastal biology, fire ecology, endangered and exotic species, and resource analysis. Topics also include the impact of <span class="hlt">park</span> visitors on natural resources in the <span class="hlt">parks</span>, resource management, planning, and mapping. Information about specific <span class="hlt">parks</span> including Rocky Mountain, Great Smoky Mountains, Redwood, Grand Canyon, Sequoia, <span class="hlt">Glacier</span> Bay, and others is presented. (Contains a minimum of 55 citations and includes a subject term index and title list.)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/wa0325.photos.040644p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/wa0325.photos.040644p/"><span>1. July 1988 <span class="hlt">EAST</span> (MAIN) ELEVATION, PROTECTION ASSISTANT'S RESIDENCE (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>1. July 1988 <span class="hlt">EAST</span> (MAIN) ELEVATION, PROTECTION ASSISTANT'S RESIDENCE (BUILDING 1092) - <span class="hlt">Glacier</span> Ranger Station, Protection Assistant's Residence, Washington State Route 542, <span class="hlt">Glacier</span>, Whatcom County, WA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70188098','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70188098"><span>Seafloor habitat mapping and classification in <span class="hlt">Glacier</span> Bay, Alaska: Phase 1 & 2 1996-2004</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Hooge, Philip N.; Carlson, Paul R.; Mondragon, Jennifer; Etherington, Lisa L.; Cochran, G.R.</p> <p>2004-01-01</p> <p><span class="hlt">Glacier</span> Bay is a diverse fjord ecosystem with multiple sills, numerous tidewater <span class="hlt">glaciers</span> and a highly complex oceanographic system. The Bay was completely glaciated prior to the 1700’s and subsequently experienced the fastest glacial retreat recorded in historical times. Currently, some of the highest sedimentation rates ever observed occur in the Bay, along with rapid uplift (up to 2.5 cm/year) due to a combination of plate tectonics and isostatic rebound. <span class="hlt">Glacier</span> Bay is the second deepest fjord in Alaska, with depths over 500 meters. This variety of physical processes and bathymetry creates many diverse habitats within a relatively small area (1,255 km2 ). Habitat can be defined as the locality, including resources and environmental conditions, occupied by a species or population of organisms (Morrison et al 1992). Mapping and characterization of benthic habitat is crucial to an understanding of marine species and can serve a variety of purposes including: understanding species distributions and improving stock assessments, designing special management areas and marine protected areas, monitoring and protecting important habitats, and assessing habitat change due to natural or human impacts. In 1996, Congress recognized the importance of understanding benthic habitat for fisheries management by reauthorizing the Magnuson-Stevens Fishery Conservation and Management Act and amending it with the Sustainable Fisheries Act (SFA). This amendment emphasizes the importance of habitat protection to healthy fisheries and requires identification of essential fish habitat in management decisions. Recently, the National <span class="hlt">Park</span> Service’s Ocean Stewardship Strategy identified the creation of benthic habitat maps and sediment maps as crucial components to complete basic ocean <span class="hlt">park</span> resource inventories (Davis 2003). <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> managers currently have very limited knowledge about the bathymetry, sediment types, and various marine habitats of ecological</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED387493.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED387493.pdf"><span>Graduation by Portfolio at Central <span class="hlt">Park</span> <span class="hlt">East</span> Secondary School. A Series on Authentic Assessment and Accountability.</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>Darling-Hammond, Linda; Ancess, Jacqueline</p> <p></p> <p>Central <span class="hlt">Park</span> <span class="hlt">East</span> Secondary School (New York) is a school committed to authentic and learner-centered education. The school has developed an approach to assessing student performance that is active, authentic, and learner-centered. Students in the school's Senior Institute, a division comparable to the traditional grades 11 and 12, prepare 14…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70178374','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70178374"><span>Climate-induced <span class="hlt">glacier</span> and snow loss imperils alpine stream insects</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Giersch, J. Joseph; Hotaling, Scott; Kovach, Ryan; Jones, Leslie A.; Muhlfeld, Clint C.</p> <p>2017-01-01</p> <p>Climate warming is causing rapid loss of <span class="hlt">glaciers</span> and snowpack in mountainous regions worldwide. These changes are predicted to negatively impact the habitats of many range-restricted species, particularly endemic, mountaintop species dependent on the unique thermal and hydrologic conditions found only in <span class="hlt">glacier</span>-fed and snowmelt-driven alpine streams. Though progress has been made, existing understanding of the status, distribution, and ecology of alpine aquatic species, particularly in North America, is lacking, thereby hindering conservation and management programs. Two aquatic insects – the meltwater stonefly Lednia tumana and the <span class="hlt">glacier</span> stonefly Zapada <span class="hlt">glacier</span> – were recently proposed for listing under the U.S. Endangered Species Act due to climate-change-induced habitat loss. Using a large dataset (272 streams, 482 total sites) with high-resolution climate and habitat information, we describe the distribution, status, and key environmental features that limit L. tumana and Z. <span class="hlt">glacier</span> across the northern Rocky Mountains. Lednia tumana was detected in 113 streams (175 sites) within <span class="hlt">Glacier</span> National <span class="hlt">Park</span> (GNP) and surrounding areas. The probability of L. tumana occurrence increased with cold stream temperatures and close proximity to <span class="hlt">glaciers</span> and permanent snowfields. Similarly, densities of L. tumana declined with increasing distance from stream source. Zapada <span class="hlt">glacier</span> was only detected in 10 streams (20 sites), six in GNP and four in mountain ranges up to ~600 km southwest. Our results show that both L. tumana and Z. <span class="hlt">glacier</span> inhabit an extremely narrow distribution, restricted to short sections of cold, alpine streams often below <span class="hlt">glaciers</span> predicted to disappear over the next two decades. Climate warming-induced <span class="hlt">glacier</span> and snow loss clearly imperils the persistence of L. tumana and Z. <span class="hlt">glacier</span> throughout their ranges, highlighting the role of mountaintop aquatic invertebrates as sentinels of climate change in mid-latitude regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMPA51A..08M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMPA51A..08M"><span>Repeat Photography of Alaskan <span class="hlt">Glaciers</span> and Landscapes as Both Art and as a Means of Communicating Climat Change</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molnia, B. F.</p> <p>2013-12-01</p> <p>For nearly 15 years, I have used repeat photography of Alaskan <span class="hlt">glaciers</span> and landscapes to communicate to fellow scientists, policymakers, the media, and society that Alaskan <span class="hlt">glaciers</span> and landscapes have been experiencing significant change in response to post-Little Ice Age climate change. I began this pursuit after being contacted by a U.S. Department of the Interior senior official who requested unequivocal and unambiguous documentation that climate change was real and underway. After considering several options as to how best respond to this challenge, I decided that if a picture is worth a thousand words, then a pair of photographs, both with the same field of view, spanning a century or more, and showing dramatic differences, would speak volumes to documenting that dynamic climate change is occurring over a very broad region of Alaska. To me, understating the obvious with photographic pairs was the best mechanism to present irrefutable, unambiguous, nonjudgmental, as well as unequivocal visual documentation that climate change was both underway and real. To date, more than 150 pairs that meet these criteria have been produced. What has surprised me most is that the many of the photographs contained in the pairs present beautiful images of stark, remote landscapes that convey the majestic nature of this dynamic region with its unique topography and landscapes. Typically, over periods of just several decades, the photographed landscapes change from black and white to blue and green. White ice becomes blue water and dark rock becomes lush vegetation. Repeat photography is a technique in which a historical photograph and a modern photograph, both having the same field of view, are compared and contrasted to quantitatively and qualitatively determine their similarities and differences. I have used this technique from both ground-based photo stations and airborne platforms at Alaskan locations in Kenai Fjords National <span class="hlt">Park</span>, <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve</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/2018Geomo.311..127A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Geomo.311..127A"><span>Glaciation of alpine valleys: The <span class="hlt">glacier</span> - debris-covered <span class="hlt">glacier</span> - rock <span class="hlt">glacier</span> continuum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anderson, Robert S.; Anderson, Leif S.; Armstrong, William H.; Rossi, Matthew W.; Crump, Sarah E.</p> <p>2018-06-01</p> <p>Alpine ice varies from pure ice <span class="hlt">glaciers</span> to partially debris-covered <span class="hlt">glaciers</span> to rock <span class="hlt">glaciers</span>, as defined by the degree of debris cover. In many low- to mid-latitude mountain ranges, the few bare ice <span class="hlt">glaciers</span> that do exist in the present climate are small and are found where snow is focused by avalanches and where direct exposure to radiation is minimized. Instead, valley heads are more likely to be populated by rock <span class="hlt">glaciers</span>, which can number in the hundreds. These rock-cloaked <span class="hlt">glaciers</span> represent some of the most identifiable components of the cryosphere today in low- to mid-latitude settings, and the over-steepened snouts pose an often overlooked hazard to travel in alpine terrain. Geomorphically, rock <span class="hlt">glaciers</span> serve as conveyor belts atop which rock is pulled away from the base of cliffs. In this work, we show how rock <span class="hlt">glaciers</span> can be treated as an end-member case that is captured in numerical models of <span class="hlt">glaciers</span> that include ice dynamics, debris dynamics, and the feedbacks between them. Specifically, we focus on the transition from debris-covered <span class="hlt">glaciers</span>, where the modern equilibrium line altitude (ELA) intersects the topography, to rock <span class="hlt">glaciers</span>, where the modern ELA lies above the topography. On debris-covered <span class="hlt">glaciers</span> (i.e., <span class="hlt">glaciers</span> with a partial rock mantle), rock delivered to the <span class="hlt">glacier</span> from its headwall, or from sidewall debris swept into the <span class="hlt">glacier</span> at tributary junctions, travels englacially to emerge below the ELA. There it accumulates on the surface and damps the rate of melt of underlying ice. This allows the termini of debris-covered <span class="hlt">glaciers</span> to extend beyond debris-free counterparts, thereby decreasing the ratio of accumulation area to total area of the <span class="hlt">glacier</span> (AAR). In contrast, rock <span class="hlt">glaciers</span> (i.e., <span class="hlt">glaciers</span> with a full rock mantle) occur where and when the environmental ELA rises above the topography. They require avalanches and rockfall from steep headwalls. The occurrence of rock <span class="hlt">glaciers</span> reflects this dependence on avalanche sources</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/46693','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/46693"><span>Collection and production of indigenous plant material for national <span class="hlt">park</span> restoration</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Mark Majerus</p> <p>1999-01-01</p> <p>The National <span class="hlt">Park</span> Service is taking the "Restoration" approach to reestablishing native plant communities by salvaging topsoil and by seeding and planting native indigenous plant materials. In this way, they are making every effort to protect the genetic integrity of the often unique native plant resource. Since 1985, Yellowstone and <span class="hlt">Glacier</span> National <span class="hlt">Parks</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Flight+AND+planning&pg=5&id=ED231896','ERIC'); return false;" href="https://eric.ed.gov/?q=Flight+AND+planning&pg=5&id=ED231896"><span>Reviving an Inner City Community: The Drama of Urban Change in <span class="hlt">East</span> Humboldt <span class="hlt">Park</span> in Chicago.</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>Marciniak, Ed</p> <p></p> <p>In 1974, residents of <span class="hlt">East</span> Humboldt <span class="hlt">Park</span>, one of the oldest working class communities in Chicago, Illinois, gathered together in a common effort to reverse the process of urban decay and deterioration in their community. With the help of a hired consultant, the citizens planned the future of their community, a process that was completed in 1976,…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e000241.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e000241.html"><span>Biscayne National <span class="hlt">Park</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>On February 25, 2016, the Operational Land Imager (OLI) on the Landsat 8 satellite acquired this natural-color image of Biscayne National <span class="hlt">Park</span>. The <span class="hlt">park</span> encompasses the northernmost Florida Keys, starting from Miami to just north of Key Largo. The keys run like a spine through the center of the <span class="hlt">park</span>, with Biscayne Bay to the west and the Atlantic Ocean to the <span class="hlt">east</span>. The water-covered areas span more than 660 square kilometers (250 square miles) of the <span class="hlt">park</span>, making it the largest marine <span class="hlt">park</span> in the U.S. National <span class="hlt">Park</span> System. Biscayne protects the longest stretch of mangrove forest on the U.S. <span class="hlt">East</span> Coast, and one of the most extensive stretches of coral reef in the world. Read more: go.nasa.gov/1SWs1a3 Credit: NASA/Landsat8 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.C31A1247M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.C31A1247M"><span>Meltwater Induced <span class="hlt">Glacier</span> Landslides - Waxell Ridge, AK</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molnia, B. F.; Angeli, K. M.; Bratton, D. A.; Keeler, R. H.; Noyles, C.</p> <p>2006-12-01</p> <p>Within the past year, two large landslides have originated from south-facing peaks on Waxell Ridge, the bedrock massif that separates the Bagley Icefield from Bering <span class="hlt">Glacier</span>, Alaska. Each involves a near-summit hanging <span class="hlt">glacier</span>. In each instance, the presence of meltwater appears to be a triggering factor. The largest of the two, which occurred on September 14, 2005, originated from just below the summit of 3,236-m-high Mt Steller and landed on the surface of Bering <span class="hlt">Glacier</span>, nearly 2,500 m below. The Alaska Volcano Observatory estimated the volume of this landslide, which consisted of rock, <span class="hlt">glacier</span> ice, and snow, to be approximately 50 million cubic meters. Unlike most large Alaskan <span class="hlt">glacier</span>-related landslides, this one was not triggered by an earthquake. However, the energy that the slide released was intense enough to generate a seismic signal that was recorded around the world with magnitudes of 3.8 to greater than 5. The slide extended ~10 km down the Bering <span class="hlt">Glacier</span> from the point of impact. Much of the surface on which the slide occurred had a slope >50 degrees. The second landslide, located ~6 km to the west of Mt Steller, originated from a secondary summit of a 2,500- m-high unnamed peak. The date of its occurrence is unknown, but its toe sits on winter 2005-2006 snow. Both slides have been examined from helicopter and fixed-wing overflights, and with a variety of vertical and oblique aerial photographs. Oblique aerial photographs obtained of the Mt Steller slide on September 15, 2005 depict a 10-15-m-diameter moulin or englacial stream channel in the truncated 30-m-thick <span class="hlt">glacier</span> ice that comprises the <span class="hlt">east</span> wall of the landslide scarp. The presence of this unusual glacial-hydrologic feature at an elevation above 3,000 m, suggests that a large volume of water had recently been flowing on Mt Steller's <span class="hlt">east</span> ridge and that the water might have had a role in triggering the landslide. Similarly, there is evidence of an englacial channel on the west flank of the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70156580','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70156580"><span>Avalanche ecology and large magnitude avalanche events: <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fagre, Daniel B.; Peitzsch, Erich H.</p> <p>2010-01-01</p> <p>Large magnitude snow avalanches play an important role ecologically in terms of wildlife habitat, vegetation diversity, and sediment transport within a watershed. Ecological effects from these infrequent avalanches can last for decades. Understanding the frequency of such large magnitude avalanches is also critical to avalanche forecasting for the Going-to-the-Sun Road (GTSR). In January 2009, a large magnitude avalanche cycle occurred in and around <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana. The study site is the Little Granite avalanche path located along the GTSR. The study is designed to quantify change in vegetative cover immediately after a large magnitude event and document ecological response over a multi-year period. GPS field mapping was completed to determine the redefined perimeter of the avalanche path. Vegetation was inventoried using modified U.S. Forest Service Forest Inventory and Analysis plots, cross sections were taken from over 100 dead trees throughout the avalanche path, and an avalanche chronology was developed. Initial results indicate that the perimeter of this path was expanded by 30%. The avalanche travelled approximately 1200 vertical meters and 3 linear kilometers. Stands of large conifers as old as 150 years were decimated by the avalanche, causing a shift in dominant vegetation types in many parts of the avalanche path. Woody debris is a major ground cover up to 3 m in depth on lower portions of the avalanche path and will likely affect tree regrowth. Monitoring and measuring the post-avalanche vegetation recovery of this particular avalanche path provides a unique dataset for determining the ecological role of avalanches in mountain landscapes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1334445-estimates-glacier-mass-loss-contribution-streamflow-wind-river-range-wyoming-case-study','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1334445-estimates-glacier-mass-loss-contribution-streamflow-wind-river-range-wyoming-case-study"><span>Estimates of <span class="hlt">Glacier</span> Mass Loss and Contribution to Streamflow in the Wind River Range in Wyoming: Case Study</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>Marks, Jeffrey; Piburn, Jesse; Tootle, Glenn</p> <p>2014-09-11</p> <p>The Wind River Range is a continuous mountain range, approximately 160 km in length, in west-central Wyoming. The presence of <span class="hlt">glaciers</span> results in meltwater contributions to streamflow during the late summer (July, August, and September: JAS) when snowmelt is decreasing; temperatures are high; precipitation is low; evaporation rates are high; and municipal, industrial, and irrigation water are at peak demands. Therefore, the quantification of <span class="hlt">glacier</span> meltwater (e.g., volume and mass) contributions to late summer/early fall streamflow is important, given that this resource is dwindling owing to <span class="hlt">glacier</span> recession. The current research expands upon previous research efforts and identifies two glaciatedmore » watersheds, one on the <span class="hlt">east</span> slope (Bull Lake Creek) and one on the west slope (Green River) of the Wind River Range, in which unimpaired streamflow is available from 1966 to 2006. <span class="hlt">Glaciers</span> were delineated within each watershed and area estimates (with error) were obtained for the years 1966, 1989, and 2006. <span class="hlt">Glacier</span> volume (mass) loss (with error) was estimated by using empirically based volume-area scaling relationships. For 1966 to 2006, <span class="hlt">glacier</span> mass contributions to JAS streamflow on the <span class="hlt">east</span> slope were approximately 8%, whereas those on the west slope were approximately 2%. Furthermore, the volume-area scaling <span class="hlt">glacier</span> mass estimates compared favorably with measured (stereo pair remote sensed data) estimates of <span class="hlt">glacier</span> mass change for three <span class="hlt">glaciers</span> (Teton, Middle Teton, and Teepe) in the nearby Teton Range and one <span class="hlt">glacier</span> (Dinwoody) in the Wind River Range.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70006885','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70006885"><span>Marine predator surveys in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Bodkin, James L.; Kloecker, Kimberly A.; Coletti, Heather A.; Esslinger, George G.; Monson, Daniel H.; Ballachey, Brenda E.</p> <p>2002-01-01</p> <p>Since 1999, vessel based surveys to estimate species composition, distribution and relative abundance of marine birds and mammals have been conducted along coastal and pelagic (offshore) transects in <span class="hlt">Glacier</span> Bay, Alaska. Surveys have been conducted during winter (November-March) and summer (June). This annual report presents the results of those surveys conducted in March and June of 2001. Following completion of surveys in 2002 we will provide a final report of the results of all surveys conducted between 1999 and 2002.<span class="hlt">Glacier</span> Bay supports diverse and abundant assemblages of marine birds and mammals. In 2001 we identified 58 species of bird, 7 species of marine mammal, and 6 species of terrestrial mammal on transects sampled during winter and summer. Of course all species are not equally abundant. Among all taxa, in both seasons, sea ducks were the numerically dominant group. In their roles as consumers and because of their generally large size, marine mammals are also likely important in the consumption of energy produced in the <span class="hlt">Glacier</span> Bay ecosystem. Most common and abundant marine birds and mammals can be placed in either a fish based (e.g. alcids and pinnipeds), or a benthic invertebrate (e.g. sea ducks and sea otters) based food web.Distinct differences in the species composition and abundance of marine birds were observed between winter and summer surveys. Winter marine bird assemblages were dominated numerically (> 11,000; 65% of all birds) by a relatively few species of sea ducks (scoters, goldeneye, Bufflehead, Harlequin and Long-tailed ducks). The sea ducks were distributed almost exclusively along near shore habitats. The prevalence of sea ducks during the March surveys indicates the importance of <span class="hlt">Glacier</span> Bay as a wintering area for this poorly understood group of animals that occupy a high trophic position in a principally benthic invertebrate (mussel and clam) food web. Marine mammal assemblages were generally consistent between seasons, although</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C11E..02J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C11E..02J"><span>Rapid Holocene thinning of outlet <span class="hlt">glaciers</span> followed by readvance in the western Ross Embayment, Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jones, R. S.; Whitmore, R.; Mackintosh, A.; Norton, K. P.; Eaves, S.; Stutz, J.</p> <p>2017-12-01</p> <p>Investigating Antarctic deglaciation following the LGM provides an opportunity to better understand patterns, mechanisms and drivers of ice sheet retreat. In the Ross Sea sector, geomorphic features preserved on the seafloor indicate that streaming <span class="hlt">East</span> Antarctic outlet <span class="hlt">glaciers</span> once extended >100 km offshore of South Victoria Land prior to back-stepping towards their modern configurations. In order to adequately interpret the style and causes of this retreat, the timing and magnitude of corresponding ice thickness change is required. We present new constraints on ice surface lowering from Mawson <span class="hlt">Glacier</span>, an outlet of the <span class="hlt">East</span> Antarctic Ice Sheet that flows into the western Ross Sea. Surface-exposure (10Be) ages from samples collected in elevation transects above the modern ice surface reveal that rapid thinning occurred at 5-8 ka, broadly coeval with new ages of grounding-line retreat at 6 ka and rapid thinning recorded at nearby Mackay <span class="hlt">Glacier</span> at 7 ka. Our data also show that a moraine formed near to the modern ice margin of Mawson <span class="hlt">Glacier</span> at 0.8 ka, which, together with historical observations, indicates that <span class="hlt">glaciers</span> in this region readvanced during the last thousand years. We argue that 1) the accelerated thinning of outlet <span class="hlt">glaciers</span> was driven by local grounding-line retreat through overdeepened basins during the early-mid Holocene, and 2) the <span class="hlt">glaciers</span> subsequently readvanced, possibly linked to late Holocene sea-ice expansion, before retreating to their current positions. Our work demonstrates that these outlet <span class="hlt">glaciers</span> were closely coupled to environmental and topography-induced perturbations near their termini throughout the Holocene.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2009/1169/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2009/1169/"><span>Black and Brown Bear Activity at Selected Coastal Sites in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, Alaska: A Preliminary Assessment Using Noninvasive Procedures</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Partridge, Steve; Smith, Tom; Lewis, Tania</p> <p>2009-01-01</p> <p>A number of efforts in recent years have sought to predict bear activity in various habitats to minimize human disturbance and bear/human conflicts. Alaskan coastal areas provide important foraging areas for bears (Ursus americanus and U. arctos), particularly following den emergence when there may be no snow-free foraging alternatives. Additionally, coastal areas provide important food items for bears throughout the year. <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve (GLBA) in southeastern Alaska has extensive coastal habitats, and the National <span class="hlt">Park</span> Service (NPS) has been long interested in learning more about the use of these coastal habitats by bears because these same habitats receive extensive human use by <span class="hlt">park</span> visitors, especially kayaking recreationists. This study provides insight regarding the nature and intensity of bear activity at selected coastal sites within GLBA. We achieved a clearer understanding of bear/habitat relationships within GLBA by analyzing bear activity data collected with remote cameras, bear sign mapping, scat collections, and genetic analysis of bear hair. Although we could not quantify actual levels of bear activity at study sites, agreement among measures of activity (for example, sign counts, DNA analysis, and video record) lends support to our qualitative site assessments. This work suggests that habitat evaluation, bear sign mapping, and periodic scat counts can provide a useful index of bear activity for sites of interest.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/1016398','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/1016398"><span>Estimated ultraviolet radiation doses in wetlands in six national <span class="hlt">parks</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>Diamond, S.A.; Trenham, P.C.; Adams, Michael J.; Hossack, B.R.; Knapp, R.A.; Stark, L.; Bradford, D.; Corn, P.S.; Czarnowski, K.; Brooks, P.D.; Fagre, D.B.; Breen, B.; Dentenbeck, N.E.; Tonnessen, K.</p> <p>2005-01-01</p> <p>Ultraviolet-B radiation (UV-B, 280–320-nm wavelengths) doses were estimated for 1024 wetlands in six national <span class="hlt">parks</span>: Acadia (Acadia), <span class="hlt">Glacier</span> (<span class="hlt">Glacier</span>), Great Smoky Mountains (Smoky), Olympic (Olympic), Rocky Mountain (Rocky), and Sequoia/Kings Canyon (Sequoia). Estimates were made using ground-based UV-B data (Brewer spectrophotometers), solar radiation models, GIS tools, field characterization of vegetative features, and quantification of DOC concentration and spectral absorbance. UV-B dose estimates were made for the summer solstice, at a depth of 1 cm in each wetland. The mean dose across all wetlands and <span class="hlt">parks</span> was 19.3 W-h m−2 (range of 3.4–32.1 W-h m−2). The mean dose was lowest in Acadia (13.7 W-h m−2) and highest in Rocky (24.4 W-h m−2). Doses were significantly different among all <span class="hlt">parks</span>. These wetland doses correspond to UV-B flux of 125.0 μW cm−2 (range 21.4–194.7 μW cm−2) based on a day length, averaged among all <span class="hlt">parks</span>, of 15.5 h. Dissolved organic carbon (DOC), a key determinant of water-column UV-B flux, ranged from 0.6 (analytical detection limit) to 36.7 mg C L−1 over all wetlands and <span class="hlt">parks</span>, and reduced potential maximal UV-B doses at 1-cm depth by 1%–87 %. DOC concentration, as well as its effect on dose, was lowest in Sequoia and highest in Acadia (DOC was equivalent in Acadia, <span class="hlt">Glacier</span>, and Rocky). Landscape reduction of potential maximal UV-B doses ranged from zero to 77% and was lowest in Sequoia. These regional differences in UV-B wetland dose illustrate the importance of considering all aspects of exposure in evaluating the potential impact of UV-B on aquatic organisms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H43C1661G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H43C1661G"><span>Vegetative Succession in Recently Deglaciated Land in Kenai Fjords National <span class="hlt">Park</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Green, C.; Klein, A. G.; Cairns, D. M.</p> <p>2017-12-01</p> <p>Poleward vegetation expansion has affected Alaska for decades and due to recently increased rates of warming, the expansion will accelerate. Glacial recession in Kenai Fjords National <span class="hlt">Park</span> has exposed previously ice-covered land with vegetation succession occurring just a few years following glacial retreat. Land cover changes in recently deglaciated areas are affected by surface-air interactions, temperature gradients, and ecosystem development. Using satellite images from Landsat 5, 7, and 8 and the previous extents of four retreating <span class="hlt">glaciers</span> from 1985 to 2015 within Kenai Fjords National <span class="hlt">Park</span>, this study examines the relationship between deglaciation rates and vegetation greening. The <span class="hlt">glaciers</span>, Exit (-15.04 m/yr), Petrof (-31.12 m/yr), Lowell (-33.14 m/yr), and Yalik (-51.32 m/yr) were selected based on their location, whether they were land or lake terminating, and their average retreat rate measured between 1985 and 2015. These <span class="hlt">glaciers</span> have also been extensively studied. Combining historic <span class="hlt">glacier</span> extents with 371 summer (JJA) Landsat images gathered from Google's Earth Engine platform we identified annual summer changes in NDVI of locations that were deglaciated between 1985, 1995, 2005, and 2015. Summer temperature maximums were determined to be more correlated with deglaciation, as measured using NDSI, than mean summer temperatures. Using NDVI, heightened deglaciation rates were found to be reasonably correlated with vegetation succession. The faster retreating <span class="hlt">glaciers</span>, Lowell and Yalik, exhibited higher mean and maximum rates of increase of NDVI in their terminus areas than Exit and Petrof, the two slower retreating <span class="hlt">glaciers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27862701','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27862701"><span>Climate-induced <span class="hlt">glacier</span> and snow loss imperils alpine stream insects.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Giersch, J Joseph; Hotaling, Scott; Kovach, Ryan P; Jones, Leslie A; Muhlfeld, Clint C</p> <p>2017-07-01</p> <p>Climate warming is causing rapid loss of <span class="hlt">glaciers</span> and snowpack in mountainous regions worldwide. These changes are predicted to negatively impact the habitats of many range-restricted species, particularly endemic, mountaintop species dependent on the unique thermal and hydrologic conditions found only in <span class="hlt">glacier</span>-fed and snow melt-driven alpine streams. Although progress has been made, existing understanding of the status, distribution, and ecology of alpine aquatic species, particularly in North America, is lacking, thereby hindering conservation and management programs. Two aquatic insects - the meltwater stonefly (Lednia tumana) and the <span class="hlt">glacier</span> stonefly (Zapada <span class="hlt">glacier</span>) - were recently proposed for listing under the U.S. Endangered Species Act due to climate-change-induced habitat loss. Using a large dataset (272 streams, 482 total sites) with high-resolution climate and habitat information, we describe the distribution, status, and key environmental features that limit L. tumana and Z. <span class="hlt">glacier</span> across the northern Rocky Mountains. Lednia tumana was detected in 113 streams (175 sites) within <span class="hlt">Glacier</span> National <span class="hlt">Park</span> (GNP) and surrounding areas. The probability of L. tumana occurrence increased with cold stream temperatures and close proximity to <span class="hlt">glaciers</span> and permanent snowfields. Similarly, densities of L. tumana declined with increasing distance from stream source. Zapada <span class="hlt">glacier</span> was only detected in 10 streams (24 sites), six in GNP and four in mountain ranges up to ~600 km southwest. Our results show that both L. tumana and Z. <span class="hlt">glacier</span> inhabit an extremely narrow distribution, restricted to short sections of cold, alpine streams often below <span class="hlt">glaciers</span> predicted to disappear over the next two decades. Climate warming-induced <span class="hlt">glacier</span> and snow loss clearly imperils the persistence of L. tumana and Z. <span class="hlt">glacier</span> throughout their ranges, highlighting the role of mountaintop aquatic invertebrates as sentinels of climate change in mid-latitude regions. © 2016</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011TCD.....5.3541D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011TCD.....5.3541D"><span>A new <span class="hlt">glacier</span> inventory for 2009 reveals spatial and temporal variability in <span class="hlt">glacier</span> response to atmospheric warming in the Northern Antarctic Peninsula, 1988-2009</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Davies, B. J.; Carrivick, J. L.; Glasser, N. F.; Hambrey, M. J.; Smellie, J. L.</p> <p>2011-12-01</p> <p>The Northern Antarctic Peninsula has recently exhibited ice-shelf disintegration, <span class="hlt">glacier</span> recession and acceleration. However, the dynamic response of land-terminating, ice-shelf tributary and tidewater <span class="hlt">glaciers</span> has not yet been quantified or assessed for variability, and there are sparse published data for <span class="hlt">glacier</span> classification, morphology, area, length or altitude. This paper firstly uses ASTER images from 2009 and a SPIRIT DEM from 2006 to classify the area, length, altitude, slope, aspect, geomorphology, type and hypsometry of 194 <span class="hlt">glaciers</span> on Trinity Peninsula, Vega Island and James Ross Island. Secondly, this paper uses LANDSAT-4 and ASTER images from 1988 and 2001 and data from the Antarctic Digital Database (ADD) from 1997 to document <span class="hlt">glacier</span> change 1988-2009. From 1988-2001, 90 % of <span class="hlt">glaciers</span> receded, and from 2001-2009, 79 % receded. <span class="hlt">Glaciers</span> on the western side of Trinity Peninsula retreated relatively little. On the eastern side of Trinity Peninsula, the rate of recession of ice-shelf tributary <span class="hlt">glaciers</span> has slowed from 12.9 km2 a-1 (1988-2001) to 2.4 km2 a-1 (2001-2009). Tidewater <span class="hlt">glaciers</span> on the drier, cooler Eastern Trinity Peninsula experienced fastest recession from 1988-2001, with limited frontal retreat after 2001. Land-terminating <span class="hlt">glaciers</span> on James Ross Island also retreated fastest in the period 1988-2001. Large tidewater <span class="hlt">glaciers</span> on James Ross Island are now declining in areal extent at rates of up to 0.04 km2 a-1. This <span class="hlt">east</span>-west difference is largely a result of orographic temperature and precipitation gradients across the Antarctic Peninsula. Strong variability in tidewater <span class="hlt">glacier</span> recession rates may result from the influence of <span class="hlt">glacier</span> length, altitude, slope and hypsometry on <span class="hlt">glacier</span> mass balance. High snowfall means that the <span class="hlt">glaciers</span> on the Western Peninsula are not currently rapidly receding. Recession rates on the eastern side of Trinity Peninsula are slowing as the floating ice tongues retreat into the fjords and the <span class="hlt">glaciers</span> reach a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70027285','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70027285"><span>Posteruption <span class="hlt">glacier</span> development within the crater of Mount St. Helens, Washington, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Schilling, S.P.; Carrara, P.E.; Thompson, R.A.; Iwatsubo, E.Y.</p> <p>2004-01-01</p> <p>The cataclysmic eruption of Mount St. Helens on May 18, 1980, resulted in a large, north-facing amphitheater, with a steep headwall rising 700 m above the crater floor. In this deeply shaded niche a <span class="hlt">glacier</span>, here named the Amphitheater <span class="hlt">glacier</span>, has formed. Tongues of ice-containing crevasses extend from the main ice mass around both the <span class="hlt">east</span> and the west sides of the lava dome that occupies the center of the crater floor. Aerial photographs taken in September 1996 reveal a small <span class="hlt">glacier</span> in the southwest portion of the amphitheater containing several crevasses and a bergschrund-like feature at its head. The extent of the <span class="hlt">glacier</span> at this time is probably about 0.1 km2. By September 2001, the debris-laden <span class="hlt">glacier</span> had grown to about 1 km2 in area, with a maximum thickness of about 200 m, and contained an estimated 120,000,000 m3 of ice and rock debris. Approximately one-third of the volume of the <span class="hlt">glacier</span> is thought to be rock debris derived mainly from rock avalanches from the surrounding amphitheater walls. The newly formed Amphitheater <span class="hlt">glacier</span> is not only the largest <span class="hlt">glacier</span> on Mount St. Helens but its aerial extent exceeds that of all other remaining <span class="hlt">glaciers</span> combined. Published by University of Washington.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e000243.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e000243.html"><span>Acadia National <span class="hlt">Park</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Acadia National <span class="hlt">Park</span> is one of the most visited <span class="hlt">parks</span> in America, drawing more than 2.5 million visitors per year to the craggy, jagged coast of Maine. The <span class="hlt">park</span> is celebrating its 100th anniversary in 2016. On September 6, 2015, the Operational Land Imager (OLI) on the Landsat 8 satellite acquired these images of Acadia National <span class="hlt">Park</span> and its surroundings. Mountains and hills roll right up to the Atlantic Ocean in this rocky landscape carved by <span class="hlt">glaciers</span> at the end of the last Ice Age. Since the beginning of the 20th Century, the <span class="hlt">park</span> has been pieced together by donations and acquisitions of once-private lands, and it is still growing. Of the park’s 47,000 acres, more than 12,000 are privately owned lands under conservation agreements, while the rest is held by the National <span class="hlt">Park</span> Service. Mount Desert Island is the focal point of the <span class="hlt">park</span>, which also includes lands around a former naval base (Schoodic Peninsula), Isle au Haut, and several smaller islands. Read more: go.nasa.gov/2adyd8J Credit: NASA/Landsat8 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.coldregions.org/vufind/ajus/ajus','USGSPUBS'); return false;" href="http://www.coldregions.org/vufind/ajus/ajus"><span>Velocities of antarctic outlet <span class="hlt">glaciers</span> determined from sequential Landsat images</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>MacDonald, Thomas R.; Ferrigno, Jane G.; Williams, Richard S.; Lucchitta, Baerbel K.</p> <p>1989-01-01</p> <p>Approximately 91.0 percent of the volume of present-day <span class="hlt">glacier</span> ice on Earth is in Antarctica; Greenland contains about another 8.3 percent of the volume. Thus, together, these two great ice sheets account for an estimated 99.3 percent of the total. Long-term changes in the volume of <span class="hlt">glacier</span> ice on our planet are the result of global climate change. Because of the relationship of global ice volume to sea level (± 330 cubic kilometers of <span class="hlt">glacier</span> ice equals ± 1 millimeter sea level), changes in the mass balance of the antarctic ice sheet are of particular importance.Whether the mass balance of the <span class="hlt">east</span> and west antarctic ice sheets is positive or negative is not known. Estimates of mass input by total annual precipitation for the continent have been made from scattered meteorological observations (Swithinbank 1985). The magnitude of annual ablation of the ice sheet from calving of outlet <span class="hlt">glaciers</span> and ice shelves is also not well known. Although the velocities of outlet <span class="hlt">glaciers</span> can be determined from field measurements during the austral summer,the technique is costly, does not cover a complete annual cycle,and has been applied to just a few <span class="hlt">glaciers</span>. To increase the number of outlet <span class="hlt">glaciers</span> in Antarctica for which velocities have been determined and to provide additional data for under-standing the dynamics of the antarctic ice sheets and their response to global climate change, sequential Landsat image of several outlet <span class="hlt">glaciers</span> were measured.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70037469','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70037469"><span>Contribution of <span class="hlt">glacier</span> runoff to freshwater discharge into the Gulf of Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Neal, E.G.; Hood, E.; Smikrud, K.</p> <p>2010-01-01</p> <p>Watersheds along the Gulf of Alaska (GOA) are undergoing climate warming, <span class="hlt">glacier</span> volume loss, and shifts in the timing and volume of freshwater delivered to the eastern North Pacific Ocean. We estimate recent mean annual freshwater discharge to the GOA at 870 km3 yr-1. Small distributed coastal drainages contribute 78% of the freshwater discharge with the remainder delivered by larger rivers penetrating coastal ranges. Discharge from <span class="hlt">glaciers</span> and icefields accounts for 47% of total freshwater discharge, with 10% coming from <span class="hlt">glacier</span> volume loss associated with rapid thinning and retreat of <span class="hlt">glaciers</span> along the GOA. Our results indicate the region of the GOA from Prince William Sound to the <span class="hlt">east</span>, where <span class="hlt">glacier</span> runoff contributes 371 km3 yr -1, is vulnerable to future changes in freshwater discharge as a result of <span class="hlt">glacier</span> thinning and recession. Changes in timing and magnitude of freshwater delivery to the GOA could impact coastal circulation as well as biogeochemical fluxes to near-shore marine ecosystems and the eastern North Pacific Ocean. Copyright ?? 2010 by the American Geophysical Union.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22519575','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22519575"><span>Mercury distribution and deposition in <span class="hlt">glacier</span> snow over western China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Qianggong; Huang, Jie; Wang, Feiyue; Mark, Loewen; Xu, Jianzhong; Armstrong, Debbie; Li, Chaoliu; Zhang, Yulan; Kang, Shichang</p> <p>2012-05-15</p> <p>Western China is home to the largest aggregate of <span class="hlt">glaciers</span> outside the polar regions, yet little is known about how the <span class="hlt">glaciers</span> in this area affect the transport and cycling of mercury (Hg) regionally and globally. From 2005 to 2010, extensive <span class="hlt">glacier</span> snow sampling campaigns were carried out in 14 snowpits from 9 <span class="hlt">glaciers</span> over western China, and the vertical distribution profiles of Hg were obtained. The Total Hg (THg) concentrations in the <span class="hlt">glacier</span> snow ranged from <1 to 43.6 ng L(-1), and exhibited clear seasonal variations with lower values in summer than in winter. Spatially, higher THg concentrations were typically observed in <span class="hlt">glacier</span> snows from the northern region where atmospheric particulate loading is comparably high. <span class="hlt">Glacier</span> snowpit Hg was largely dependent on particulate matters and was associated with particulate Hg, which is less prone to postdepositional changes, thus providing a valuable record of atmospheric Hg deposition. Estimated atmospheric Hg depositional fluxes ranged from 0.74 to 7.89 μg m(-2) yr(-1), agreeing very well with the global natural values, but are one to two orders of magnitude lower than that of the neighboring <span class="hlt">East</span> Asia. Elevated Hg concentrations were observed in refrozen ice layers in several snowpits subjected to intense melt, indicating that Hg can be potentially released to meltwater.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e000236.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e000236.html"><span>Yosemite National <span class="hlt">Park</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Naked summits alternate with forested lowlands in Yosemite Valley, part of California’s Yosemite National <span class="hlt">Park</span>. During the Pleistocene Ice Age, <span class="hlt">glaciers</span> sculpted the underlying rocks in this region, leaving behind canyons, waterfalls, rugged peaks, and granite domes. As the ice retreated, forests grew, but forests only extend as high as 2,900 meters (9,500 feet) above sea level. Above the tree line are rocky landscapes with sparse alpine vegetation. So from the sky, Yosemite Valley appears as a light-and-dark patchwork of forest, rock, and shadow. The Enhanced Thematic Mapper Plus on NASA’s Landsat 7 satellite captured this true-color image of part of Yosemite Valley on August 18, 2001. The valley runs roughly <span class="hlt">east</span>-west, and tall granite peaks lining the valley’s southern side cast long shadows across the valley floor. On the valley’s northern side, steep slopes appear almost white. Along the valley floor, roadways form narrow, meandering lines of off-white, past <span class="hlt">parking</span> lots, buildings, and meadows. On the north side of Yosemite Valley is El Capitan. Shooting straight up more than 915 meters (3,000 feet) above the valley floor, El Capitan is considered the largest granite monolith in the world. This granite monolith sits across the valley from Bridalveil Fall, one of the valley’s most prominent waterfalls. Read more: go.nasa.gov/2bzGo3d Credit: NASA/Landsat7 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</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/2014AGUFM.C44B..06C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.C44B..06C"><span>Laboratory Experiments Investigating <span class="hlt">Glacier</span> Submarine Melt Rates and Circulation in an <span class="hlt">East</span> Greenland Fjord</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cenedese, C.</p> <p>2014-12-01</p> <p>Idealized laboratory experiments investigate the <span class="hlt">glacier</span>-ocean boundary dynamics near a vertical '<span class="hlt">glacier</span>' (i.e. no floating ice tongue) in a two-layer stratified fluid, similar to Sermilik Fjord where Helheim <span class="hlt">Glacier</span> terminates. In summer, the discharge of surface runoff at the base of the <span class="hlt">glacier</span> (subglacial discharge) intensifies the circulation near the <span class="hlt">glacier</span> and increases the melt rate with respect to that in winter. In the laboratory, the effect of subglacial discharge is simulated by introducing fresh water at melting temperatures from either point or line sources at the base of an ice block representing the <span class="hlt">glacier</span>. The circulation pattern observed both with and without subglacial discharge resembles those observed in previous studies. The buoyant plume of cold meltwater and subglacial discharge water entrains ambient water and rises vertically until it finds either the interface between the two layers or the free surface. The results suggest that the meltwater deposits within the interior of the water column and not entirely at the free surface, as confirmed by field observations. The submarine melt rate increases with the subglacial discharge rate. Furthermore, the same subglacial discharge causes greater submarine melting if it exits from a point source rather than from a line source. When the subglacial discharge exits from two point sources, two buoyant plumes are formed which rise vertically and interact. The results suggest that the distance between the two subglacial discharges influences the entrainment in the plumes and consequently the amount of submarine melting and the final location of the meltwater within the water column. Hence, the distribution and number of sources of subglacial discharge may play an important role in glacial melt rates and fjord stratification and circulation. Support was given by NSF project OCE-113008.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.C51B0465M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.C51B0465M"><span>Low-frequency radar sounder over <span class="hlt">Glaciers</span> in Alaska, Greenland and Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mouginot, J.; Rignot, E. J.; Gim, Y.; Kirchner, D. L.; Merritt, S.; Robison, W. T.</p> <p>2009-12-01</p> <p>Ice-thickness and basal layer topography measurements are needed to calculate fluxes through fast-flowing outlet <span class="hlt">glaciers</span> in Greenland, Alaska, Patagonia and Antarctica. However, relatively high attenuation of radio waves by dielectric absorption and volume scattering from englacial water restrains detection of the bed through warm deep ice. Using a low-frequency (1-5 MHz) airborne radar, we have sounded outlet fast <span class="hlt">glaciers</span> over Greenland (Store, Upernavik, Hellheim, …), <span class="hlt">East</span> Antarctica (David, Mertz, Dibble, Byrd, …) and Alaska (Bering, Maslapina, Bagley, …). We will show that we detected the bed through temperate ice up to 1000m thick over Bering and Maslapina <span class="hlt">Glaciers</span> and also point out difficulty in detecting bed of other Alaska <span class="hlt">glaciers</span> due to off-nadir returns. We will also make direct comparison of this radar and previous airborne measurements in Greenland and Antarctica in order to discuss a potential improvement of bedrock detectability in temperate ice.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70155353','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70155353"><span>Climate change links fate of <span class="hlt">glaciers</span> and an endemic alpine invertebrate</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Muhlfeld, Clint C.; Giersch, J. Joseph; Hauer, F. Richard; Pederson, Gregory T.; Luikart, Gordon; Peterson, Douglas P.; Downs, Christopher C.; Fagre, Daniel B.</p> <p>2011-01-01</p> <p>Climate warming in the mid- to high-latitudes and high-elevation mountainous regions is occurring more rapidly than anywhere else on Earth, causing extensive loss of <span class="hlt">glaciers</span> and snowpack. However, little is known about the effects of climate change on alpine stream biota, especially invertebrates. Here, we show a strong linkage between regional climate change and the fundamental niche of a rare aquatic invertebrate—themeltwater stonefly Lednia tumana—endemic toWaterton- <span class="hlt">Glacier</span> International Peace <span class="hlt">Park</span>, Canada and USA. L. tumana has been petitioned for listing under the U.S. Endangered Species Act due to climate-change-induced <span class="hlt">glacier</span> loss, yet little is known on specifically how climate impacts may threaten this rare species and many other enigmatic alpine aquatic species worldwide. During 14 years of research, we documented that L. tumana inhabits a narrow distribution, restricted to short sections (∼500 m) of cold, alpine streams directly below <span class="hlt">glaciers</span>, permanent snowfields, and springs. Our simulation models suggest that climate change threatens the potential future distribution of these sensitive habitats and persistence of L. tumana through the loss of <span class="hlt">glaciers</span> and snowfields. Mountaintop aquatic invertebrates are ideal early warning indicators of climate warming in mountain ecosystems. Research on alpine invertebrates is urgently needed to avoid extinctions and ecosystem change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70181183','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70181183"><span>Colonization and development of stream communities across a 200-year gradient in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Milner, Alexander M.; Knudsen, E. Eric; Soiseth, Chad; Robertson, Anne L.; Schell, Don; Phillips, Ian T.; Magnusson, Katrina</p> <p>2000-01-01</p> <p>In May 1997, physical and biological variables were studied in 16 streams of different ages and contrasting stages of development following glacial recession in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, southeast Alaska. The number of microcrustacean and macroinvertebrate taxa and juvenile fish abundance and diversity were significantly greater in older streams. Microcrustacean diversity was related to the amount of instream wood and percent pool habitat, while the number of macroinvertebrate taxa was related to bed stability, amount of instream wood, and percent pool habitat. The percent contribution of Ephemeroptera to stream benthic communities increased significantly with stream age and the amount of coarse benthic organic matter. Juvenile Dolly Varden (Salvelinus malma) were dominant in the younger streams, but juvenile coho salmon (Oncorhynchus kisutch) abundance was greater in older streams associated with increased pool habitat. Upstream lakes significantly influenced channel stability, percent Chironomidae, total macroinvertebrate and meiofaunal abundance, and percent fish cover. Stable isotope analyses indicated nitrogen enrichment from marine sources in macroinvertebrates and juvenile fish in older streams with established salmon runs. The findings are encapsulated in a conceptual summary of stream development that proposes stream assemblages to be determined by direct interactions with the terrestrial, marine, and lake ecosystems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.herpconbio.org/contents_vol3_issue1.html','USGSPUBS'); return false;" href="http://www.herpconbio.org/contents_vol3_issue1.html"><span>Post-breeding habitat use by adult Boreal Toads (Bufo boreas) after wildfire in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Guscio, C.G.; Hossack, B.R.; Eby, L.A.; Corn, P.S.</p> <p>2008-01-01</p> <p>Effects of wildfire on amphibians are complex, and some species may benefit from the severe disturbance of stand-replacing fire. Boreal Toads (Bufo boreas boreas) in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA increased in occurrence after fires in 2001 and 2003. We used radio telemetry to track adult B. boreas in a mosaic of terrestrial habitats with different burn severities to better understand factors related to the post-fire pulse in breeding activity. Toads used severely burned habitats more than expected and partially burned habitats less than expected. No toads were relocated in unburned habitat, but little of the study area was unburned and the expected number of observations in unburned habitat was < 3. High vagility of B. boreas and preference for open habitats may predispose this species to exploit recently disturbed landscapes. The long-term consequences of fire suppression likely have had different effects in different parts of the range of B. boreas. More information is needed, particularly in the northern Rocky Mountains, where toads are more likely to occupy habitats that have diverged from historic fire return intervals. Copyright ?? 2008. C. Gregory Guscio. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.C31A1235R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.C31A1235R"><span>An Adjoint Force-restore Model for <span class="hlt">Glacier</span> Terminus Fluctuations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ren, D.; Leslie, L.; Karoly, D.</p> <p>2006-12-01</p> <p>A linear inverse formula comprises the basis for an individual treatment of 7 central Asian (25-55°N; 70-95°E) <span class="hlt">glaciers</span>. The linear forward model is based on first order <span class="hlt">glacier</span> dynamics, and requires the knowledge of reference states of forcing and <span class="hlt">glacier</span> perturbation magnitude. In this study, the adjoint based 4D-var method was applied to optimally determine the reference states and make it possible to start the integration at an arbitrarily chosen time, and thus suitable to use the availability of the coupled general circulation model (CGCM) predictions of future temperature scenarios. Two sensitive yet uncertain <span class="hlt">glacier</span> parameters and reference states at year 1900 are inferred from observed <span class="hlt">glacier</span> length records distributed irregularly over the 20th century and the regional mean annual temperature anomaly (against 1961-1990 reference) time series. We rotated the temperature forcing for the Hadley Centre- Climatic Research Unit of the University of <span class="hlt">East</span> Anglia (HadCRUT2), the Global Historical Climatology Network (GHCN) observations, and the ensemble mean of multiple CGCM runs and compared the retrieval results. Because of the high resemblance between the three data sources after 1960, it was decided practicable to use the observed temperature as forcing in retrieving the model parameters and initial states and then run an extended period with forcing from ensemble mean CGCM temperature of the next century. The length fluctuation is estimated for the transient climate period with 9 CGCM simulations under SRES A2 (a strong emission scenario from the Special report on Emissions Scenarios). For the 60-year period 2000- 2060, all <span class="hlt">glaciers</span> experienced salient shrinkage, especially those with gentle slopes. Although nearly one-third the year 2000 length will be reduced for some small <span class="hlt">glaciers</span>, the very existence of the <span class="hlt">glaciers</span> studied here is not threatened by year 2060. The differences in individual <span class="hlt">glacier</span> responses are very large. No straightforward</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/wa0227.photos.369744p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/wa0227.photos.369744p/"><span>Historic view of penstock repair following 1910's landslide just <span class="hlt">east</span> ...</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>Historic view of penstock repair following 1910's landslide just <span class="hlt">east</span> of powerhouse; looking <span class="hlt">east</span>. (photographer unknown, ca. 1910's) - Nooksack Falls Hydroelectric Plant, Route 542, <span class="hlt">Glacier</span>, Whatcom County, WA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://arc.lib.montana.edu/snow-science/item/1139','USGSPUBS'); return false;" href="http://arc.lib.montana.edu/snow-science/item/1139"><span>Natural avalanches and transportation: A case study from <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Reardon, B.A.; Fagre, Daniel B.; Steiner, R.W.</p> <p>2004-01-01</p> <p>In January 2004, two natural avalanches (destructive class 3) derailed a freight train in John F. Stevens Canyon, on the southern boundary of <span class="hlt">Glacier</span> National <span class="hlt">Park</span>. The railroad tracks were closed for 29 hours due to cleanup and lingering avalanche hazard, backing up 112km of trains and shutting down Amtrak’s passenger service. The incident marked the fourth time in three winters that natural avalanches have disrupted transportation in the canyon, which is also the route of U.S. Highway 2. It was the latest in a 94-year history of accidents that includes three fatalities and the destruction of a major highway bridge. Despite that history and the presence of over 40 avalanche paths in the 16km canyon, mitigation is limited to nine railroad snow sheds and occasional highway closures. This case study examines natural avalanche cycles of the past 28 winters using data from field observations, a Natural Resources Conservation Service (NRCS) SNOTEL station, and data collected since 2001 at a high-elevation weather station. The avalanches occurred when storms with sustained snowfall buried a persistent near-surface faceted layer and/or were followed by rain-on-snow or dramatic warming (as much as 21oC in 30 minutes). Natural avalanche activity peaked when temperatures clustered near freezing (mean of -1.5oC at 1800m elev.). Avalanches initiated through rapid loading, rain falling on new snow, and/ or temperature-related changes in the mechanical properties of slabs. Lastly, the case study describes how recent incidents have prompted a unique partnership of land management agencies, private corporations and non-profit organizations to develop an avalanche mitigation program for the transportation corridor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025461','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025461"><span>The health of <span class="hlt">glaciers</span>: Recent changes in <span class="hlt">glacier</span> regime</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Meier, M.F.; Dyurgerov, M.B.; McCabe, G.J.</p> <p>2003-01-01</p> <p><span class="hlt">Glacier</span> wastage has been pervasive during the last century; small <span class="hlt">glaciers</span> and those in marginal environments are disappearing, large mid-latitude <span class="hlt">glaciers</span> are shrinking slightly, and arctic <span class="hlt">glaciers</span> are warming. Net mass balances during the last 40 years are predominately negative and both winter and summer balances (accumulation and ablation) and mass turnover are increasing, especially after 1988. Two principal components of winter balance time-series explain about 50% of the variability in the data. <span class="hlt">Glacier</span> winter balances in north and central Europe correlate with the Arctic Oscillation, and <span class="hlt">glaciers</span> in western North America correlate with the Southern Oscillation and Northern Hemisphere air temperature. The degree of synchronization for distant <span class="hlt">glaciers</span> relates to changes in time of atmospheric circulation patterns as well as differing dynamic responses.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/prof/p1386b/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/prof/p1386b/"><span><span class="hlt">Glaciers</span> of Antarctica</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Williams, Richard S.; Ferrigno, Jane G.</p> <p>1988-01-01</p> <p> have been included. Again, these represent only a small fraction of the large number of aerial photographs now available in various national collections. The chapter is divided into five geographic sections. The first is the Transantarctic Mountains in the Ross Sea area. Some very large outlet <span class="hlt">glaciers</span> flow from the <span class="hlt">East</span> Antarctic ice sheet through the Transantarctic Mountains to the Ross Ice Shelf. Byrd <span class="hlt">Glacier</span>, one of the largest in the world, drains an area of more than 1,000,000 km2. Next, images from the Indian Ocean sector are discussed. These include the Lambert <span class="hlt">Glacier</span>- Amery Ice Shelf system, so large that about 25 images must be mosaicked to cover its complex system of tributary <span class="hlt">glaciers</span>. Shirase <span class="hlt">Glacier</span>, a tidal outlet <span class="hlt">glacier</span> in the sector, flows at a speed of 2.5 km a-l. About 200 km inland and 200 km west of Shirase <span class="hlt">Glacier</span> lie the Queen Fabiola (?Yamato?) Mountains, whose extensive exposures of `blue ice? lay claim to being the world?s most important meteorite-collecting locality, with more than 4,700 meteorite fragments discovered since 1969. The Atlantic Ocean sector is fringed by ice shelves into which flow large ice streams like Jutulstraumen, Stancomb-Wills, Slessor, and Recovery <span class="hlt">Glaciers</span>. Filchner and Ronne Ice Shelves together cover an area two-thirds the size of Texas. From the western margin of the Ronne Ice Shelf, the north-trending arc of the Antarctic Peninsula, with its fjord and alpine landscape and fringing ice shelves, stretches towards South America. The Pacific Ocean sector begins with the Ellsworth Mountains, which include the highest peaks (Vinson Massif at 4,897 m) in Antarctica. The area between the Ellsworth Mountains and the eastern margin of the Ross Ice Shelf is fringed with small ice shelves and some major outlet <span class="hlt">glaciers</span>. One of these, Pine Island <span class="hlt">Glacier</span>, was found from comparing 1973 and 1975 images to have an average ice-front velocity of 2.4 km a-l. This part of Antarctica</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004glac.book.....H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004glac.book.....H"><span><span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hambrey, Michael; Alean, Jürg</p> <p>2004-12-01</p> <p><span class="hlt">Glaciers</span> are among the most beautiful natural wonders on Earth, as well as the least known and understood, for most of us. Michael Hambrey describes how <span class="hlt">glaciers</span> grow and decay, move and influence human civilization. Currently covering a tenth of the Earth's surface, <span class="hlt">glacier</span> ice has shaped the landscape over millions of years by scouring away rocks and transporting and depositing debris far from its source. <span class="hlt">Glacier</span> meltwater drives turbines and irrigates deserts, and yields mineral-rich soils as well as a wealth of valuable sand and gravel. However, <span class="hlt">glaciers</span> also threaten human property and life. Our future is indirectly connected with the fate of <span class="hlt">glaciers</span> and their influence on global climate and sea level. Including over 200 stunning photographs, the book takes the reader from the High-Arctic through North America, Europe, Asia, Africa, New Zealand and South America to the Antarctic. Michael Hambrey is Director of the Centre for Glaciology at the University of Wales, Aberystwyth. A past recipient of the Polar Medal, he was also given the Earth Science Editors' Outstanding Publication Award for the first edition of <span class="hlt">Glaciers</span> (Cambridge, 1995). Hambrey is also the author of Glacial Environments (British Columbia, 1994). JÜrg Alean is Professor of Geography at the Kantonsschule ZÜrcher Unterland in BÜlach, Switzerland.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/md1513.photos.216790p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/md1513.photos.216790p/"><span>Perspective view looking from the <span class="hlt">east</span> to the <span class="hlt">east</span> northeast ...</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 looking from the <span class="hlt">east</span> to the <span class="hlt">east</span> northeast facade, with Swiss Chalet in background, to replicate the view shown in MD-1109-J-18 - National <span class="hlt">Park</span> Seminary, Japanese Pagoda, 2805 Linden Lane, Silver Spring, Montgomery County, MD</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/1386e/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/1386e/report.pdf"><span><span class="hlt">Glaciers</span> of Europe</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Williams, Richard S.; Ferrigno, Jane G.</p> <p>1993-01-01</p> <p>ALPS: AUSTRIAN: An overview is provided on the occurrence of the <span class="hlt">glaciers</span> in the Eastern Alps of Austria and on the climatic conditions in this area, Historical documents on the <span class="hlt">glaciers</span> have been available since the Middle Ages. Special glaciological observations and topographic surveys of individual <span class="hlt">glaciers</span> were initiated as early as 1846. Recent data in an inventory based on aerial photographs taken in 1969 show 925 <span class="hlt">glaciers</span> in the Austrian Alps with a total area of 542 square kilometers. Present research topics include studies of mass and energy balance, relations of <span class="hlt">glaciers</span> and climate, physical glaciology, a complete inventory of the <span class="hlt">glaciers</span>, and testing of remote sensing methods. The location of the <span class="hlt">glacier</span> areas is shown on Landsat multispectral scanner images; the improved capabilities of the Landsat thematic mapper are illustrated with an example from the Oztaler Alpen group. ALPS: SWISS: According to a <span class="hlt">glacier</span> inventory published in 1976, which is based on aerial photography of 1973, there are 1,828 <span class="hlt">glacier</span> units in the Swiss Alps that cover a total area of 1fl42 square kilometers. The Rhonegletscher, currently the ninth largest in the country, was one of the first to be studied in detail. Its surface has been surveyed repeatedly; velocity profiles were measured, and the fluctuations of its terminus were mapped and recorded from 1874 to 1914. Recent research on the <span class="hlt">glacier</span> has included climatological, hydrological, and massbalance studies. Glaciological research has been conducted on various other <span class="hlt">glaciers</span> in Switzerland concerning <span class="hlt">glacier</span> hydrology, <span class="hlt">glacier</span> hazards, fluctuations of <span class="hlt">glacier</span> termini, ice mechanics, ice cores, and mass balance. Good maps are available showing the extent of <span class="hlt">glaciers</span> from the latter decades of the 19th century. More recently, the entire country has been mapped at scales of 1:25,000, 1:50,000, 1:100,000, 1:200,000, and 1:500,000. The 1:25,000-scale series very accurately represents the <span class="hlt">glaciers</span> as well as locates</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMIN33B1311N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMIN33B1311N"><span>Of Images, Archives, and Anonymity: <span class="hlt">Glacier</span> Photographs from Louise Arner Boyd's <span class="hlt">East</span> Greenland Expeditions, 1933, 1937, and 1938</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nelson, F. E.; Peschel, S. M.; Hall, D. K.</p> <p>2010-12-01</p> <p>Louise A. Boyd (1887-1972) was raised to wealth and privilege in San Raphael, CA. Her inherited fortune allowed unlimited travel, and in 1924 she arrived in Spitsbergen as a tourist. Infatuated by Arctic landscapes, Boyd resolved to return north in a more assertive role and ran three "preliminary" expeditions to Greenland, in 1926, 1928, and 1931. Boyd’s expeditions to <span class="hlt">East</span> Greenland in 1933, 1937, and 1938 were predictive of the type of campaign that after WWII would characterize government-sponsored and international scientific efforts. “Planned as a unit,” these campaigns, sponsored by the American Geographical Society (AGS), were thoroughly integrated scientific expeditions incorporating glaciology, periglacial and glacial geomorphology, bedrock geology, botany, hydrography, topographic surveys, tides and currents, and magnetic observations within representative areas. The goal of the expeditions was to provide comprehensive characterization of the physical environment. The volumes resulting from this work contain many large-scale hydrographic and topographic maps, photomosaics, <span class="hlt">glacier</span> maps, and chapters on the geology, glacial history, botany, and hydrology of the region. Boyd received extensive publicity for her Arctic expeditions, although much of it was concerned with the novelty of expeditions to remote locations being led by a woman. Boyd’s expeditions employed scientists who eventually became highly influential in their respective fields. Boyd employed, among others, the earth scientists J.H. Bretz, R.F. Flint, and A.L. Washburn. Other important personnel on these expeditions included AGS cartographer/surveyor O.M. Miller and his assistant, W.A. Wood, who employed novel ground-based photogrammetric techniques to construct a series of <span class="hlt">glacier</span> maps at scales as large as 1:5000. The maps featured detailed error analyses, and are probably the first large-scale maps of known accuracy to be made of the Greenland Ice Sheet’s outlet <span class="hlt">glaciers</span>. Boyd</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://arc.lib.montana.edu/snow-science/item.php?id=500','USGSPUBS'); return false;" href="http://arc.lib.montana.edu/snow-science/item.php?id=500"><span>Natural glide slab avalanches, <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, USA: A unique hazard and forecasting challenge</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Reardon, Blase; Fagre, Daniel B.; Dundas, Mark; Lundy, Chris</p> <p>2006-01-01</p> <p>In a museum of avalanche phenomena, glide cracks and glide avalanches might be housed in the “strange but true” section. These oddities are uncommon in most snow climates and tend to be isolated to specific terrain features such as bedrock slabs. Many glide cracks never result in avalanches, and when they do, the wide range of time between crack formation and slab failure makes them highly unpredictable. Despite their relative rarity, glide cracks and glide avalanches pose a regular threat and complex forecasting challenge during the annual spring opening of the Going-to-the-Sun Road in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, U.S.A. During the 2006 season, a series of unusual glide cracks delayed snow removal operations by over a week and provided a unique opportunity to record detailed observations of glide avalanches and characterize their occurrence and associated weather conditions. Field observations were from snowpits, crown profiles and where possible, measurements of slab thickness, bed surface slope angle, substrate and other physical characteristics. Weather data were recorded at one SNOTEL site and two automated stations located from 0.6-10 km of observed glide slab avalanches. Nearly half (43%) of the 35 glide slab avalanches recorded were Class D2-2.5, with 15% Class D3-D3.5. The time between glide crack opening and failure ranged from 2 days to over six weeks, and the avalanches occurred in cycles associated with loss of snow water equivalent and spikes in temperature and radiation. We conclude with suggest ions for further study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70118022','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70118022"><span>Modeling mountain pine beetle disturbance in <span class="hlt">Glacier</span> National <span class="hlt">Park</span> using multiple lines of evidence</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Assal, Timothy; Sibold, Jason</p> <p>2013-01-01</p> <p>Temperate forest ecosystems are subject to various disturbances which contribute to ecological legacies that can have profound effects on the structure of the ecosystem. Impacts of disturbance can vary widely in extent, duration and severity over space and time. Given that global climate change is expected to increase rates of forest disturbance, an understanding of these events are critical in the interpretation of contemporary forest patterns and those of the near future. We seek to understand the impact of the 1970s mountain pine beetle outbreak on the landscape of <span class="hlt">Glacier</span> National <span class="hlt">Park</span> and investigate any connection between this event and subsequent decades of extensive wildfire. The lack of spatially explicit data on the mountain pine beetle disturbance represents a major data gap and inhibits our ability to test for correlations between outbreak severity and fire severity. To overcome this challenge, we utilized multiple lines of evidence to model forest canopy mortality as a proxy for outbreak severity. We used historical aerial and landscape photos, reports, aerial survey data, a six year collection of Landsat imagery and abiotic data in combination with regression analysis. The use of remotely sensed data is critical in large areas where subsequent disturbance (fire) has erased some of the evidence from the landscape. Results indicate that this method is successful in capturing the spatial heterogeneity of the outbreak in a topographically complex landscape. Furthermore, this study provides an example on the use of existing data to reduce levels of uncertainty associated with an historic disturbance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1060156','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1060156"><span>INDEPENDENT VERIFICATION SURVEY REPORT FOR ZONE 1 OF THE <span class="hlt">EAST</span> TENNESSEE TECHNOLOGY <span class="hlt">PARK</span> IN OAK RIDGE, TENNESSEE</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>King, David A.</p> <p>2012-08-16</p> <p>Oak Ridge Associated Universities (ORAU) conducted in-process inspections and independent verification (IV) surveys in support of DOE's remedial efforts in Zone 1 of <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> (ETTP) in Oak Ridge, Tennessee. Inspections concluded that the remediation contractor's soil removal and survey objectives were satisfied and the dynamic verification strategy (DVS) was implemented as designed. Independent verification (IV) activities included gamma walkover surveys and soil sample collection/analysis over multiple exposure units (EUs).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMGC22C..06H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMGC22C..06H"><span>The potential for retreating alpine <span class="hlt">glaciers</span> to alter alpine ecosystems in the Colorado Front Range</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hall, E.; Baron, J.</p> <p>2013-12-01</p> <p><span class="hlt">Glaciers</span> are retreating at an unprecedented rate. In mid-latitude alpine ecosystems the presence of <span class="hlt">glaciers</span> and rock <span class="hlt">glaciers</span> govern rates and ecology of alpine and sub-alpine ecosystems. Changes in the thermal environment due to the loss of isothermal habitat and inputs from <span class="hlt">glacier</span> melt chemistry are altering alpine ecosystems in unpredictable ways. In particular, <span class="hlt">glacier</span> may be a source of nitrogen that is altering alpine ecosystem dynamics. Loch Vale Watershed (LVWS) located within Rocky Mountain National <span class="hlt">Park</span>. LVWS contains a surface <span class="hlt">glacier</span> (Andrew's <span class="hlt">glacier</span>) and a rock <span class="hlt">glacier</span> (Taylor's <span class="hlt">glacier</span>) at the headwater of each of the two drainages within the watershed. We collected precipitation from a National Atmospheric Deposition Site and surface water from multiple alpine lakes and streams during a particularly high and low snow year in the Colorado Front Range. We also sampled stream and lake sediments at each site to analyze the associated microbial community. Concentrations of nitrate and ammonium, relative abundance of amoA (the gene responsible for a key step in the microbial nitrification pathway), and the dual isotope signal to nitrate all point to snow melt as a key deliverer of nitrogen to ecosystems along the Colorado Front Range. However, late summer surface water chemistry is isotopically similar to the chemistry of glacial ice. This suggests that retreating <span class="hlt">glacier</span> may be an additional source of N to alpine ecosystems and have the potential to alter microbial community composition, biogeochemical rate processes, and ecosystem function. These dynamics are most likely not unique to the Colorado Front Range and should be globally distributed as <span class="hlt">glaciers</span> continue to retreat in high altitude ecosystems around the world.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://pubs.water.usgs.gov/sir20045057','USGSPUBS'); return false;" href="http://pubs.water.usgs.gov/sir20045057"><span>Glacial history and runoff components of the Tlikakila River Basin, Lake Clark National <span class="hlt">Park</span> and Preserve, Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Brabets, Timothy P.; March, Rod S.; Trabant, Dennis C.</p> <p>2004-01-01</p> <p>The Tlikakila River is located in Lake Clark National <span class="hlt">Park</span> and Preserve and drains an area of 1,610 square kilometers (622 square miles). Runoff from the Tlikakila River Basin accounts for about one half of the total inflow to Lake Clark. <span class="hlt">Glaciers</span> occupy about one third of the basin and affect the runoff characteristics of the Tlikakila River. As part of a cooperative study with the National <span class="hlt">Park</span> Service, <span class="hlt">glacier</span> changes and runoff characteristics in the Tlikakila River Basin were studied in water years 2001 and 2002. Based on analyses of remote sensing data and on airborne laser profiling, most <span class="hlt">glaciers</span> in the Tlikakila River Basin have retreated and thinned from 1957 to the present. Volume loss from 1957-2001 from the Tanaina <span class="hlt">Glacier</span>, the largest <span class="hlt">glacier</span> in the Tlikakila River Basin, was estimated to be 6.1 x 109 cubic meters or 1.4 x 108 cubic meters per year. For the 2001 water year, mass balance measurements made on the three largest <span class="hlt">glaciers</span> in the Tlikakila River BasinTanaina, <span class="hlt">Glacier</span> Fork, and North Forkall indicate a negative mass balance. Runoff measured near the mouth of the Tlikakila River for water year 2001 was 1.70 meters. Of this total, 0.18 meters (11 percent) was from <span class="hlt">glacier</span> ice melt, 1.27 meters (75 percent) was from snowmelt, 0.24 meters (14 percent) was from rainfall runoff, and 0.01 meters (1 percent) was from ground water. Although ground water is a small component of runoff, it provides a critical source of warm water for fish survival in the lower reaches of the Tlikakila River.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70197897','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70197897"><span>Local topography increasingly influences the mass balance of a retreating cirque <span class="hlt">glacier</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>Florentine, Caitlyn; Harper, Joel T.; Fagre, Daniel B.; Moore, Johnnie; Peitzsch, Erich H.</p> <p>2018-01-01</p> <p>Local topographically driven processes – such as wind drifting, avalanching, and shading – are known to alter the relationship between the mass balance of small cirque <span class="hlt">glaciers</span> and regional climate. Yet partitioning such local effects from regional climate influence has proven difficult, creating uncertainty in the climate representativeness of some <span class="hlt">glaciers</span>. We address this problem for Sperry <span class="hlt">Glacier</span> in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, USA, using field-measured surface mass balance, geodetic constraints on mass balance, and regional climate data recorded at a network of meteorological and snow stations. Geodetically derived mass changes during 1950–1960, 1960–2005, and 2005–2014 document average mass change rates during each period at −0.22 ± 0.12, −0.18 ± 0.05, and −0.10 ± 0.03 m w.e. yr−1, respectively. A correlation of field-measured mass balance and regional climate variables closely (i.e., within 0.08 m w.e. yr−1) predicts the geodetically measured mass loss from 2005 to 2014. However, this correlation overestimates <span class="hlt">glacier</span> mass balance for 1950–1960 by +1.20 ± 0.95 m w.e. yr−1. Our analysis suggests that local effects, not represented in regional climate variables, have become a more dominant driver of the net mass balance as the <span class="hlt">glacier</span> lost 0.50 km2 and retreated further into its cirque.</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://www.jstor.org/stable/i269589','USGSPUBS'); return false;" href="http://www.jstor.org/stable/i269589"><span>A half century of change in alpine treeline patterns at <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, U.S.A.</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Klasner, F.L.; Fagre, D.B.</p> <p>2002-01-01</p> <p>Using sequential aerial photography, we identified changes in the spatial distribution of subalpine fir (Abies lasiocarpa) habitat at the alpine treeline ecotone. Six 40-ha study sites in the McDonald Creek drainage of <span class="hlt">Glacier</span> National <span class="hlt">Park</span> contained subalpine fir forests that graded into alpine tundra. Over a 46-yr period, altitudinal changes in the location of alpine treeline ecotone were not observed. However, over this 46-yr period the area of krummholz, patch-forest, and continuous canopy forest increased by 3.4%, and tree density increased within existing patches of krummholz and patch-forest. Change in subalpine fir vegetation patterns within 100 m of trails was also compared to areas without trails. Within 100 m of trails, the number of small, discrete krummholz stands increased compared to areas without trails, but there was no significant change in total krummholz area. We used historical terrestrial photography to expand the period (to 70 yr) considered. This photography supported the conclusions that a more abrupt ecotone transition developed from forest to tundra at alpine treeline, that tree density within forested areas increased, and that krummholz became fragmented along trails. This local assessment of fine-grained change in the alpine treeline ecotone provides a comparative base for looking at ecotone change in other mountain regions throughout the world.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C12B..06B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C12B..06B"><span>North Atlantic Oscillation Drives Regional Greenland <span class="hlt">Glacier</span> Volume During the 20th Century</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bjork, A. A.; Aagaard, S.; Hallander, A. M.; Khan, S. A.; Box, J. E.; Kjeldsen, K. K.; Larsen, N. K.; Korsgaard, N. J.; Cappelen, J.; Colgan, W. T.; Machguth, H.; Andresen, C. S.; Kjaer, K. H.</p> <p>2016-12-01</p> <p>While most areas of the Greenland ice sheet have undergone rapid mass loss since c. 1990, the central eastern section of the ice sheet has advanced and gained mass. This contrasting regional trend has been attributed to positive surface mass balance (SMB) in the absence of significant dynamic mass loss. To constrain the atypical behavior in this region, we mapped <span class="hlt">glacier</span> length fluctuations of nearly 200 peripheral <span class="hlt">glaciers</span> and ice caps (PGICs) over a 103-year period, and compare the results with c. 150 new <span class="hlt">glacier</span> length records from central west Greenland. We demonstrate that the regional response in ice volume is closely correlated to changes in precipitation, governed by circulation patterns associated with the North Atlantic Oscillation (NAO) and secondarily influenced by temperature forcing in certain periods. More broadly, we find that the NAO contributes to contrasting precipitation variability in <span class="hlt">East</span> and West Greenland, where it appears to be responsible for at least 10% and more than 25%, respectively, of the variability in ice sheet accumulation rate. This <span class="hlt">east</span>-west asymmetry, which influences both LGICs and the ice sheet, illustrates how substantial uncertainty in NAO projections directly contributes to uncertainty in mass balance projections.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/48392','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/48392"><span>Application of the MAGIC model to the <span class="hlt">Glacier</span> Lakes catchments</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>John O. Reuss</p> <p>1994-01-01</p> <p>The MAGIC model (Cosby et al. 1985, 1986) was calibrated for <span class="hlt">East</span> and West <span class="hlt">Glacier</span> Lakes, two adjacent high-altitude (3200 m- 3700 m) catchments in the Medicine Bow National Forest of southern Wyoming. This model uses catchment characteristics including weathering rates, soil chemical characteristics, hydrological parameters, and precipitation amounts and composition...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.C13B0432L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.C13B0432L"><span>Middle to late Holocene fluctuations of the Vindue <span class="hlt">glacier</span>, an outlet <span class="hlt">glacier</span> of the Greenland Ice Sheet, central <span class="hlt">East</span> Greenland.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Levy, L.; Hammer, S. K.; Kelly, M. A.; Lowell, T. V.; Hall, B. L.; Howley, J. A.; Wilcox, P.; Medford, A.</p> <p>2014-12-01</p> <p>The margins of the Greenland Ice Sheet are currently responding to present-day climate changes. Determining how the ice sheet margins have responded to past climate changes provides a means to understand how they may respond in the future. Here we present a multi-proxy record used to reconstruct the Holocene fluctuations of the Vindue <span class="hlt">glacier</span>, an ice sheet outlet <span class="hlt">glacier</span> in eastern Greenland. Lake sediment cores from Qiviut lake (informal name), located ~0.75 km from the present-day Vindue <span class="hlt">glacier</span> margin contain a sharp transition from medium sand/coarse silt to laminated gyttja just prior to 6,340±130 cal yr BP. We interpret this transition to indicate a time when the Vindue <span class="hlt">glacier</span> retreated sufficiently to cease glacial sedimentation into the lake basin. Above this contact the core contains laminated gyttja with prominent, ~0.5 cm thick, silt layers. 10Be ages of boulders on bedrock located between Qiviut lake and the present-day ice margin date to 6.81 ± 0.67 ka (n = 3), indicating the time of deglaciation. These ages also agree well with the radiocarbon age of the silt-gyttja transition in Qiviut lake cores. 10Be ages on boulders on bedrock located more proximal to the ice margin (~0.5 km) yield ages of 2.67 ± 0.18 ka (n = 2). These ages indicate either the continued recession of the ice margin during the late Holocene or an advance at this time. Boulders on the historical moraines show that ice retreated from the moraine by AD 1620 ± 20 yrs (n = 2). These results are in contrast with some areas of the western margin of the ice sheet where 10Be ages indicate that the ice sheet was behind its Historical limit from the middle Holocene (~6-7 ka) to Historical time. This may indicate that the eastern margin may have responded to late Holocene cooling more sensitively or that the advance associated with the Historical moraines overran any evidence of late Holocene fluctuations along the western margin of the ice sheet.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca1656.photos.190898p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca1656.photos.190898p/"><span>2. HORSESHOE CURVE IN <span class="hlt">GLACIER</span> POINT ROAD NEAR <span class="hlt">GLACIER</span> POINT. ...</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. HORSESHOE CURVE IN <span class="hlt">GLACIER</span> POINT ROAD NEAR <span class="hlt">GLACIER</span> POINT. HALF DOME AT CENTER REAR. LOOKING NNE. GIS N-37 43 44.3 / W-119 34 14.1 - <span class="hlt">Glacier</span> Point Road, Between Chinquapin Flat & <span class="hlt">Glacier</span> Point, Yosemite Village, Mariposa County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca1656.color.218148c/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca1656.color.218148c/"><span>HORSESHOE CURVE IN <span class="hlt">GLACIER</span> POINT ROAD NEAR <span class="hlt">GLACIER</span> POINT. HALF ...</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>HORSESHOE CURVE IN <span class="hlt">GLACIER</span> POINT ROAD NEAR <span class="hlt">GLACIER</span> POINT. HALF DOME AT CENTER REAR. SAME VIEW AT CA-157-2. LOOKING NNE. GIS: N-37' 43 44.3 / W-119 34 14.1 - <span class="hlt">Glacier</span> Point Road, Between Chinquapin Flat & <span class="hlt">Glacier</span> Point, Yosemite Village, Mariposa County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.4931H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.4931H"><span>A 70-year record of outlet <span class="hlt">glacier</span> retreat in northern Greenland</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hill, Emily; Carr, Rachel; Stokes, Chris; Gudmundsson, Hilmar</p> <p>2017-04-01</p> <p>Over the past two decades, the Greenland Ice Sheet (GrIS) has undergone accelerated mass loss increasing its contribution to sea level rise. This is partly attributed to increased mass loss from dynamic marine-terminating outlet <span class="hlt">glaciers</span>. Despite marine-terminating outlet <span class="hlt">glaciers</span> in northern Greenland draining 40% of the ice sheet by area, they are comparatively less well-studied than other regions of the ice sheet (e.g. central west or south-<span class="hlt">east</span>). This region could be susceptible to marine-ice sheet instability due to large proportions of the bedrock rested below sea level and is also unique in the presence of large floating ice tongues. Here, we use a range of satellite imagery sources, accompanied by historical maps, to examine multi-decadal front position changes at 21 outlet <span class="hlt">glaciers</span> in northern Greenland between 1948 and 2016. We accompany these terminus changes, with annual records of ice velocity, climate-ocean forcing data, and <span class="hlt">glacier</span>-specific factors (e.g. fjord-width and basal topography) to understand the dominant forcing on <span class="hlt">glacier</span> dynamics in the region. Over the last 70 years, there has been a clear pattern of <span class="hlt">glacier</span> retreat in northern Greenland. This is particularly notable during the last two decades, where 62% of our study <span class="hlt">glaciers</span> showed accelerated retreat. This was most notable at Humboldt, Tracy, Hagen Brae, C. H. Ostenfeld and Petermann <span class="hlt">Glaciers</span>, and in the case of the latter three <span class="hlt">glaciers</span>, this involved substantial retreat of their floating ice tongues (> 10 km). Alongside retreat, several study <span class="hlt">glaciers</span> underwent simultaneous velocity increases. However, the collapse of floating ice tongues did not always result in increased velocity. Similar to other regions of the ice sheet, recent <span class="hlt">glacier</span> retreat in the northern regions of the Greenland Ice Sheet could be linked to climatic-oceanic forcing, but at this stage this remains largely unknown. This response to external forcing is further complicated by the presence of <span class="hlt">glacier</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFMNS22A..03S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFMNS22A..03S"><span>Afghanistan <span class="hlt">Glacier</span> Diminution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shroder, J. F.; Bishop, M.; Haritashya, U.; Olsenholler, J.</p> <p>2008-12-01</p> <p><span class="hlt">Glaciers</span> in Afghanistan represent a late summer - early fall source of melt water for late season crop irrigation in a chronically drought-torn region. Precise river discharge figures associated with <span class="hlt">glacierized</span> drainage basins are generally unavailable because of the destruction of hydrological gauging stations built in pre-war times although historic discharge data and prior (1960s) mapped <span class="hlt">glacier</span> regions offer some analytical possibilities. The best satellite data sets for <span class="hlt">glacier</span>-change detection are declassified Cornona and Keyhole satellite data sets, standard Landsat sources, and new ASTER images assessed in our GLIMS (Global Land Ice Measurements from Space) Regional Center for Southwest Asia (Afghanistan and Pakistan). The new hyperspectral remote sensing survey of Afghanistan completed by the US Geological Survey and the Afghanistan Ministry of Mines offers potential for future detailed assessments. Long-term climate change in southwest Asia has decreased precipitation for millennia so that <span class="hlt">glaciers</span>, rivers and lakes have all declined from prehistoric and historic highs. As many <span class="hlt">glaciers</span> declined in ice volume, they increased in debris cover until they were entirely debris-covered or became rock <span class="hlt">glaciers</span>, and the ice was protected thereby from direct solar radiation, to presumably reduce ablation rates. We have made a preliminary assessment of <span class="hlt">glacier</span> location and extent for the country, with selected, more-detailed, higher-resolution studies underway. In the Great Pamir of the Wakhan Corridor where the largest <span class="hlt">glaciers</span> occur, we assessed fluctuations of a randomly selected 30 <span class="hlt">glaciers</span> from 1976 to 2003. Results indicate that 28 <span class="hlt">glacier</span>-terminus positions have retreated, and the largest average retreat rate was 36 m/yr. High albedo, non-vegetated <span class="hlt">glacier</span> forefields formed prior to 1976, and geomorphological evidence shows apparent <span class="hlt">glacier</span>-surface downwasting after 1976. Climatic conditions and <span class="hlt">glacier</span> retreat have resulted in disconnection of tributary</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/29498','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/29498"><span>Developing a framework for evaluating proposals for research in wilderness: Science to protect and learn from <span class="hlt">parks</span></span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Lewis C. Sharman; Peter Landres; Susan Boudreau</p> <p>2007-01-01</p> <p>In designated <span class="hlt">park</span> wilderness, the requirements for scientific research often conflict with requirements designed to protect wilderness resources and values. Managers who wish to realize the benefits of scientific research must have a process by which to evaluate those benefits as well as their associated wilderness impacts. <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, in...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002AGUFM.C61A..05H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002AGUFM.C61A..05H"><span><span class="hlt">Glacier</span> Instability, Rapid <span class="hlt">Glacier</span> Lake Growth and Related Hazards at Belvedere <span class="hlt">Glacier</span>, Macugnaga, Italy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huggel, C.; Kaeaeb, A.; Haeberli, W.; Mortara, G.; Chiarle, M.; Epifani, F.</p> <p>2002-12-01</p> <p>Starting in summer 2000, Belvedere <span class="hlt">Glacier</span>, near Macugnaga, Italian Alps, developed an extraordinary change in flow, geometry and surface appearance. A surge-type flow acceleration started in the lower parts of the Monte-Rosa <span class="hlt">east</span> face, leading to strong crevassing and deformation of Belvedere <span class="hlt">Glacier</span>, accompanied by bulging of its orographic right margin. In September 2001, a small supraglacial lake developed on the <span class="hlt">glacier</span>. High water pressure and accelerated movement lasted into winter 2001/2002. The ice, in places, started to override moraines from the Little Ice Age. In late spring and early summer 2002, the supraglacial lake grew at extraordinary rates reaching a maximum area of more than 150'000 m2 by end of June. The evolution of such a large supraglacial lake, a rather unique feature in the Alps, was probably enabled by changes in the subglacial drainage system in the course of the surge-like developments with high water pressure in the <span class="hlt">glacier</span>. At the end of June, an enhanced growth of the lake level with a rise of about 1 m per day was observed such that the supraglacial lake became a urgent hazard problem for the community of Macugnaga. Emergency measures had to be taken by the Italian Civil Protection. The authors thereby acted as the official expert advisers. Temporal evacuations were ordered and a permanent monitoring and alarm system was installed. Pumps with a maximum output of 1 m3/s were brought to the lake. Bathymetric studies yielded a maximum lake depth of 55 m and a volume of 3.3 millions of cubic meters of water. Aerial photography of 1995, 1999, September 2001 and October 2001 was used to calculate ice flow velocities and changes in surface altitude. Compared to the period of 1995 to 1999, the flow accelerated by about five times in 2001 (max. speeds up to 200 m/yr). Surface uplift measured was about 10-15 m/yr. The results of the photogrammetric studies were used to evaluate different possible lake-outburst scenarios, in particular</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA03386.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA03386.html"><span>Malaspina <span class="hlt">Glacier</span>, Alaska</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2003-05-01</p> <p>Malaspina <span class="hlt">Glacier</span> in southeastern Alaska is considered the classic example of a piedmont <span class="hlt">glacier</span>. Piedmont <span class="hlt">glaciers</span> occur where valley <span class="hlt">glaciers</span> exit a mountain range onto broad lowlands, are no longer laterally confined, and spread to become wide lobes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C51D..01H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C51D..01H"><span>How do <span class="hlt">glacier</span> inventory data aid global <span class="hlt">glacier</span> assessments and projections?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hock, R.</p> <p>2017-12-01</p> <p>Large-scale <span class="hlt">glacier</span> modeling relies heavily on datasets that are collected by many individuals across the globe, but managed and maintained in a coordinated fashion by international data centers. The Global Terrestrial Network for <span class="hlt">Glaciers</span> (GTN-G) provides the framework for coordinating and making available a suite of data sets such as the Randolph <span class="hlt">Glacier</span> Inventory (RGI), the <span class="hlt">Glacier</span> Thickness Dataset or the World <span class="hlt">Glacier</span> Inventory (WGI). These datasets have greatly increased our ability to assess global-scale <span class="hlt">glacier</span> mass changes. These data have also been vital for projecting the <span class="hlt">glacier</span> mass changes of all mountain <span class="hlt">glaciers</span> in the world outside the Greenland and Antarctic ice sheet, a total >200,000 <span class="hlt">glaciers</span> covering an area of more than 700,000 km2. Using forcing from 8 to 15 GCMs and 4 different emission scenarios, global-scale <span class="hlt">glacier</span> evolution models project multi-model mean net mass losses of all <span class="hlt">glaciers</span> between 7 cm and 24 cm sea-level equivalent by the end of the 21st century. Projected mass losses vary greatly depending on the choice of the forcing climate and emission scenario. Insufficiently constrained model parameters likely are an important reason for large differences found among these studies even when forced by the same emission scenario, especially on regional scales.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70041040','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70041040"><span>Listening to <span class="hlt">Glaciers</span>: Passive hydroacoustics near marine-terminating <span class="hlt">glaciers</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>Pettit, E.C.; Nystuen, J.A.; O'Neel, Shad</p> <p>2012-01-01</p> <p>The catastrophic breakup of the Larsen B Ice Shelf in the Weddell Sea in 2002 paints a vivid portrait of the effects of <span class="hlt">glacier</span>-climate interactions. This event, along with other unexpected episodes of rapid mass loss from marine-terminating <span class="hlt">glaciers</span> (i.e., tidewater <span class="hlt">glaciers</span>, outlet <span class="hlt">glaciers</span>, ice streams, ice shelves) sparked intensified study of the boundaries where marine-terminating <span class="hlt">glaciers</span> interact with the ocean. These dynamic and dangerous boundaries require creative methods of observation and measurement. Toward this effort, we take advantage of the exceptional sound-propagating properties of seawater to record and interpret sounds generated at these glacial ice-ocean boundaries from distances safe for instrument deployment and operation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C41C1237P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C41C1237P"><span>Sensitivity of Totten <span class="hlt">Glacier</span> Ice Shelf extent and grounding line to oceanic forcing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pelle, T.; Morlighem, M.; Choi, Y.</p> <p>2017-12-01</p> <p>Totten <span class="hlt">Glacier</span> is a major outlet <span class="hlt">glacier</span> of the <span class="hlt">East</span> Antarctic Ice Sheet and has been shown to be vulnerable to ocean-induced melt in both its past and present states. The intrusion of warm, circumpolar deep water beneath the Totten <span class="hlt">Glacier</span> Ice Shelf (TGIS) has been observed to accelerate ice shelf thinning and promote iceberg calving, a primary mechanism of mass discharge from Totten. As such, accurately simulating TGIS's ice front dynamics is crucial to the predictive capabilities of ice sheet models in this region. Here, we study the TGIS using the Ice Sheet System Model (ISSM) and test the applicability of three calving laws: Crevasse Formation calving, Eigen calving, and Tensile Stress calving. We simulate the evolution of Totten <span class="hlt">Glacier</span> through 2100 under enhanced oceanic forcing in order to investigate both future changes in ice front dynamics and possible thresholds of instability. In addition, we artificially retreat Totten's ice front position and allow the model to proceed dynamically in order to analyze the response of the <span class="hlt">glacier</span> to calving events. Our analyses show that Tensile Stress calving most accurately reproduces Totten <span class="hlt">Glacier</span>'s observed ice front position. Furthermore, unstable grounding line retreat is projected when Totten is simulated under stronger oceanic thermal forcing scenarios and when the calving front is significantly retreated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AAS...21715815N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AAS...21715815N"><span>Stars Above, Earth Below: Astronomy in the National <span class="hlt">Parks</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nordgren, Tyler E.</p> <p>2011-01-01</p> <p>The U.S. national <span class="hlt">parks</span> that protect our enjoyment of the landscape around us by day, also protect our enjoyment of the sky above at night. With the growth of light pollution, the view of the stars and Milky Way overhead has become as rare as the views of <span class="hlt">glaciers</span>, geysers, and grizzlies that bring millions of visitors to the <span class="hlt">parks</span> every year. Through the pristine view of a starry sky at night <span class="hlt">park</span> visitors are primed to learn about our planet, its place in the solar system, and the larger Universe in which we live. The national <span class="hlt">parks</span> are therefore the largest informal educational setting for reaching millions of people from all over the world who might not otherwise encounter astronomical outreach. The material in this presentation has been field tested in national <span class="hlt">parks</span>, campgrounds, lodges, and visitor centers over the last four years and is elaborated on in the just released book: "Stars Above, Earth Below: A Guide to Astronomy in the National <span class="hlt">Parks</span>.” Funding for this project was provided by The Planetary Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://arc.lib.montana.edu/snow-science/item/1138','USGSPUBS'); return false;" href="http://arc.lib.montana.edu/snow-science/item/1138"><span>Forecasting for natural avalanches during spring opening of Going-to-the-Sun Road, <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Reardon, Blase; Lundy, Chris</p> <p>2004-01-01</p> <p>The annual spring opening of the Going-to-the-Sun Road in <span class="hlt">Glacier</span> National <span class="hlt">Park</span> presents a unique avalanche forecasting challenge. The highway traverses dozens of avalanche paths mid-track in a 23-kilometer section that crosses the Continental Divide. Workers removing seasonal snow and avalanche debris are exposed to paths that can produce avalanches of destructive class 4. The starting zones for most slide paths are within proposed Wilderness, and explosive testing or control are not currently used. Spring weather along the Divide is highly variable; rain-on-snow events are common, storms can bring several feet of new snow as late as June, and temperature swings can be dramatic. Natural avalanches - dry and wet slab, dry and wet loose, and glide avalanches - present a wide range of hazards and forecasting issues. This paper summarizes the forecasting program instituted in 2002 for the annual snow removal operations. It focuses on tools and techniques for forecasting natural wet snow avalanches by incorporating two case studies, including a widespread climax wet slab cycle in 2003. We examine weather and snowpack conditions conducive to wet snow avalanches, indicators for instability, and suggest a conceptual model for wet snow stability in a northern intermountain snow climate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Geomo.296..142J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Geomo.296..142J"><span>An inventory and estimate of water stored in firn fields, <span class="hlt">glaciers</span>, debris-covered <span class="hlt">glaciers</span>, and rock <span class="hlt">glaciers</span> in the Aconcagua River Basin, Chile</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Janke, Jason R.; Ng, Sam; Bellisario, Antonio</p> <p>2017-11-01</p> <p>An inventory of firn fields, <span class="hlt">glaciers</span>, debris-covered <span class="hlt">glaciers</span>, and rock <span class="hlt">glaciers</span> was conducted in the Aconcagua River Basin of the semiarid Andes of central Chile. A total of 916 landforms were identified, of which rock <span class="hlt">glaciers</span> were the most abundant (669) and occupied the most total area. <span class="hlt">Glaciers</span> and debris-covered <span class="hlt">glaciers</span> were less numerous, but were about five times larger in comparison. The total area occupied by <span class="hlt">glaciers</span> and debris-covered <span class="hlt">glaciers</span> was roughly equivalent to the total area of rock <span class="hlt">glaciers</span>. Debris-covered <span class="hlt">glaciers</span> and rock <span class="hlt">glaciers</span> were subcategorized into six ice-content classes based on interpretation of surface morphology with high-resolution satellite imagery. Over 50% of rock <span class="hlt">glaciers</span> fell within a transitional stage; 85% of debris-covered <span class="hlt">glaciers</span> were either fully covered or buried. Most landforms occupied elevations between 3500 and 4500 m. <span class="hlt">Glaciers</span> and firn occurred at higher elevations compared to rock <span class="hlt">glaciers</span> and debris-covered <span class="hlt">glaciers</span>. Rock <span class="hlt">glaciers</span> had a greater frequency in the northern part of the study area where arid climate conditions exist. Firn and <span class="hlt">glaciers</span> were oriented south, debris-covered <span class="hlt">glaciers</span> west, and rock <span class="hlt">glaciers</span> southwest. An analysis of water contribution of each landform in the upper Andes of the Aconcagua River Basin was conducted using formulas that associate the size of the landforms to estimates of water stored. Minimum and maximum water storage was calculated based on a range of debris to ice content ratios for debris-covered <span class="hlt">glaciers</span> and rock <span class="hlt">glaciers</span>. In the Aconcagua River Basin, rock <span class="hlt">glaciers</span> accounted for 48 to 64% of the water stored within the landforms analyzed; <span class="hlt">glaciers</span> accounted for 15 to 25%; debris-covered <span class="hlt">glaciers</span> were estimated at 15 to 19%; firn fields contained only about 5 to 8% of the water stored. Expansion of agriculture, prolonged drought, and removal of ice-rich landforms for mining have put additional pressure on already scarce water resources. To develop long</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/1386i/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/1386i/report.pdf"><span><span class="hlt">Glaciers</span> of South America</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Williams, Richard S.; Ferrigno, Jane G.</p> <p>1998-01-01</p> <p>Landsat images, together with maps and aerial photographs, have been used to produce <span class="hlt">glacier</span> inventories, define <span class="hlt">glacier</span> locations, and study <span class="hlt">glacier</span> dynamics in the countries of South America, along with the Andes Mountains. In Venezuela, Colombia, Ecuador, and Bolivia, the small <span class="hlt">glaciers</span> have been undergoing extensive <span class="hlt">glacier</span> recession since the late 1800's. <span class="hlt">Glacier</span>-related hazards (outburst floods, mud flows, and debris avalanches) occur in Colombia, in Ecuador, and associated with the more extensive (2,600 km2) <span class="hlt">glaciers</span> of Peru. The largest area of <span class="hlt">glacier</span> ice is found in Argentina and Chile, including the northern Patagonian ice field (about 4,200 km2) and the southern Patagonian ice field (about 13,000 km2), the largest <span class="hlt">glacier</span> in the Southern Hemisphere outside Antarctica.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GPC...160..123J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GPC...160..123J"><span>The distribution and hydrological significance of rock <span class="hlt">glaciers</span> in the Nepalese Himalaya</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jones, D. B.; Harrison, S.; Anderson, K.; Selley, H. L.; Wood, J. L.; Betts, R. A.</p> <p>2018-01-01</p> <p>In the Nepalese Himalaya, there is little information on the number, spatial distribution and morphometric characteristics of rock <span class="hlt">glaciers</span>, and this information is required if their hydrological contribution is to be understood. Based on freely available fine spatial resolution satellite data accessible through Google Earth, we produced the first comprehensive Nepalese rock <span class="hlt">glacier</span> inventory, supported through statistical validation and field survey. The inventory includes the location of over 6000 rock <span class="hlt">glaciers</span>, with a mean specific density of 3.4%. This corresponds to an areal coverage of 1371 km2. Our approach subsampled approximately 20% of the total identified rock <span class="hlt">glacier</span> inventory (n = 1137) and digitised their outlines so that quantitative/qualitative landform attributes could be extracted. Intact landforms (containing ice) accounted for 68% of the subsample, and the remaining were classified as relict (not containing ice). The majority (56%) were found to have a northerly aspect (NE, N, and NW), and landforms situated within north- to west-aspects reside at lower elevations than those with south- to- <span class="hlt">east</span> aspects. In Nepal, we show that rock <span class="hlt">glaciers</span> are situated between 3225 and 5675 m a.s.l., with the mean minimum elevation at the front estimated to be 4977 ± 280 m a.s.l. for intact landforms and 4541 ± 346 m a.s.l. for relict landforms. The hydrological significance of rock <span class="hlt">glaciers</span> in Nepal was then established by statistically upscaling the results from the subsample to estimate that these cryospheric reserves store between 16.72 and 25.08 billion m3 of water. This study, for the first time, estimates rock <span class="hlt">glacier</span> water volume equivalents and evaluates their relative hydrological importance in comparison to ice <span class="hlt">glaciers</span>. Across the Nepalese Himalaya, rock <span class="hlt">glacier</span> to ice <span class="hlt">glacier</span> water volume equivalent is 1:9, and generally increases westwards (e.g., ratio = 1:3, West region). This inventory represents a preliminary step for understanding the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/dc0657.photos.028057p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/dc0657.photos.028057p/"><span>2. D Street facade and rear (<span class="hlt">east</span>) blank wall 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>2. D Street facade and rear (<span class="hlt">east</span>) blank wall of <span class="hlt">parking</span> garage. Farther <span class="hlt">east</span> is 408 8th Street (National Art And Frame Company). - PMI <span class="hlt">Parking</span> Garage, 403-407 Ninth Street, Northwest, Washington, District of Columbia, DC</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('http://adsabs.harvard.edu/abs/2014EGUGA..16.6254L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16.6254L"><span>Reconstruction of late Holocene <span class="hlt">glacier</span> retreat and relevant climatic and topographic patterns in southeastern Tibet by <span class="hlt">glacier</span> mapping and equilibrium line altitude calculation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loibl, David; Lehmkuhl, Frank</p> <p>2014-05-01</p> <p> limitation, and detachment of <span class="hlt">glacier</span> tributaries, are criteria that prevent reliable ELA calculations. In order to determine the best-fitting TRAM ratio value and to test the quality of the calculated ELAs, a remote sensing approach was applied: the altitudes of transient snowlines visible in the late summer Landsat scene were measured from the DEM and compared to TRAM results for each <span class="hlt">glacier</span>. The interpolated ELA results show a southeast-northwest gradient ranging from 4,400 to 5,600 m a.s.l. and an average ELA rise of ~ 98 m since the LIA. Due to the large amount of measurements, the ELA distribution reveals topographic effects down to the catchment scale, i.e. orographic rainfalls and leeward shielding. Contrasting to the expectations for subtropical settings, <span class="hlt">glaciers</span> on south facing slopes have not retreated strongest and ELAs on south facing slopes did not rise furthest. Instead, highly heterogeneous spatial patterns emerge that show a strong imprint of both, topography and monsoonal dynamics. The interpretation of these patterns provides insights into the monsoonal system and the characteristics of late Holocene <span class="hlt">glacier</span> change in southeastern Tibet. For example, the ELA distribution reveals that the study area is influenced by both, Indian summer monsoon and <span class="hlt">East</span> Asian summer monsoon, but that the latter does not reach the Tibetan Plateau.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014Geomo.210...59C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014Geomo.210...59C"><span>Evolution of <span class="hlt">glacier</span>-dammed lakes through space and time; Brady <span class="hlt">Glacier</span>, Alaska, USA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Capps, Denny M.; Clague, John J.</p> <p>2014-04-01</p> <p><span class="hlt">Glacier</span>-dammed lakes and their associated jökulhlaups cause severe flooding in downstream areas and substantially influence <span class="hlt">glacier</span> dynamics. Brady <span class="hlt">Glacier</span> in southeast Alaska is well suited for a study of these phenomena because it presently dams 10 large (> 1 km2) lakes. Our objectives are to demonstrate how Brady <span class="hlt">Glacier</span> and its lakes have co-evolved in the past and to apply this knowledge to predict how the <span class="hlt">glacier</span> and its lakes will likely evolve in the future. To accomplish these objectives, we georeferenced a variety of maps, airphotos, and optical satellite imagery to characterize the evolution of the <span class="hlt">glacier</span> and lakes. We also collected bathymetry data and created bathymetric maps of select lakes. Despite small advances and retreats, the main terminus of Brady <span class="hlt">Glacier</span> has changed little since 1880. However, it downwasted at rates of 2-3 m/y between 1948 and 2000, more than the regional average. The most dramatic retreat (2 km) and downwasting (120 m) have occurred adjacent to <span class="hlt">glacier</span>-dammed lakes and are primarily the result of calving. Brady <span class="hlt">Glacier</span> is a former tidewater <span class="hlt">glacier</span>. With continued downwasting, Brady <span class="hlt">Glacier</span> may return to a tidewater regime and enter into a phase of catastrophic retreat. The situation at Brady <span class="hlt">Glacier</span> is not unique, and the lessons learned here can be applied elsewhere to identify future <span class="hlt">glacier</span>-dammed lakes, jökulhlaups, and <span class="hlt">glacier</span> instability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70160237','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70160237"><span>Long-duration drought variability and impacts on ecosystem services: A case study from <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pederson, Gregory T.; Gray, Stephen T.; Fagre, Daniel B.; Graumlich, Lisa J.</p> <p>2006-01-01</p> <p>Using a suite of paleoproxy reconstructions and information from previous studies examining the relationship between climate variability and natural processes, the authors explore how such persistent moisture anomalies affect the delivery of vital goods and services provided by <span class="hlt">Glacier</span> NP and surrounding areas. These analyses show that regional water resources and tourism are particularly vulnerable to persistent moisture anomalies in the <span class="hlt">Glacier</span> NP area. Many of these same decadal-scale wet and dry events were also seen among a wider network of hydroclimatic reconstructions along a north–south transect of the Rocky Mountains. Such natural climate variability can, in turn, have enormous impacts on the sustainable provision of natural resources over wide areas. Overall, these results highlight the susceptibility of goods and services provided by protected areas like <span class="hlt">Glacier</span> NP to natural climate variability, and show that this susceptibility will likely be compounded by the effects of future human-induced climate change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009EGUGA..11.7719B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009EGUGA..11.7719B"><span>The CARIPANDA project: Climate change and water resources in the Adamello Natural <span class="hlt">Park</span> of Italy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bocchiola, D.</p> <p>2009-04-01</p> <p>The three years (2007-2009) CARIPANDA project funded by the Cariplo Foundation of Italy is aimed to evaluate scenarios for water resources in the Adamello natural <span class="hlt">Park</span> of Italy in a window of 50 years or so (until 2050). The project is led by Ente Parco Adamello and involves Politecnico di Milano, Università Statale di Milano, Università di Brescia, and ARPA Lombardia as scientific partners, while ENEL hydropower Company of Italy joins the project as stake holder. The Adamello Natural <span class="hlt">Park</span> is a noteworthy resource in the Italian Alps. The Adamello Group is made of several <span class="hlt">glacierized</span> areas (c. 24 km2), of both debris covered and free ice types, including the widest Italian <span class="hlt">Glacier</span>, named Adamello, spreading on an area of about c. 18 km2. Also the Adamello Natural Reserve, covering 217 km2 inside the Adamello <span class="hlt">Park</span> and including the Adamello <span class="hlt">glaciers</span>, hosts a number of high altitude safeguarded vegetal and animal species, the safety of which is a primary task of the Reserve. Project's activity involves analysis of local climate trend, field campaigns on <span class="hlt">glaciers</span>, hydrological modelling and remote sensing of snow and ice covered areas, aimed to build a consistent model of the present hydrological conditions and of the areas. Then, properly tailored climate change projections for the area, obtained using local data driven downscaling of climate change projections from GCMs model, are used to infer the likely response to expected climate change conditions. With two years in the project now some preliminary findings can be highlighted and some preliminary trend analysis carried out. The proposed poster provides a resume of the main results of the project insofar, of interest as a benchmark for similar ongoing and foregoing projects about climate change impact on European mountainous natural areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.C23C0510A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.C23C0510A"><span>Automated <span class="hlt">Glacier</span> Surface Velocity using Multi-Image/Multi-Chip (MIMC) Feature Tracking</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ahn, Y.; Howat, I. M.</p> <p>2009-12-01</p> <p>Remote sensing from space has enabled effective monitoring of remote and inhospitable polar regions. <span class="hlt">Glacier</span> velocity, and its variation in time, is one of the most important parameters needed to understand <span class="hlt">glacier</span> dynamics, <span class="hlt">glacier</span> mass balance and contribution to sea level rise. Regular measurements of ice velocity are possible from large and accessible satellite data set archives, such as ASTER and LANDSAT-7. Among satellite imagery, optical imagery (i.e. passive, visible to near-infrared band sensors) provides abundant data with optimal spatial resolution and repeat interval for tracking <span class="hlt">glacier</span> motion at high temporal resolution. Due to massive amounts of data, computation of ice velocity from feature tracking requires 1) user-friendly interface, 2) minimum local/user parameter inputs and 3) results that need minimum editing. We focus on robust feature tracking, applicable to all currently available optical satellite imagery, that is ASTER, SPOT and LANDSAT etc. We introduce the MIMC (multiple images/multiple chip sizes) matching approach that does not involve any user defined local/empirical parameters except approximate average <span class="hlt">glacier</span> speed. We also introduce a method for extracting velocity from LANDSAT-7 SLC-off data, which has 22 percent of scene data missing in slanted strips due to failure of the scan line corrector. We apply our approach to major outlet <span class="hlt">glaciers</span> in west/<span class="hlt">east</span> Greenland and assess our MIMC feature tracking technique by comparison with conventional correlation matching and other methods (e.g. InSAR).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2007/5278/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2007/5278/"><span>Distribution of ground-nesting marine birds along shorelines in <span class="hlt">Glacier</span> Bay, southeastern Alaska: An assessment related to potential disturbance by back-country users</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Arimitsu, Mayumi L.; Piatt, John F.; Romano, Marc D.</p> <p>2007-01-01</p> <p>With the exception of a few large colonies, the distribution of ground-nesting marine birds in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> in southeastern Alaska is largely unknown. As visitor use increases in back-country areas of the <span class="hlt">park</span>, there is growing concern over the potential impact of human activities on breeding birds. During the 2003–05 breeding seasons, the shoreline of <span class="hlt">Glacier</span> Bay was surveyed to locate ground-nesting marine birds and their nesting areas, including wildlife closures and historical sites for egg collection by Alaska Native peoples. The nesting distribution of four common ground-nesting marine bird species was determined: Arctic Tern (Sterna paradisaea), Black Oystercatcher (Haematopus bachmani), Mew Gull (Larus canus), and Glaucous-winged Gull (Larus glaucescens). Observations of less abundant species also were recorded, including Herring Gull (Larus argentatus), Red-throated Loon (Gavia stellata), Canada Goose (Branta canadensis), Willow Ptarmigan (Lagopus lagopus), Semipalmated Plover (Charadrius semipalmatus), Spotted Sandpiper (Actitis macularia), Least Sandpiper (Calidris minutilla), Parasitic Jaeger (Stercorarius parasiticus), and Aleutian Tern (Sterna aleutica). Nesting distribution for Arctic Terns was largely restricted to the upper arms of the bay and a few treeless islets in the lower bay, whereas Black Oystercatchers were more widely distributed along shorelines in the <span class="hlt">park</span>. Mew Gulls nested throughout the upper bay in Geikie Inlet and in Fingers and Berg Bays, and most Glaucous-winged Gull nests were found at wildlife closures in the central and lower bays. Several areas were identified where human disturbance could affect breeding birds. This study comprises the first bay-wide survey for the breeding distribution of ground-nesting marine birds in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, providing a minimum estimate of their numbers and distribution within the <span class="hlt">park</span>. This information can be used to assess future human disturbance and track natural</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1812583P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1812583P"><span><span class="hlt">Glacier</span>-derived climate for the Younger Dryas in Europe</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pellitero, Ramon; Rea, Brice R.; Spagnolo, Matteo; Hughes, Philip; Braithwaite, Roger; Renssen, Hans; Ivy-Ochs, Susan; Ribolini, Adriano; Bakke, Jostein; Lukas, Sven</p> <p>2016-04-01</p> <p>We have reconstructed and calculated the <span class="hlt">glacier</span> equilibrium line altitudes (ELA) for 120 Younger Dryas palaeoglaciers from Morocco in the south to Svalbard in the north and from Ireland in the west to Turkey in the <span class="hlt">east</span>. The chronology of these landform were checked and, when derived from cosmogenic dates, these were recalculated based on newer production rates. Frontal moraines/limits for the palaeoglaciers were used to reconstruct palaeoglacier extent by using a GIS tool which implements a discretised solution for the assumption of perfect-plasticity ice rheology for a single flowline and extents this out to a 3D ice surface. From the resulting equilibrium profile, palaeoglaciers palaeo-ELAs were calculated using another GIS tool. Where several <span class="hlt">glaciers</span> were reconstructed in a region, a single ELA value was generated following the methodology of Osmaston (2005). In order to utilise these ELAs for quantitative palaeo-precipitation reconstructions an independent regional temperature analysis was undertaken. A database of 121 sites was compiled where the temperature was determined from palaeoproxies other than <span class="hlt">glaciers</span> (e.g. pollen, diatoms, choleoptera, chironimids…) in both terrestrial and offshore environments. These proxy data provides estimates of average annual, summer and winter temperatures. These data were merged and interpolated to generate maps of average temperature for the warmest and coldest months and annual average temperature. From these maps the temperature at the ELA was obtained using a lapse rate of 0.65°C/100m. Using the ELA temperature range and summer maximum in a degree-day model allows determination of the potential melt which can be taken as equivalent to precipitation given the assumption a <span class="hlt">glacier</span> is in equilibrium with climate. Results show that during the coldest part of the Younger Dryas precipitation was high in the British Isles, the NW of the Iberian Peninsula and the Vosges. There is a general trend for declining precipitation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17367725','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17367725"><span>Food and <span class="hlt">park</span> environments: neighborhood-level risks for childhood obesity in <span class="hlt">east</span> Los Angeles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kipke, Michele D; Iverson, Ellen; Moore, Deborah; Booker, Cara; Ruelas, Valerie; Peters, Anne L; Kaufman, Francine</p> <p>2007-04-01</p> <p>The rapid increase in obesity over the past two decades suggests that behavioral and environmental influences, including poor nutrition and physical inactivity, are fueling what is now widely recognized as a public health crisis. Yet, limited research has been conducted to examine how environmental factors, such as neighborhood-level characteristics, may be associated with increased risk for obesity. Community-level risk associated with childhood obesity was examined in <span class="hlt">East</span> Los Angeles, a community with one of the highest rates of childhood obesity in Los Angeles by triangulating: 1) spatial data for the number and location of food establishments relative to the location of schools; 2) observations regarding the availability and quality of fruits and vegetables in local grocery stores; and 3) observations regarding the quality and utilization of local <span class="hlt">parks</span>. The findings revealed that there were 190 food outlets in the study community, of which 93 (49%) were fast-food restaurants. Of the fast-food restaurants, 63% were within walking distance of a school. In contrast, there were 62 grocery stores, of which only 18% sold fresh fruits and/or vegetables of good quality. Of the stores that did sell fruits and/or vegetables, only four were within walking distance of a school. Although well maintained, the five <span class="hlt">parks</span> in this community accounted for only 37.28 acres, or 0.543 acres per 1000 residents. These findings suggest that children have easy access to fast food, and limited access to both healthy food options and <span class="hlt">parks</span> in which to engage in physical fitness activities. This was particularly true in areas around schools. The implications for these findings with regards to policy-related prevention and future research are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25123485','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25123485"><span><span class="hlt">Glaciers</span>. Attribution of global <span class="hlt">glacier</span> mass loss to anthropogenic and natural causes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Marzeion, Ben; Cogley, J Graham; Richter, Kristin; Parkes, David</p> <p>2014-08-22</p> <p>The ongoing global <span class="hlt">glacier</span> retreat is affecting human societies by causing sea-level rise, changing seasonal water availability, and increasing geohazards. Melting <span class="hlt">glaciers</span> are an icon of anthropogenic climate change. However, <span class="hlt">glacier</span> response times are typically decades or longer, which implies that the present-day <span class="hlt">glacier</span> retreat is a mixed response to past and current natural climate variability and current anthropogenic forcing. Here we show that only 25 ± 35% of the global <span class="hlt">glacier</span> mass loss during the period from 1851 to 2010 is attributable to anthropogenic causes. Nevertheless, the anthropogenic signal is detectable with high confidence in <span class="hlt">glacier</span> mass balance observations during 1991 to 2010, and the anthropogenic fraction of global <span class="hlt">glacier</span> mass loss during that period has increased to 69 ± 24%. Copyright © 2014, American Association for the Advancement of Science.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1914921H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1914921H"><span><span class="hlt">Glacier</span>Rocks - <span class="hlt">Glacier</span>-Headwall Interaction and its Influence on Rockfall Activity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hartmeyer, Ingo; Keuschnig, Markus; Krautblatter, Michael; Helfricht, Kay; Leith, Kerry; Otto, Jan-Christoph</p> <p>2017-04-01</p> <p>Climate models predict continued climate warming and a decrease of Austrian <span class="hlt">glaciers</span> to less than 20% of their present area by the end of this century. Rockfall from freshly exposed headwalls has been documented as an increasing risk factor with considerable significance for man and high-alpine infrastructure. Recent findings of a five-year terrestrial laserscanning campaign (2011-2016) monitoring glacial headwalls at the Kitzsteinhorn (3.203 m a.s.l.), Hohe Tauern Range, Austria, show the dramatic impact of <span class="hlt">glacier</span> thinning on adjacent headwalls: 80 % of the detected rockfall volumes were triggered from areas located less than 20 m above the current <span class="hlt">glacier</span> surface. Despite these implications, little is known about the thermal, mechanical and hydrological processes that operate at the <span class="hlt">glacier</span>-headwall interface (randkluft). Systemic in-situ monitoring of stability-relevant parameters are lacking, leaving fundamental gaps in the understanding of rockfall preconditioning in glacial headwalls and the geomorphological evolution of glaciated catchments. In this contribution we introduce the recently approved research project '<span class="hlt">Glacier</span>Rocks', which starts in 2017 and will run for at least three years. '<span class="hlt">Glacier</span>Rocks' will establish the worldwide first research site for long-term monitoring of stability-relevant processes inside a randkluft system. Based on the acquired monitoring data '<span class="hlt">Glacier</span>Rocks' is pursuing three overall aims at (1) gaining a better understanding of rockfall preconditioning in randklufts and related geomorphological shaping of headwalls, (2) analyzing poorly understood glacial thinning dynamics near headwalls, and (3) estimating present and future rockfall hazard potential in headwalls on a regional scale. The three system components (headwall, <span class="hlt">glacier</span>, randkluft) will be investigated by combining geomorphological, glaciological and meteorological methods. '<span class="hlt">Glacier</span>Rocks' will continuously monitor rock temperature, rock moisture, frost cracking</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.C32A..07S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.C32A..07S"><span><span class="hlt">Glacier</span> surge triggered by massive rock avalanche: Teleseismic and satellite image study of long-runout landslide onto RGO <span class="hlt">Glacier</span>, Pamirs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stark, C. P.; Wolovick, M.; Ekstrom, G.</p> <p>2012-12-01</p> <p><span class="hlt">Glacier</span> surges are thought to result from changes in resistance to sliding at the base of the ice mass. The reasons for such changes in basal conditions are not entirely understood, and this is in part because empirical constraints are severely limited. Recent work in the Karakoram and Pamir mountains, home to the majority of Earth's surging mountain <span class="hlt">glaciers</span>, has boosted observational data, but has led to diametrically opposed interpretations of their <span class="hlt">glacier</span> surging mechanics, ranging from thermal to hydrological switching. In this context we describe a surge of the RGO (Russian Geographical Society) <span class="hlt">Glacier</span> in the Pamirs triggered by a massive rock avalanche off Mt Garmo in 2001. Initial reports pegged the RGO <span class="hlt">Glacier</span> landslide as having been triggered in 2002 by strong ground motion originating from a nearby tectonic earthquake. We used multitemporal satellite imagery to establish failure must have struck in August-September 2001. This revised date was confirmed by reexamining teleseismic data recorded at stations in central Asia: it became clear that a landslide seismic source of magnitude Msw≈5.4 on 2001/09/02 had been misinterpreted as two tectonic sources located within kilometers of Mt Garmo. Exploiting a new technique we have developed for inverting long-period seismic waveforms, we show that a mass of rock and ice around 2.8×{}1011 kg collapsed to the SSE from an elevation of around 5800m, accelerated to a peak speed of about 60m/s, collided with the valley wall ˜ 2 km to the south and turned <span class="hlt">east</span> to run out a further 6km over significant fractions of the accumulation and ablation zones of the RGO <span class="hlt">Glacier</span>. Based on this estimate of landslide mass, we deduce that the supraglacial debris blanket generated by this rock avalanches averaged about 20m in thickness. By this reckoning, the Mt Garmo landslide is one of the largest in the last 33 years. Next we mapped the velocity field of the RGO <span class="hlt">Glacier</span> over time using multitemporal satellite imagery. We</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70168968','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70168968"><span>The differing biogeochemical and microbial signatures of <span class="hlt">glaciers</span> and rock <span class="hlt">glaciers</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>Fegel, Timothy S.; Baron, Jill S.; Fountain, Andrew G.; Johnson, Gunnar F.; Hall, Edward K.</p> <p>2016-01-01</p> <p><span class="hlt">Glaciers</span> and rock <span class="hlt">glaciers</span> supply water and bioavailable nutrients to headwater mountain lakes and streams across all regions of the American West. Here we present a comparative study of the metal, nutrient, and microbial characteristics of glacial and rock glacial influence on headwater ecosystems in three mountain ranges of the contiguous U.S.: The Cascade Mountains, Rocky Mountains, and Sierra Nevada. Several meltwater characteristics (water temperature, conductivity, pH, heavy metals, nutrients, complexity of dissolved organic matter (DOM), and bacterial richness and diversity) differed significantly between <span class="hlt">glacier</span> and rock <span class="hlt">glacier</span> meltwaters, while other characteristics (Ca2+, Fe3+, SiO2 concentrations, reactive nitrogen, and microbial processing of DOM) showed distinct trends between mountain ranges regardless of meltwater source. Some characteristics were affected both by <span class="hlt">glacier</span> type and mountain range (e.g. temperature, ammonium (NH4+) and nitrate (NO3- ) concentrations, bacterial diversity). Due to the ubiquity of rock <span class="hlt">glaciers</span> and the accelerating loss of the low latitude <span class="hlt">glaciers</span> our results point to the important and changing influence that these frozen features place on headwater ecosystems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title36-vol1/pdf/CFR-2010-title36-vol1-sec7-33.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title36-vol1/pdf/CFR-2010-title36-vol1-sec7-33.pdf"><span>36 CFR 7.33 - Voyageurs National <span class="hlt">Park</span>.</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>... following lakes and trails within Voyageurs National <span class="hlt">Park</span> are open to snowmobile use: (i) The frozen waters... Railroad Grade from the <span class="hlt">park</span> boundary north to Ash River, and then <span class="hlt">east</span> to Moose Bay, Namakan Lake. (iii... intersection with the Black Bay to Moose Bay portage, across Locator, War Club, Quill, Loiten, and Shoepack...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004AGUFM.V31E..07W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004AGUFM.V31E..07W"><span>Effects of lava-dome emplacement on the Mount St. Helens crater <span class="hlt">glacier</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Walder, J. S.; Schilling, S. P.; Denlinger, R. P.; Vallance, J. W.</p> <p>2004-12-01</p> <p>Since the end of the 1981-1986 episode of lava-dome growth at Mount St. Helens, an unusual <span class="hlt">glacier</span> has grown rapidly within the crater of the volcano. The <span class="hlt">glacier</span>, which is fed primarily by avalanching from the crater walls, contains about 30% rock debris by volume, has a maximum thickness of about 220 m and a volume of about 120 million cubic m, and forms a crescent that wraps around the old lava dome on both <span class="hlt">east</span> and west sides. The new (October 2004) lava dome in the south of the crater began to grow centered roughly on the contact between the old lava dome and the <span class="hlt">glacier</span>, in the process uplifting both ice and old dome rock. As the new dome is spreading to the south, the adjacent <span class="hlt">glacier</span> is bulging upward. Firn layers on the outer flank of the <span class="hlt">glacier</span> bulge have been warped upward almost vertically. In contrast, ice adjacent to the new dome has been thoroughly fractured. The overall style of deformation is reminiscent of that associated with salt-dome intrusion. Drawing an analogy to sand-box experiments, we suggest that the <span class="hlt">glacier</span> is being deformed by high-angle reverse faults propagating upward from depth. Comparison of Lidar images of the <span class="hlt">glacier</span> from September 2003 and October 2004 reveals not only the volcanogenic bulge but also elevated domains associated with the passage of kinematic waves, which are caused by <span class="hlt">glacier</span>-mass-balance perturbations and have nothing to do with volcanic activity. As of 25 October 2004, growth of the new lava dome has had negligible hydrological consequences. Ice-surface cauldrons are common consequences of intense melting caused by either subglacial eruptions (as in Iceland) or subglacial venting of hot gases (as presently taking place at Mount Spurr, Alaska). However, there has been a notable absence of ice-surface cauldrons in the Mount St. Helens crater <span class="hlt">glacier</span>, aside from a short-lived pond formed where the 1 October eruption pierced the <span class="hlt">glacier</span>. We suggest that heat transfer to the <span class="hlt">glacier</span> base is inefficient because</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24797737','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24797737"><span>Seasonal variations in the sources of natural and anthropogenic lead deposited at the <span class="hlt">East</span> Rongbuk <span class="hlt">Glacier</span> in the high-altitude Himalayas.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Burn-Nunes, Laurie; Vallelonga, Paul; Lee, Khanghyun; Hong, Sungmin; Burton, Graeme; Hou, Shugui; Moy, Andrew; Edwards, Ross; Loss, Robert; Rosman, Kevin</p> <p>2014-07-15</p> <p>Lead (Pb) isotopic compositions and concentrations, and barium (Ba) and indium (In) concentrations have been analysed at sub-annual resolution in three sections from a <110 m ice core dated to the 18th and 20th centuries, as well as snow pit samples dated to 2004/2005, recovered from the <span class="hlt">East</span> Rongbuk <span class="hlt">Glacier</span> in the high-altitude Himalayas. Ice core sections indicate that atmospheric chemistry prior to ~1,953 was controlled by mineral dust inputs, with no discernible volcanic or anthropogenic contributions. Eighteenth century monsoon ice core chemistry is indicative of dominant contributions from local Himalayan sources; non-monsoon ice core chemistry is linked to contributions from local (Himalayan), regional (Indian/Thar Desert) and long-range (North Africa, Central Asia) sources. Twentieth century monsoon and non-monsoon ice core data demonstrate similar seasonal sources of mineral dust, however with a transition to less-radiogenic isotopic signatures that suggests local and regional climate/environmental change. The snow pit record demonstrates natural and anthropogenic contributions during both seasons, with increased anthropogenic influence during non-monsoon times. Monsoon anthropogenic inputs are most likely sourced to South/South-<span class="hlt">East</span> Asia and/or India, whereas non-monsoon anthropogenic inputs are most likely sourced to India and Central Asia. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.C23A0594C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.C23A0594C"><span>Reconstructing the history of major Greenland <span class="hlt">glaciers</span> since the Little Ice Age</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Csatho, B. M.; Schenk, A. F.; van der Veen, C. J.; Stearns, L.; Babonis, G. S.</p> <p>2008-12-01</p> <p>The Greenland Ice Sheet may have been responsible for rapid sea level rise during the last interglacial period and recent studies indicate that it is likely to make a faster contribution to sea-level rise than previously believed. Rapid thinning and velocity increase has been observed on most major outlet <span class="hlt">glaciers</span> with terminus retreat that might lead to increased discharge from the interior and consequent further thinning and retreat. Potentially, such behavior could have serious implications for global sea level. However, the current thinning may simply be a manifestation of longer-term behavior of the ice sheet as it responds to the general warming following the Little Ice Age (LIA). Although Greenland outlet <span class="hlt">glaciers</span> have been comprehensively monitored since the 1980s, studies of long-term changes mostly rely on records of the calving front position. Such records can be misleading because the <span class="hlt">glacier</span> terminus, particularly if it is afloat, can either advance or retreat as ice further upstream thins and accelerates. To assess whether recent trends deviate from longer-term behavior, we examined three rapidly thinning and retreating outlet <span class="hlt">glaciers</span>, Jakobshavn Isbrae in west, Kangerdlussuaq <span class="hlt">Glacier</span> in <span class="hlt">east</span> and Petermann <span class="hlt">Glacier</span> in northwest Greenland. <span class="hlt">Glacier</span> surface and trimline elevations, as well as terminus positions were measured using historical photographs and declassified satellite imagery acquired between the 1940s and 1985. These results were combined with data from historical records, ground surveys, airborne laser altimetry, satellite observations and field mapping of lateral moraines and trimlines, to reconstruct the history of changes since the (LIA) up to the present. We identified several episodes of rapid thinning and ice shelf break-up, including thinning episodes that occurred when the calving front was stationary. Coastal weather station data are used to assess the influence of air temperatures and intensity of surface melting, and to isolate</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JSAES..77..218V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JSAES..77..218V"><span><span class="hlt">Glacier</span> monitoring and <span class="hlt">glacier</span>-climate interactions in the tropical Andes: A review</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Veettil, Bijeesh Kozhikkodan; Wang, Shanshan; Florêncio de Souza, Sergio; Bremer, Ulisses Franz; Simões, Jefferson Cardia</p> <p>2017-08-01</p> <p>In this review, we summarized the evolution of <span class="hlt">glacier</span> monitoring in the tropical Andes during the last few decades, particularly after the development of remote sensing and photogrammetry. Advantages and limitations of <span class="hlt">glacier</span> mapping, applied so far, in Venezuela, Colombia, Ecuador, Peru and Bolivia are discussed in detail. <span class="hlt">Glacier</span> parameters such as the equilibrium line altitude, snowline and mass balance were given special attention in understanding the complex cryosphere-climate interactions, particularly using remote sensing techniques. <span class="hlt">Glaciers</span> in the inner and the outer tropics were considered separately based on the precipitation and temperature conditions within a new framework. The applicability of various methods to use <span class="hlt">glacier</span> records to understand and reconstruct the tropical Andean climate between the Last Glacial Maximum (11,700 years ago) and the present is also explored in this paper. Results from various studies published recently were analyzed and we tried to understand the differences in the magnitudes of <span class="hlt">glacier</span> responses towards the climatic perturbations in the inner tropics and the outer tropics. Inner tropical <span class="hlt">glaciers</span>, particularly those in Venezuela and Colombia near the January Intertropical Convergence Zone (ITCZ), are more vulnerable to increase in temperature. Surface energy balance experiments show that outer tropical <span class="hlt">glaciers</span> respond to precipitation variability very rapidly in comparison with the temperature variability, particularly when moving towards the subtropics. We also analyzed the gradients in <span class="hlt">glacier</span> response to climate change from the Pacific coast towards the Amazon Basin as well as with the elevation. Based on the current trends synthesised from recent studies, it is hypothesized that the <span class="hlt">glaciers</span> in the inner tropics and the southern wet outer tropics will disappear first as a response to global warming whereas <span class="hlt">glaciers</span> in the northern wet outer tropics and dry outer tropics show resistance to warming trends due to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca2540.photos.326599p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca2540.photos.326599p/"><span>View of Chapel <span class="hlt">Park</span>, showing bomb shelters at right 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>View of Chapel <span class="hlt">Park</span>, showing bomb shelters at right foreground, from building 746 <span class="hlt">parking</span> lot across Walnut Avenue; camera facing north. - Mare Island Naval Shipyard, <span class="hlt">East</span> of Nave Drive, Vallejo, Solano County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70135865','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70135865"><span>Velocity measurements and changes in position of Thwaites <span class="hlt">Glacier</span>/iceberg tongue from aerial photography, Landsat images and NOAA AVHRR 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>Ferrigno, Jane G.; Lucchitta, Baerbel K.; Mullinsallison, A. L.; Allen, Robert J.; Gould, W. G.</p> <p>1993-01-01</p> <p>The Thwaites <span class="hlt">Glacier</span>/iceberg tongue complex has been a significant feature of the Antarctic coastline for at least 50 years. In 1986, major changes began to occur in this area. Fast ice melted and several icebergs calved from the base of the iceberg tongue and the terminus of Thwaites <span class="hlt">Glacier</span>. The iceberg tongue rotated to an <span class="hlt">east</span>-west orientation and drifted westward. Between 1986 and 1992, a total of 140 km of drift has occurred. Remote digital velocity measurements were made on Thwaites <span class="hlt">Glacier</span> using sequential Landsat images to try to determine if changes in velocity had occurred in conjunction with the changes in ice position. Examination of the morphology of the <span class="hlt">glacier</span>/iceberg tongue showed no evidence of surge activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.9407L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.9407L"><span>Remotely-sensed and field-based observations of <span class="hlt">glacier</span> change in the Annapurna-Manaslu region, Nepal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lovell, Arminel; Carr, Rachel; Stokes, Chris</p> <p>2017-04-01</p> <p>Himalayan <span class="hlt">glaciers</span> have shrunk rapidly during the past twenty years. Understanding the factors controlling these losses is vital for forecasting changes in water resources, as the Himalaya houses the headwaters of major river systems, with densely populated catchments downstream. However, our knowledge of Himalayan <span class="hlt">glaciers</span> is comparatively limited, due to their high-altitude, remote location. This is particularly the case in the Annapurna-Manaslu region, which has received relatively little scientific attention to date. Here, we present initial findings from remotely sensed data analysis, and our first field campaign in October 2016. Feature tracking of Band 8 Landsat imagery demonstrates that velocities in the region reach a maximum of 70-100 m a-1 , which is somewhat faster than those reported in the Khumbu region (e.g. Quincey et al 2009). A number of <span class="hlt">glaciers</span> have substantial stagnant ice tongues, and most are flowing faster in the upper ablation zone than in the lower sections. The most rapidly flowing <span class="hlt">glaciers</span> are located in the south-<span class="hlt">east</span> of the Annapurna-Manaslu region and tend to also be the largest. Interestingly, initial observations suggest that the debris-covered ablation zones in the south-<span class="hlt">east</span> are flowing more rapidly than the smaller, clean-ice <span class="hlt">glaciers</span> in the north of the region. Comparison of velocities between 2000-2001 and 2014-2015 suggests deceleration on some <span class="hlt">glacier</span> tongues. In October 2016, we conducted fieldwork on Annapurna South <span class="hlt">Glacier</span>, located at the foot of Annapurna I. Here, we collected a number of datasets, with the aim of assessing the relationship between surface elevation change, ice velocities and debris cover. These included: i) installing ablation stakes in areas with varying debris cover; ii) quantifying debris characteristics, using Wolman counting and by measuring thickness; iii) surveying the <span class="hlt">glacier</span> surface, using a differential GPS; iv) monitoring ice cliff melting, using Structure from Motion and; v) measuring surface</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('http://adsabs.harvard.edu/abs/2016AGUFMEP41D..04A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMEP41D..04A"><span>Modeling the Rock <span class="hlt">Glacier</span> Cycle</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anderson, R. S.; Anderson, L. S.</p> <p>2016-12-01</p> <p>Rock <span class="hlt">glaciers</span> are common in many mountain ranges in which the ELA lies above the peaks. They represent some of the most identifiable components of today's cryosphere in these settings. Their oversteepened snouts pose often-overlooked hazards to travel in alpine terrain. Rock <span class="hlt">glaciers</span> are supported by avalanches and by rockfall from steep headwalls. The winter's avalanche cone must be sufficiently thick not to melt entirely in the summer. The spatial distribution of rock <span class="hlt">glaciers</span> reflects this dependence on avalanche sources; they are most common on lee sides of ridges where wind-blown snow augments the avalanche source. In the absence of rockfall, this would support a short, cirque <span class="hlt">glacier</span>. Depending on the relationship between rockfall and avalanche patterns, "talus-derived" and "<span class="hlt">glacier</span>-derived" rock <span class="hlt">glaciers</span> are possible. Talus-derived: If the spatial distribution of rock delivery is similar to the avalanche pattern, the rock-ice mixture will travel an englacial path that is downward through the short accumulation zone before turning upward in the ablation zone. Advected debris is then delivered to the base of a growing surface debris layer that reduces the ice melt rate. The physics is identical to the debris-covered <span class="hlt">glacier</span> case. <span class="hlt">Glacier</span>-derived: If on the other hand rockfall from the headwall rolls beyond the avalanche cone, it is added directly to the ablation zone of the <span class="hlt">glacier</span>. The avalanche accumulation zone then supports a pure ice core to the rock <span class="hlt">glacier</span>. We have developed numerical models designed to capture the full range of <span class="hlt">glacier</span> to debris-covered <span class="hlt">glacier</span> to rock <span class="hlt">glacier</span> behavior. The hundreds of meter lengths, tens of meters thicknesses, and meter per year speeds of rock <span class="hlt">glaciers</span> are well described by the models. The model can capture both "talus-derived" and "<span class="hlt">glacier</span>-derived" rock <span class="hlt">glaciers</span>. We explore the dependence of <span class="hlt">glacier</span> behavior on climate histories. As climate warms, a pure ice debris-covered <span class="hlt">glacier</span> can transform to a much shorter rock</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e002000.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e002000.html"><span>Susitna <span class="hlt">Glacier</span>, Alaska</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>NASA image acquired August 27, 2009 Like rivers of liquid water, <span class="hlt">glaciers</span> flow downhill, with tributaries joining to form larger rivers. But where water rushes, ice crawls. As a result, <span class="hlt">glaciers</span> gather dust and dirt, and bear long-lasting evidence of past movements. Alaska’s Susitna <span class="hlt">Glacier</span> revealed some of its long, grinding journey when the Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) on NASA’s Terra satellite passed overhead on August 27, 2009. This satellite image combines infrared, red, and green wavelengths to form a false-color image. Vegetation is red and the glacier’s surface is marbled with dirt-free blue ice and dirt-coated brown ice. Infusions of relatively clean ice push in from tributaries in the north. The <span class="hlt">glacier</span> surface appears especially complex near the center of the image, where a tributary has pushed the ice in the main <span class="hlt">glacier</span> slightly southward. A photograph taken by researchers from the U.S. Geological Survey (archived by the National Snow and Ice Data Center) shows an equally complicated Susitna <span class="hlt">Glacier</span> in 1970, with dirt-free and dirt-encrusted surfaces forming stripes, curves, and U-turns. Susitna flows over a seismically active area. In fact, a 7.9-magnitude quake struck the region in November 2002, along a previously unknown fault. Geologists surmised that earthquakes had created the steep cliffs and slopes in the <span class="hlt">glacier</span> surface, but in fact most of the jumble is the result of surges in tributary <span class="hlt">glaciers</span>. <span class="hlt">Glacier</span> surges—typically short-lived events where a <span class="hlt">glacier</span> moves many times its normal rate—can occur when melt water accumulates at the base and lubricates the flow. This water may be supplied by meltwater lakes that accumulate on top of the <span class="hlt">glacier</span>; some are visible in the lower left corner of this image. The underlying bedrock can also contribute to <span class="hlt">glacier</span> surges, with soft, easily deformed rock leading to more frequent surges. NASA Earth Observatory image created by Jesse Allen and Robert</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003AGUFM.C11C0844M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003AGUFM.C11C0844M"><span>Recent Observations and Structural Analysis of Surge-Type <span class="hlt">Glaciers</span> in the <span class="hlt">Glacier</span> Bay Area</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mayer, H.; Herzfeld, U. C.</p> <p>2003-12-01</p> <p>The Chugach-St.-Elias Mountains in North America hold the largest non-polar connected glaciated area of the world. Most of its larger <span class="hlt">glaciers</span> are surge-type <span class="hlt">glaciers</span>. In the summer of 2003, we collected aerial photographic and GPS data over numerous <span class="hlt">glaciers</span> in the eastern St. Elias Mountains, including the <span class="hlt">Glacier</span> Bay area. Observed <span class="hlt">glaciers</span> include Davidson, Casement, McBride, Riggs, Cushing, Carroll, Rendu, Tsirku, Grand Pacific, Melbern, Ferris, Margerie, Johns Hopkins, Lamplugh, Reid, Burroughs, Morse, Muir and Willard <span class="hlt">Glaciers</span>, of which Carroll, Rendu, Ferris, Grand Pacific, Johns Hopkins and Margerie <span class="hlt">Glaciers</span> are surge-type <span class="hlt">glaciers</span>. Our approach utilizes a quantitative analysis of surface patterns, following the principles of structural geology for the analysis of brittle-deformation patterns (manifested in crevasses) and ductile deformation patterns (visible in folded moraines). First results will be presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22768074','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22768074"><span>Protected areas: mixed success in conserving <span class="hlt">East</span> Africa's evergreen forests.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pfeifer, Marion; Burgess, Neil D; Swetnam, Ruth D; Platts, Philip J; Willcock, Simon; Marchant, Robert</p> <p>2012-01-01</p> <p>In <span class="hlt">East</span> Africa, human population growth and demands for natural resources cause forest loss contributing to increased carbon emissions and reduced biodiversity. Protected Areas (PAs) are intended to conserve habitats and species. Variability in PA effectiveness and 'leakage' (here defined as displacement of deforestation) may lead to different trends in forest loss within, and adjacent to, existing PAs. Here, we quantify spatial variation in trends of evergreen forest coverage in <span class="hlt">East</span> Africa between 2001 and 2009, and test for correlations with forest accessibility and environmental drivers. We investigate PA effectiveness at local, landscape and national scales, comparing rates of deforestation within <span class="hlt">park</span> boundaries with those detected in <span class="hlt">park</span> buffer zones and in unprotected land more generally. Background forest loss (BFL) was estimated at -9.3% (17,167 km(2)), but varied between countries (range: -0.9% to -85.7%; note: no BFL in South Sudan). We document high variability in PA effectiveness within and between PA categories. The most successful PAs were National <span class="hlt">Parks</span>, although only 26 out of 48 <span class="hlt">parks</span> increased or maintained their forest area (i.e. Effective <span class="hlt">parks</span>). Forest Reserves (Ineffective <span class="hlt">parks</span>, i.e. <span class="hlt">parks</span> that lose forest from within boundaries: 204 out of 337), Nature Reserves (six out of 12) and Game <span class="hlt">Parks</span> (24 out of 26) were more likely to lose forest cover. Forest loss in buffer zones around PAs exceeded background forest loss, in some areas indicating leakage driven by Effective National <span class="hlt">Parks</span>. Human pressure, forest accessibility, protection status, distance to fires and long-term annual rainfall were highly significant drivers of forest loss in <span class="hlt">East</span> Africa. Some of these factors can be addressed by adjusting <span class="hlt">park</span> management. However, addressing close links between livelihoods, natural capital and poverty remains a fundamental challenge in <span class="hlt">East</span> Africa's forest conservation efforts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025232','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025232"><span>The slow advance of a calving <span class="hlt">glacier</span>: Hubbard <span class="hlt">Glacier</span>, Alaska, U.S.A</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Trabant, D.C.; Krimmel, R.M.; Echelmeyer, K.A.; Zirnheld, S.L.; Elsberg, D.H.</p> <p>2003-01-01</p> <p>Hubbard <span class="hlt">Glacier</span> is the largest tidewater <span class="hlt">glacier</span> in North America. In contrast to most <span class="hlt">glaciers</span> in Alaska and northwestern Canada, Hubbard <span class="hlt">Glacier</span> thickened and advanced during the 20th century. This atypical behavior is an important example of how insensitive to climate a <span class="hlt">glacier</span> can become during parts of the calving <span class="hlt">glacier</span> cycle. As this <span class="hlt">glacier</span> continues to advance, it will close the seaward entrance to 50 km long Russell Fjord and create a <span class="hlt">glacier</span>-dammed, brackish-water lake. This paper describes measured changes in ice thickness, ice speed, terminus advance and fjord bathymetry of Hubbard <span class="hlt">Glacier</span>, as determined from airborne laser altimetry, aerial photogrammetry, satellite imagery and bathymetric measurements. The data show that the lower regions of the <span class="hlt">glacier</span> have thickened by as much as 83 m in the last 41 years, while the entire <span class="hlt">glacier</span> increased in volume by 14.1 km3. Ice speeds are generally decreasing near the calving face from a high of 16.5 m d-1 in 1948 to 11.5 m d-1 in 2001. The calving terminus advanced at an average rate of about 16 m a-1 between 1895 and 1948 and accelerated to 32 m a-1 since 1948. However, since 1986, the advance of the part of the terminus in Disenchantment Bay has slowed to 28 m a-1. Bathymetric data from the lee slope of the submarine terminal moraine show that between 1978 and 1999 the moraine advanced at an average rate of 32 m a-1, which is the same as that of the calving face.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27828967','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27828967"><span>Quantification and Analysis of Icebergs in a Tidewater <span class="hlt">Glacier</span> Fjord Using an Object-Based Approach.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>McNabb, Robert W; Womble, Jamie N; Prakash, Anupma; Gens, Rudiger; Haselwimmer, Christian E</p> <p>2016-01-01</p> <p>Tidewater <span class="hlt">glaciers</span> are <span class="hlt">glaciers</span> that terminate in, and calve icebergs into, the ocean. In addition to the influence that tidewater <span class="hlt">glaciers</span> have on physical and chemical oceanography, floating icebergs serve as habitat for marine animals such as harbor seals (Phoca vitulina richardii). The availability and spatial distribution of <span class="hlt">glacier</span> ice in the fjords is likely a key environmental variable that influences the abundance and distribution of selected marine mammals; however, the amount of ice and the fine-scale characteristics of ice in fjords have not been systematically quantified. Given the predicted changes in <span class="hlt">glacier</span> habitat, there is a need for the development of methods that could be broadly applied to quantify changes in available ice habitat in tidewater <span class="hlt">glacier</span> fjords. We present a case study to describe a novel method that uses object-based image analysis (OBIA) to classify floating <span class="hlt">glacier</span> ice in a tidewater <span class="hlt">glacier</span> fjord from high-resolution aerial digital imagery. Our objectives were to (i) develop workflows and rule sets to classify high spatial resolution airborne imagery of floating <span class="hlt">glacier</span> ice; (ii) quantify the amount and fine-scale characteristics of floating <span class="hlt">glacier</span> ice; (iii) and develop processes for automating the object-based analysis of floating <span class="hlt">glacier</span> ice for large number of images from a representative survey day during June 2007 in Johns Hopkins Inlet (JHI), a tidewater <span class="hlt">glacier</span> fjord in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, southeastern Alaska. On 18 June 2007, JHI was comprised of brash ice ([Formula: see text] = 45.2%, SD = 41.5%), water ([Formula: see text] = 52.7%, SD = 42.3%), and icebergs ([Formula: see text] = 2.1%, SD = 1.4%). Average iceberg size per scene was 5.7 m2 (SD = 2.6 m2). We estimate the total area (± uncertainty) of iceberg habitat in the fjord to be 455,400 ± 123,000 m2. The method works well for classifying icebergs across scenes (classification accuracy of 75.6%); the largest classification errors occur in areas with</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5102356','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5102356"><span>Quantification and Analysis of Icebergs in a Tidewater <span class="hlt">Glacier</span> Fjord Using an Object-Based Approach</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>McNabb, Robert W.; Womble, Jamie N.; Prakash, Anupma; Gens, Rudiger; Haselwimmer, Christian E.</p> <p>2016-01-01</p> <p>Tidewater <span class="hlt">glaciers</span> are <span class="hlt">glaciers</span> that terminate in, and calve icebergs into, the ocean. In addition to the influence that tidewater <span class="hlt">glaciers</span> have on physical and chemical oceanography, floating icebergs serve as habitat for marine animals such as harbor seals (Phoca vitulina richardii). The availability and spatial distribution of <span class="hlt">glacier</span> ice in the fjords is likely a key environmental variable that influences the abundance and distribution of selected marine mammals; however, the amount of ice and the fine-scale characteristics of ice in fjords have not been systematically quantified. Given the predicted changes in <span class="hlt">glacier</span> habitat, there is a need for the development of methods that could be broadly applied to quantify changes in available ice habitat in tidewater <span class="hlt">glacier</span> fjords. We present a case study to describe a novel method that uses object-based image analysis (OBIA) to classify floating <span class="hlt">glacier</span> ice in a tidewater <span class="hlt">glacier</span> fjord from high-resolution aerial digital imagery. Our objectives were to (i) develop workflows and rule sets to classify high spatial resolution airborne imagery of floating <span class="hlt">glacier</span> ice; (ii) quantify the amount and fine-scale characteristics of floating <span class="hlt">glacier</span> ice; (iii) and develop processes for automating the object-based analysis of floating <span class="hlt">glacier</span> ice for large number of images from a representative survey day during June 2007 in Johns Hopkins Inlet (JHI), a tidewater <span class="hlt">glacier</span> fjord in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, southeastern Alaska. On 18 June 2007, JHI was comprised of brash ice (x¯ = 45.2%, SD = 41.5%), water (x¯ = 52.7%, SD = 42.3%), and icebergs (x¯ = 2.1%, SD = 1.4%). Average iceberg size per scene was 5.7 m2 (SD = 2.6 m2). We estimate the total area (± uncertainty) of iceberg habitat in the fjord to be 455,400 ± 123,000 m2. The method works well for classifying icebergs across scenes (classification accuracy of 75.6%); the largest classification errors occur in areas with densely-packed ice, low contrast between</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/1998/0791/pdf/ofr98-791.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/1998/0791/pdf/ofr98-791.pdf"><span>Physical characteristics of dungeness crab and halibut habitats in <span class="hlt">Glacier</span> Bay</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Cochrane, Guy R.; Carlson, Paul R.; Denny, Jane F.; Boyle, Michael E.; Taggart, S. James; Hooge, Philip N.</p> <p>1998-01-01</p> <p>In <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span>, Alaska there are ongoing studies of Dungeness Crab (Cancer magister) and Pacific Halibut (Hippoglosus stenolepis). Scientists of the United States Geological Survey (USGS) are attempting to ascertain life history, distribution, and abundance, and to determine the effects of commercial fishing in the <span class="hlt">park</span> (Carlson et al., 1998). Statistical sampling studies suggest that seafloor characteristics and bathymetry affect the distribution, abundance and behavior of benthic species. Examples include the distribution of Dungeness crab which varies from 78 to 2012 crabs/ha in nearshore areas to depths of 18 m (O'Clair et al., 1995), and changes in halibut foraging behavior according to bottom type (Chilton et al., 1995). This report discusses geophysical data collected in six areas within the <span class="hlt">park</span> in 1998. The geophysical surveying done in this and previous studies will be combined with existing population and sonic-tracking data sets as well as future sediment sampling, scuba, submersible, and bottom video camera observations to better understand Dungeness crab and Pacific halibut habitat relationships.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://arc.lib.montana.edu/snow-science/objects/ISSW_P-098.pdf','USGSPUBS'); return false;" href="http://arc.lib.montana.edu/snow-science/objects/ISSW_P-098.pdf"><span>Using GIS and Google Earth for the creation of the Going-to-the-Sun Road Avalanche Atlas, <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Peitzsch, Erich H.; Fagre, Daniel B.; Dundas, Mark</p> <p>2010-01-01</p> <p>Snow avalanche paths are key geomorphologic features in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, and an important component of mountain ecosystems: they are isolated within a larger ecosystem, they are continuously disturbed, and they contain unique physical characteristics (Malanson and Butler, 1984). Avalanches impact subalpine forest structure and function, as well as overall biodiversity (Bebi et al., 2009). Because avalanches are dynamic phenomena, avalanche path geometry and spatial extent depend upon climatic regimes. The USGS/GNP Avalanche Program formally began in 2003 as an avalanche forecasting program for the spring opening of the ever-popular Going-to-the-Sun Road (GTSR), which crosses through 37 identified avalanche paths. Avalanche safety and forecasting is a necessary part of the GTSR spring opening procedures. An avalanche atlas detailing topographic parameters and oblique photographs was completed for the GTSR corridor in response to a request from GNP personnel for planning and resource management. Using ArcMap 9.2 GIS software, polygons were created for every avalanche path affecting the GTSR using aerial imagery, field-based observations, and GPS measurements of sub-meter accuracy. Spatial attributes for each path were derived within the GIS. Resulting products include an avalanche atlas book for operational use, a geoPDF of the atlas, and a Google Earth flyover illustrating each path and associated photographs. The avalanche atlas aids <span class="hlt">park</span> management in worker safety, infrastructure planning, and natural resource protection by identifying avalanche path patterns and location. The atlas was created for operational and planning purposes and is also used as a foundation for research such as avalanche ecology projects and avalanche path runout modeling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMGC21G..01T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMGC21G..01T"><span>Climatic Teleconnections Recorded By Tropical Mountain <span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thompson, L. G.; Permana, D.; Mosley-Thompson, E.; Davis, M. E.</p> <p>2014-12-01</p> <p>Information from ice cores from the world's highest mountains in the Tropics demonstrates both local climate variability and a high degree of teleconnectivity across the Pacific basin. Here we examine recently recovered ice core records from <span class="hlt">glaciers</span> near Puncak Jaya in Papua, Indonesia, which lie on the highest peak between the Himalayas and the South American Andes. These <span class="hlt">glaciers</span> are located on the western side of the Tropical Pacific warm pool, which is the "center of action" for interannual climate variability dominated by El Niño-Southern Oscillation (ENSO). ENSO either directly or indirectly affects most regions of Earth and their populations. In 2010, two ice cores measuring 32.13 m and 31.25 m were recovered to bedrock from the <span class="hlt">East</span> Northwall Firn ice field. Both have been analyzed in high resolution (~3 cm sample length, 1156 and 1606 samples, respectively) for stable isotopes, dust, major ions and tritium concentrations. To better understand the controls on the oxygen isotopic (δ18 O) signal for this region, daily rainfall samples were collected between January 2013 and February 2014 at five weather stations over a distance of ~90 km ranging from 9 meters above sea level (masl) on the southern coast up to 3945 masl. The calculated isotopic lapse rate for this region is 0.24 ‰/100m. Papua, Indonesian ice core records are compared to ice core records from Dasuopu <span class="hlt">Glacier</span> in the central Himalayas and from Quelccaya, Huascarán, Hualcán and Coropuna ice fields in the tropical Andes of Peru on the eastern side of the Pacific Ocean. The composite of the annual isotopic time series from these cores is significantly (R2 =0.53) related to tropical Pacific sea surface temperatures (SSTs), reflecting the strong linkage between tropical Pacific SSTs associated with ENSO and tropospheric temperatures in the low latitudes. New data on the already well-documented concomitant loss of ice on Quelccaya, Kilimanjaro in eastern Africa and the ice fields near Puncak</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.7117S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.7117S"><span>Warm water and life beneath the grounding zone of an Antarctic outlet <span class="hlt">glacier</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sugiyama, Shin; Sawagaki, Takanobu; Fukuda, Takehiro</p> <p>2013-04-01</p> <p>Ice-ocean interaction plays a key role in rapidly changing Antarctic ice sheet margins. Recent studies demonstrated that warming ocean is eroding floating part of the ice sheet, resulting in thinning, retreat and acceleration of ice shelves and outlet <span class="hlt">glaciers</span>. Field data are necessary to understand such processes, but direct observations at the interface of ice and the ocean are lacking, particularly beneath the grounding zone. To better understand the interaction of Antarctic ice sheet and the ocean, we performed subglacial measurements through boreholes drilled in the grounding zone of Langhovde <span class="hlt">Glacier</span>, an outlet <span class="hlt">glacier</span> in <span class="hlt">East</span> Antarctica. Langhovde <span class="hlt">Glacier</span> is located at 69°12'S, 39°48'E, approximately 20 km south of a Japanese research station Syowa. The <span class="hlt">glacier</span> discharges ice into Lützow-holm Bay through a 3-km-wide floating terminus at a rate of 130 m a-1. Fast flowing feature is confined by bedrock to the west and slow moving ice to the <span class="hlt">east</span>, and it extends about 10 km upglacier from the calving front. In 2011/12 austral summer season, we operated a hot water drilling system to drill through the <span class="hlt">glacier</span> at 2.5 and 3 km from the terminus. Inspections of the boreholes revealed the ice was underlain by a shallow saline water layer. Ice and water column thicknesses were found to be 398 and 24 m at the first site, and 431 and 10 m at the second site. Judging from ice surface and bed elevations, the drilling sites were situated at within a several hundred meters from the grounding line. Sensors were lowered into the boreholes to measure temperature, salinity and current within the subglacial water layer. Salinity and temperature from the two sites were fairly uniform (34.25±0.05 PSU and -1.45±0.05°C), indicating vertical and horizontal mixing in the layer. The measured temperature was >0.7°C warmer than the in-situ freezing point, and very similar to the values measured in the open ocean near the <span class="hlt">glacier</span> front. Subglacial current was up to 3 cm/s, which</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008JGRD..113.5103R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008JGRD..113.5103R"><span>Predicting the response of seven Asian <span class="hlt">glaciers</span> to future climate scenarios using a simple linear <span class="hlt">glacier</span> model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ren, Diandong; Karoly, David J.</p> <p>2008-03-01</p> <p>Observations from seven Central Asian <span class="hlt">glaciers</span> (35-55°N; 70-95°E) are used, together with regional temperature data, to infer uncertain parameters for a simple linear model of the <span class="hlt">glacier</span> length variations. The <span class="hlt">glacier</span> model is based on first order <span class="hlt">glacier</span> dynamics and requires the knowledge of reference states of forcing and <span class="hlt">glacier</span> perturbation magnitude. An adjoint-based variational method is used to optimally determine the <span class="hlt">glacier</span> reference states in 1900 and the uncertain <span class="hlt">glacier</span> model parameters. The simple <span class="hlt">glacier</span> model is then used to estimate the <span class="hlt">glacier</span> length variations until 2060 using regional temperature projections from an ensemble of climate model simulations for a future climate change scenario (SRES A2). For the period 2000-2060, all <span class="hlt">glaciers</span> are projected to experience substantial further shrinkage, especially those with gentle slopes (e.g., <span class="hlt">Glacier</span> Chogo Lungma retreats ˜4 km). Although nearly one-third of the year 2000 length will be reduced for some small <span class="hlt">glaciers</span>, the existence of the <span class="hlt">glaciers</span> studied here is not threatened by year 2060. The differences between the individual <span class="hlt">glacier</span> responses are large. No straightforward relationship is found between <span class="hlt">glacier</span> size and the projected fractional change of its length.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1015985','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1015985"><span>Addendum to the <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> Site-Wide Residual Contamination Remedial Investigation Work Plan Oak Ridge, Tennessee</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>SAIC</p> <p>2011-04-01</p> <p>The <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> Site-Wide Residual Contamination Remedial Investigation Work Plan (DOE 2004) describes the planned fieldwork to support the remedial investigation (RI) for residual contamination at the <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> (ETTP) not addressed in previous Comprehensive Environmental Response, Compensation, and Liability Act of 1980 (CERCLA) decisions. This Addendum describes activities that will be conducted to gather additional information in Zone 1 of the ETTP for groundwater, surface water, and sediments. This Addendum has been developed from agreements reached in meetings held on June 23, 2010, August 25, 2010, October 13, 2010, November 13, 2010, December 1, 2010,more » and January 13, 2011, with representatives of the U. S. Department of Energy (DOE), U. S. Environmental Protection Agency (EPA), and Tennessee Department of Environment and Conservation (TDEC). Based on historical to recent groundwater data for ETTP and the previously completed Sitewide Remedial Investigation for the ETTP (DOE 2007a), the following six areas of concern have been identified that exhibit groundwater contamination downgradient of these areas above state of Tennessee and EPA drinking water maximum contaminant levels (MCLs): (1) K-720 Fly Ash Pile, (2) K-770 Scrap Yard, (3) Duct Island, (4) K-1085 Firehouse Burn/J.A. Jones Maintenance Area, (5) Contractor's Spoil Area (CSA), and (6) Former K-1070-A Burial Ground. The paper presents a brief summary of the history of the areas, the general conceptual models for the observed groundwater contamination, and the data gaps identified.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45.2706D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45.2706D"><span>Basal Settings Control Fast Ice Flow in the Recovery/Slessor/Bailey Region, <span class="hlt">East</span> Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Diez, Anja; Matsuoka, Kenichi; Ferraccioli, Fausto; Jordan, Tom A.; Corr, Hugh F.; Kohler, Jack; Olesen, Arne V.; Forsberg, René</p> <p>2018-03-01</p> <p>The region of Recovery <span class="hlt">Glacier</span>, Slessor <span class="hlt">Glacier</span>, and Bailey Ice Stream, <span class="hlt">East</span> Antarctica, has remained poorly explored, despite representing the largest potential contributor to future global sea level rise on a centennial to millennial time scale. Here we use new airborne radar data to improve knowledge about the bed topography and investigate controls of fast ice flow. Recovery <span class="hlt">Glacier</span> is underlain by an 800 km long trough. Its fast flow is controlled by subglacial water in its upstream and topography in its downstream region. Fast flow of Slessor <span class="hlt">Glacier</span> is controlled by the presence of subglacial water on a rough crystalline bed. Past ice flow of adjacent Recovery and Slessor <span class="hlt">Glaciers</span> was likely connected via the newly discovered Recovery-Slessor Gate. Changes in direction and speed of past fast flow likely occurred for upstream parts of Recovery <span class="hlt">Glacier</span> and between Slessor <span class="hlt">Glacier</span> and Bailey Ice Stream. Similar changes could also reoccur here in the future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://arc.lib.montana.edu/snow-science/item/1664','USGSPUBS'); return false;" href="http://arc.lib.montana.edu/snow-science/item/1664"><span>Timing of wet snow avalanche activity: An analysis from <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA.</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Peitzsch, Erich H.; Hendrikx, Jordy; Fagre, Daniel B.</p> <p>2012-01-01</p> <p>Wet snow avalanches pose a problem for annual spring road opening operations along the Going-to-the-Sun Road (GTSR) in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA. A suite of meteorological metrics and snow observations has been used to forecast for wet slab and glide avalanche activity. However, the timing of spring wet slab and glide avalanches is a difficult process to forecast and requires new capabilities. For the 2011 and 2012 spring seasons we tested a previously developed classification tree model which had been trained on data from 2003-2010. For 2011, this model yielded a 91% predictive rate for avalanche days. For 2012, the model failed to capture any of the avalanche days observed. We then investigated these misclassified avalanche days in the 2012 season by comparing them to the misclassified days from the original dataset from which the model was trained. Results showed no significant difference in air temperature variables between this year and the original training data set for these misclassified days. This indicates that 2012 was characterized by avalanche days most similar to those that the model struggled with in the original training data. The original classification tree model showed air temperature to be a significant variable in wet avalanche activity which implies that subsequent movement of meltwater through the snowpack is also important. To further understand the timing of water flow we installed two lysimeters in fall 2011 before snow accumulation. Water flow showed a moderate correlation with air temperature later in the season and no synchronous pattern associated with wet slab and glide avalanche activity. We also characterized snowpack structure as the snowpack transitioned from a dry to a wet snowpack throughout the spring. This helped to assess potential failure layers of wet snow avalanches and the timing of avalanches compared to water moving through the snowpack. These tools (classification tree model and lysimeter data), combined with</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18...73R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18...73R"><span>Debris-covered Himalayan <span class="hlt">glaciers</span> under a changing climate: observations and modelling of Khumbu <span class="hlt">Glacier</span>, Nepal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rowan, Ann; Quincey, Duncan; Egholm, David; Gibson, Morgan; Irvine-Fynn, Tristram; Porter, Philip; Glasser, Neil</p> <p>2016-04-01</p> <p>Many mountain <span class="hlt">glaciers</span> are characterised in their lower reaches by thick layers of rock debris that insulate the <span class="hlt">glacier</span> surface from solar radiation and atmospheric warming. Supraglacial debris modifies the response of these <span class="hlt">glaciers</span> to climate change compared to <span class="hlt">glaciers</span> with clean-ice surfaces. However, existing modelling approaches to predicting variations in the extent and mass balance of debris-covered <span class="hlt">glaciers</span> have relied on numerical models that represent the processes governing <span class="hlt">glaciers</span> with clean-ice surfaces, and yield conflicting results. Moreover, few data exist describing the mass balance of debris-covered <span class="hlt">glaciers</span> and many observations are only made over short periods of time, but these data are needed to constrain and validate numerical modelling experiments. To investigate the impact of supraglacial debris on the response of a <span class="hlt">glacier</span> to climate change, we developed a numerical model that couples the flow of ice and debris to include important feedbacks between mass balance, ice flow and debris accumulation. We applied this model to a large debris-covered Himalayan <span class="hlt">glacier</span> - Khumbu <span class="hlt">Glacier</span> in the Everest region of Nepal. Our results demonstrate that supraglacial debris prolongs the response of the <span class="hlt">glacier</span> to warming air temperatures and causes lowering of the <span class="hlt">glacier</span> surface in situ, concealing the magnitude of mass loss when compared with estimates based on glacierised area. Since the Little Ice Age, the volume of Khumbu <span class="hlt">Glacier</span> has reduced by 34%, while <span class="hlt">glacier</span> area has reduced by only 6%. We predict a further decrease in <span class="hlt">glacier</span> volume of 8-10% by AD2100 accompanied by dynamic and physical detachment of the debris-covered tongue from the active <span class="hlt">glacier</span> within the next 150 years. For five months during the 2014 summer monsoon, we measured temperature profiles through supraglacial debris and proglacial discharge on Khumbu <span class="hlt">Glacier</span>. We found that temperatures at the ice surface beneath 0.4-0.7 m of debris were sufficient to promote considerable</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://arc.lib.montana.edu/snow-science/item/962','USGSPUBS'); return false;" href="http://arc.lib.montana.edu/snow-science/item/962"><span>High resolution tree-ring based spatial reconstructions of snow avalanche activity in <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Pederson, Gregory T.; Reardon, Blase; Caruso, C.J.; Fagre, Daniel B.</p> <p>2006-01-01</p> <p>Effective design of avalanche hazard mitigation measures requires long-term records of natural avalanche frequency and extent. Such records are also vital for determining whether natural avalanche frequency and extent vary over time due to climatic or biophysical changes. Where historic records are lacking, an accepted substitute is a chronology developed from tree-ring responses to avalanche-induced damage. This study evaluates a method for using tree-ring chronologies to provide spatially explicit differentiations of avalanche frequency and temporally explicit records of avalanche extent that are often lacking. The study area - part of John F. Stevens Canyon on the southern border of <span class="hlt">Glacier</span> National <span class="hlt">Park</span> – is within a heavily used railroad and highway corridor with two dozen active avalanche paths. Using a spatially geo-referenced network of avalanche-damaged trees (n=109) from a single path, we reconstructed a 96-year tree-ring based chronology of avalanche extent and frequency. Comparison of the chronology with historic records revealed that trees recorded all known events as well as the same number of previously unidentified events. Kriging methods provided spatially explicit estimates of avalanche return periods. Estimated return periods for the entire avalanche path averaged 3.2 years. Within this path, return intervals ranged from ~2.3 yrs in the lower track, to ~9-11 yrs and ~12 to >25 yrs in the runout zone, where the railroad and highway are located. For avalanche professionals, engineers, and transportation managers this technique proves a powerful tool in landscape risk assessment and decision making.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMPA54A..06M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMPA54A..06M"><span>Developing Aesthetically Compelling Visualizations for Documenting and Communicating Alaskan <span class="hlt">Glacier</span> and Landscape Change</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molnia, B. F.</p> <p>2016-12-01</p> <p>For 50 years I have investigated <span class="hlt">glacier</span> dynamics and attempted to convey this information to others. Since 2000, my focus has been on capturing and documenting decadal and century-scale Alaskan <span class="hlt">glacier</span> and landscape change using precision repeat photography and on broadly communicate these results through simple, aesthetically compelling, unambiguous visualizations. As a young geologist, I spent the summer of 1968 on the Juneau Icefield, photographing its surface features and margins. Since then, I have taken 150,000 photographs of Alaskan <span class="hlt">glaciers</span> and collected 5,000 historical Alaskan photographs taken by other, the earliest dating back to 1883. This database and my passion for photographing <span class="hlt">glaciers</span> became the basis for an on-going investigation aimed at visually documenting <span class="hlt">glacier</span> and landscapes change at more than 200 previously photographed Alaskan locations in <span class="hlt">Glacier</span> Bay and Kenai Fjords National <span class="hlt">Parks</span>, Prince William Sound, and the Coast Mountains. Repeat photography is a technique in which a historical and a modern photograph, both having similar fields of view, are compared and contrasted to quantitatively and qualitatively determine their similarities and differences. In precision repeat photography, both photographs have the same field of view, ideally being photographed from the identical location. Since 2000, I have conducted nearly 20 field campaigns to systematically revisit and re-photograph more than 225 fields of view previously captured in the historical photographs. As aesthetics are important in successfully communicating what has changed, substantial time and effort is invested in capturing new, comparable, generally cloud free photographs at each revisited site. The resulting modern images are then paired with similar field-of-view historical images to produce compelling, aesthetic photo pairs which depict long-term <span class="hlt">glacier</span>, landscape, and ecosystem changes. As a few sites have multiple historical images, photo triplets or quadruplets are</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3387152','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3387152"><span>Protected Areas: Mixed Success in Conserving <span class="hlt">East</span> Africa’s Evergreen Forests</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Pfeifer, Marion; Burgess, Neil D.; Swetnam, Ruth D.; Platts, Philip J.; Willcock, Simon; Marchant, Robert</p> <p>2012-01-01</p> <p>In <span class="hlt">East</span> Africa, human population growth and demands for natural resources cause forest loss contributing to increased carbon emissions and reduced biodiversity. Protected Areas (PAs) are intended to conserve habitats and species. Variability in PA effectiveness and ‘leakage’ (here defined as displacement of deforestation) may lead to different trends in forest loss within, and adjacent to, existing PAs. Here, we quantify spatial variation in trends of evergreen forest coverage in <span class="hlt">East</span> Africa between 2001 and 2009, and test for correlations with forest accessibility and environmental drivers. We investigate PA effectiveness at local, landscape and national scales, comparing rates of deforestation within <span class="hlt">park</span> boundaries with those detected in <span class="hlt">park</span> buffer zones and in unprotected land more generally. Background forest loss (BFL) was estimated at −9.3% (17,167 km2), but varied between countries (range: −0.9% to −85.7%; note: no BFL in South Sudan). We document high variability in PA effectiveness within and between PA categories. The most successful PAs were National <span class="hlt">Parks</span>, although only 26 out of 48 <span class="hlt">parks</span> increased or maintained their forest area (i.e. Effective <span class="hlt">parks</span>). Forest Reserves (Ineffective <span class="hlt">parks</span>, i.e. <span class="hlt">parks</span> that lose forest from within boundaries: 204 out of 337), Nature Reserves (six out of 12) and Game <span class="hlt">Parks</span> (24 out of 26) were more likely to lose forest cover. Forest loss in buffer zones around PAs exceeded background forest loss, in some areas indicating leakage driven by Effective National <span class="hlt">Parks</span>. Human pressure, forest accessibility, protection status, distance to fires and long-term annual rainfall were highly significant drivers of forest loss in <span class="hlt">East</span> Africa. Some of these factors can be addressed by adjusting <span class="hlt">park</span> management. However, addressing close links between livelihoods, natural capital and poverty remains a fundamental challenge in <span class="hlt">East</span> Africa’s forest conservation efforts. PMID:22768074</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2010/1283/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2010/1283/"><span>Development of monitoring protocols to detect change in rocky intertidal communities of <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Irvine, Gail V.</p> <p>2010-01-01</p> <p><span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve in southeastern Alaska includes extensive coastlines representing a major proportion of all coastlines held by the National <span class="hlt">Park</span> Service. The marine plants and invertebrates that occupy intertidal shores form highly productive communities that are ecologically important to a number of vertebrate and invertebrate consumers and that are vulnerable to human disturbances. To better understand these communities and their sensitivity, it is important to obtain information on species abundances over space and time. During field studies from 1997 to 2001, I investigated probability-based rocky intertidal monitoring designs that allow inference of results to similar habitat within the bay and that reduce bias. Aerial surveys of a subset of intertidal habitat indicated that the original target habitat of bedrock-dominated sites with slope less than or equal to 30 degrees was rare. This finding illustrated the value of probability-based surveys and led to a shift in the target habitat type to more mixed rocky habitat with steeper slopes. Subsequently, I investigated different sampling methods and strategies for their relative power to detect changes in the abundances of the predominant sessile intertidal taxa: barnacles -Balanomorpha, the mussel Mytilus trossulus and the rockweed Fucus distichus subsp. evanescens. I found that lower-intensity sampling of 25 randomly selected sites (= coarse-grained sampling) provided a greater ability to detect changes in the abundances of these taxa than did more intensive sampling of 6 sites (= fine-grained sampling). Because of its greater power, the coarse-grained sampling scheme was adopted in subsequent years. This report provides detailed analyses of the 4 years of data and evaluates the relative effect of different sampling attributes and management-set parameters on the ability of the sampling to detect changes in the abundances of these taxa. The intent was to provide managers with information</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('http://adsabs.harvard.edu/abs/2014E%26PSL.399...52S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014E%26PSL.399...52S"><span>Active water exchange and life near the grounding line of an Antarctic outlet <span class="hlt">glacier</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sugiyama, Shin; Sawagaki, Takanobu; Fukuda, Takehiro; Aoki, Shigeru</p> <p>2014-08-01</p> <p>The grounding line (GL) of the Antarctic ice sheet forms the boundary between grounded and floating ice along the coast. Near this line, warm oceanic water contacts the ice shelf, producing the ice sheet's highest basal-melt rate. Despite the importance of this region, water properties and circulations near the GL are largely unexplored because in-situ observations are difficult. Here we present direct evidence of warm ocean-water transport to the innermost part of the subshelf cavity (several hundred meters seaward from the GL) of Langhovde <span class="hlt">Glacier</span>, an outlet <span class="hlt">glacier</span> in <span class="hlt">East</span> Antarctica. Our measurements come from boreholes drilled through the <span class="hlt">glacier</span>'s ∼400-m-thick grounding zone. Beneath the grounding zone, we find a 10-24-m-deep water layer of uniform temperature and salinity (-1.45 °C; 34.25 PSU), values that roughly equal those measured in the ocean in front of the <span class="hlt">glacier</span>. Moreover, living organisms are found in the thin subglacial water layer. These findings indicate active transport of water and nutrients from the adjacent ocean, meaning that the subshelf environment interacts directly and rapidly with the ocean.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016RvGeo..54..220T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016RvGeo..54..220T"><span>Where <span class="hlt">glaciers</span> meet water: Subaqueous melt and its relevance to <span class="hlt">glaciers</span> in various settings</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Truffer, Martin; Motyka, Roman J.</p> <p>2016-03-01</p> <p><span class="hlt">Glacier</span> change is ubiquitous, but the fastest and largest magnitude changes occur in <span class="hlt">glaciers</span> that terminate in water. This includes the most rapidly retreating <span class="hlt">glaciers</span>, and also several advancing ones, often in similar regional climate settings. Furthermore, water-terminating <span class="hlt">glaciers</span> show a large range in morphology, particularly when ice flow into ocean water is compared to that into freshwater lakes. All water-terminating <span class="hlt">glaciers</span> share the ability to lose significant volume of ice at the front, either through mechanical calving or direct melt from the water in contact. Here we present a review of the subaqueous melt process. We discuss the relevant physics and show how different physical settings can lead to different glacial responses. We find that subaqueous melt can be an important trigger for <span class="hlt">glacier</span> change. It can explain many of the morphological differences, such as the existence or absence of floating tongues. Subaqueous melting is influenced by glacial runoff, which is largely a function of atmospheric conditions. This shows a tight connection between atmosphere, oceans and lakes, and <span class="hlt">glaciers</span>. Subaqueous melt rates, even if shown to be large, should always be discussed in the context of ice supply to the <span class="hlt">glacier</span> front to assess its overall relevance. We find that melt is often relevant to explain seasonal evolution, can be instrumental in shifting a <span class="hlt">glacier</span> into a different dynamical regime, and often forms a large part of a <span class="hlt">glacier</span>'s mass loss. On the other hand, in some cases, melt is a small component of mass loss and does not significantly affect <span class="hlt">glacier</span> response.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H43C1654B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H43C1654B"><span>Response of small <span class="hlt">glaciers</span> to climate change: runoff from <span class="hlt">glaciers</span> of the Wind River range, Wyoming</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bliss, A. K.; Stamper, B.</p> <p>2017-12-01</p> <p>Runoff from <span class="hlt">glaciers</span> affects downstream ecosystems by influencing the quantity, seasonality, and chemistry of the water. We describe the present state of <span class="hlt">glaciers</span> in the Wind River range, Wyoming and consider how these <span class="hlt">glaciers</span> will change in the future. Wind River <span class="hlt">glaciers</span> have been losing mass in recent decades, as seen with geodetic techniques and by examining <span class="hlt">glacier</span> morphology. Interestingly, the 2016/7 winter featured one of the largest snowfalls on record. Our primary focus is the Dinwoody <span class="hlt">Glacier</span> ( 3 km^2, 3300-4000 m above sea level). We present data collected in mid-August 2017 including <span class="hlt">glacier</span> ablation rates, snow line elevations, and streamflow. We compare measured <span class="hlt">glacier</span> mass loss to streamflow at the <span class="hlt">glacier</span> terminus and at a USGS stream gauge farther downstream. Using a hydrological model, we explore the fate of glacial runoff as it moves into downstream ecosystems and through ranchlands important to local people. The techniques used here can be applied to similar small-<span class="hlt">glacier</span> systems in other parts of the world.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.C13E..01B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C13E..01B"><span>Ocean forcing drives <span class="hlt">glacier</span> retreat sometimes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bassis, J. N.; Ultee, E.; Ma, Y.</p> <p>2015-12-01</p> <p>Observations show that marine-terminating <span class="hlt">glaciers</span> respond to climate forcing nonlinearly, with periods of slow or negligible <span class="hlt">glacier</span> advance punctuated by abrupt, rapid retreat. Once <span class="hlt">glacier</span> retreat has initiated, <span class="hlt">glaciers</span> can quickly stabilize with a new terminus position. Alternatively, retreat can be sustained for decades (or longer), as is the case for Columbia <span class="hlt">Glacier</span>, Alaska where retreat initiated ~1984 and continues to this day. Surprisingly, patterns of <span class="hlt">glacier</span> retreat show ambiguous or even contradictory correlations with atmospheric temperature and <span class="hlt">glacier</span> surface mass balance. Despite these puzzles, observations increasingly show that intrusion of warm subsurface ocean water into fjords can lead to <span class="hlt">glacier</span> erosion rates that can account for a substantial portion of the total mass lost from <span class="hlt">glaciers</span>. Here we use a simplified flowline model to show that even relatively modest submarine melt rates (~100 m/a) near the terminus of grounded <span class="hlt">glaciers</span> can trigger large increases in iceberg calving leading to rapid <span class="hlt">glacier</span> retreat. However, the strength of the coupling between submarine melt and calving is a strong function of the geometry of the <span class="hlt">glacier</span> (bed topography, ice thickness and <span class="hlt">glacier</span> width). This can lead to irreversible retreat when the terminus is thick and grounded deeply beneath sea level or result in little change when the <span class="hlt">glacier</span> is relatively thin, grounded in shallow water or pinned in a narrow fjord. Because of the strong dependence on <span class="hlt">glacier</span> geometry, small perturbations in submarine melting can trigger <span class="hlt">glaciers</span> in their most advanced—and geometrically precarious—state to undergo sudden retreat followed by much slower re-advance. Although many details remain speculative, our model hints that some <span class="hlt">glaciers</span> are more sensitive than others to ocean forcing and that some of the nonlinearities of <span class="hlt">glacier</span> response to climate change may be attributable to variations in difficult-to-detect subsurface water temperatures that need to be better</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012TCD.....6.4557P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012TCD.....6.4557P"><span>Quantifying present and future <span class="hlt">glacier</span> melt-water contribution to runoff in a Central Himalayan river basin</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Prasch, M.; Mauser, W.; Weber, M.</p> <p>2012-10-01</p> <p>Water supply of most lowland cultures heavily depends on rain and melt-water from the upstream mountains. Especially melt-water release of alpine mountain ranges is usually attributed a pivotal role for the water supply of large downstream regions. Water scarcity is assumed as consequence of <span class="hlt">glacier</span> shrinkage and possible disappearance due to Global Climate Change, particular for large parts of Central and South <span class="hlt">East</span> Asia. In this paper, the application and validation of a coupled modeling approach with Regional Climate Model outputs and a process-oriented <span class="hlt">glacier</span> and hydrological model is presented for a Central Himalayan river basin despite scarce data availability. Current and possible future contributions of ice-melt to runoff along the river network are spatially explicitly shown. Its role among the other water balance components is presented. Although <span class="hlt">glaciers</span> have retreated and will continue to retreat according to the chosen climate scenarios, water availability is and will be primarily determined by monsoon precipitation and snow-melt. Ice-melt from <span class="hlt">glaciers</span> is and will be a minor runoff component in summer monsoon-dominated Himalayan river basins.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/949957','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/949957"><span>Environmental Management Waste Management Facility Waste Lot Profile 155.5 for K-1015-A Laundry Pit, <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> Oak Ridge, Tennessee</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>Bechtel Jacobs, Raymer J.E.</p> <p>2008-06-12</p> <p>In 1989, the Oak Ridge Reservation (ORR), which includes the <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> (ETTP), was placed on the Comprehensive Environmental Response, Compensation, and Liability Act of 1980 (CERCLA) National Priorities List. The Federal Facility Agreement (FFA) (DOE 1992), effective January 1, 1992, now governs environmental restoration activities conducted under CERCLA at the ORR. Following signing of the FFA, U.S. Department of Energy (DOE), U.S. Environmental Protection Agency (EPA), and the state of Tennessee signed the Oak Ridge Accelerated Cleanup Plan Agreement on June 18, 2003. The purpose of this agreement is to define a streamlined decision-making process to facilitatemore » the accelerated implementation of cleanup, to resolve ORR milestone issues, and to establish future actions necessary to complete the accelerated cleanup plan by the end of fiscal year 2008. While the FFA continues to serve as the overall regulatory framework for remediation, the Accelerated Cleanup Plan Agreement supplements existing requirements to streamline the decision-making process. The disposal of the K-1015 Laundry Pit waste will be executed in accordance with the 'Record of Decision for Soil, Buried Waste, and Subsurface Structure Actions in Zone, 2, <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee' (DOB/ORAH-2161&D2) and the 'Waste Handling Plan for the Consolidated Soil and Waste Sites with Zone 2, <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee' (DOE/OR/01-2328&D1). This waste lot consists of a total of approximately 50 cubic yards of waste that will be disposed at the Environmental Management Waste Management Facility (EMWMF) as non-containerized waste. This material will be sent to the EMWMF in dump trucks. This profile is for the K-1015-A Laundry Pit and includes debris (e.g., concrete, metal rebar, pipe), incidental soil, plastic and wood, and secondary waste (such as plastic sheeting, hay bales and other erosion control materials, wooden pallets</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2013/1097/pdf/ofr20131097.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2013/1097/pdf/ofr20131097.pdf"><span>Mercury and water-quality data from Rink Creek, Salmon River, and Good River, <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, Alaska, November 2009-October 2011</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Nagorski, Sonia A.; Neal, Edward G.; Brabets, Timothy P.</p> <p>2013-01-01</p> <p><span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve (GBNPP), Alaska, like many pristine high latitude areas, is exposed to atmospherically deposited contaminants such as mercury (Hg). Although the harmful effects of Hg are well established, information on this contaminant in southeast Alaska is scarce. Here, we assess the level of this contaminant in several aquatic components (water, sediments, and biological tissue) in three adjacent, small streams in GBNPP that drain contrasting landscapes but receive similar atmospheric inputs: Rink Creek, Salmon River, and Good River. Twenty water samples were collected from 2009 to 2011 and processed and analyzed for total mercury and methylmercury (filtered and particulate), and dissolved organic carbon quantity and quality. Ancillary stream water parameters (discharge, pH, dissolved oxygen, specific conductance, and temperature) were measured at the time of sampling. Major cations, anions, and nutrients were measured four times. In addition, total mercury was analyzed in streambed sediment in 2010 and in juvenile coho salmon and several taxa of benthic macroinvertebrates in the early summer of 2010 and 2011.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMEP53C0968R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMEP53C0968R"><span>CRN Dating and Numerical <span class="hlt">Glacier</span> Modeling to Investigate Climate During the Last Glacial Maximum, and the Subsequent Deglaciation, Sawatch Range, Colorado</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Russell, C.; Leonard, E. M.</p> <p>2016-12-01</p> <p> the wetter west side of the range maintained small <span class="hlt">east</span>-side valley <span class="hlt">glaciers</span> even as the <span class="hlt">east</span>-side cirques deglaciated. Ongoing work will model a larger area of range to gain a better understanding of range-wide patterns of ice flow that could have affected deglaciation of the Lake Creek valley.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/FR-2013-02-26/pdf/2013-04342.pdf','FEDREG'); return false;" href="https://www.gpo.gov/fdsys/pkg/FR-2013-02-26/pdf/2013-04342.pdf"><span>78 FR 13081 - Draft Environmental Impact Statement for General Management Plan, Everglades National <span class="hlt">Park</span>, Florida</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>2013-02-26</p> <p>... the General Management Plan (GMP) and <span class="hlt">East</span> Everglades Wilderness Study (EEWS) for Everglades National <span class="hlt">Park</span> (<span class="hlt">park</span>). After it is finalized, the GMP/EEWS will guide the management of the <span class="hlt">park</span> over the next 20... visitor use in the <span class="hlt">Park</span>. The GMP will provide updated management direction for the entire <span class="hlt">park</span>. The EEWS...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.C31B0298T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.C31B0298T"><span>Recent thinning of Bowdoin <span class="hlt">Glacier</span>, a marine terminating outlet <span class="hlt">glacier</span> in northwestern Greenland</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsutaki, S.; Sugiyama, S.; Sakakibara, D.; Sawagaki, T.; Maruyama, M.</p> <p>2014-12-01</p> <p>Ice discharge from calving <span class="hlt">glaciers</span> has increased in the Greenland ice sheet (GrIS), and this increase plays important roles in the volume change of GrIS and its contribution to sea level rise. Thinning of GrIS calving <span class="hlt">glaciers</span> has been studied by the differentiation of digital elevation models (DEMs) derived by satellite remote-sensing (RS). Such studies rely on the accuracy of DEMs, but calibration of RS data with ground based data is difficult. This is because field data on GrIS calving <span class="hlt">glaciers</span> are few. In this study, we combined field and RS data to measure surface elevation change of Bowdoin <span class="hlt">Glacier</span>, a marine terminating outlet <span class="hlt">glacier</span> in northwestern Greenland (77°41'18″N, 68°29'47″W). The fast flowing part of the <span class="hlt">glacier</span> is approximately 3 km wide and 10 km long. Ice surface elevation within 6 km from the <span class="hlt">glacier</span> terminus was surveyed in the field in July 2013 and 2014, by using the global positioning system. We also measured the surface elevation over the <span class="hlt">glacier</span> on August 20, 2007 and September 4, 2010, by analyzing Advanced Land Observing Satellite (ALOS), Panchromatic remote-sensing Instrument for Stereo Mapping (PRISM) images. We calibrated the satellite derived elevation data with our field measurements, and generated DEM for each year with a 25 m grid mesh. The field data and DEMs were compared to calculate recent <span class="hlt">glacier</span> elevation change. Mean surface elevation change along the field survey profiles were -16.3±0.2 m (-5.3±0.1 m yr-1) in 2007-2010 and -10.8±0.2 m (-3.8±0.1 m yr-1) in 2010-2013. These rates are much greater than those observed on non-calving ice caps in the region, and similar to those reported for other calving <span class="hlt">glaciers</span> in northwestern Greenland. Loss of ice was greater near the <span class="hlt">glacier</span> terminus, suggesting the importance of ice dynamics and/or interaction with the ocean.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990099258&hterms=balance+sheet&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dbalance%2Bsheet','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990099258&hterms=balance+sheet&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dbalance%2Bsheet"><span>Velocity Estimates of Fast-Moving Outlet <span class="hlt">Glaciers</span> on the Greenland Ice Sheet</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Abdalati, Waleed; Krabill, W. B.</p> <p>1998-01-01</p> <p>In recent years, airborne laser altimetry has been used with great success to investigate the mass balance characteristics of the Greenland ice sheet. One spinoff of this activity has been the application of these measurements to the study of surface velocities in some of Greenland's fast-moving drainage <span class="hlt">glaciers</span>. This is accomplished by tracking the motion of elevation features, primarily crevasses, in pairs of aircraft laser altimetry surveys. Detailed elevation measurements are made along or across <span class="hlt">glaciers</span> of interest with a scanning swath of 150 to 200 meters, and the surveys are repeated several days later, typically to within better than 50 meters of the previous flight line. Surface elevation features are identified in each image, and their offsets are compared yielding detailed velocities over narrow regions. During the 1998 field season, repeat flights were made over three <span class="hlt">glaciers</span> for the purpose of estimating their surface velocities. These were the Kangerdlugssuaq and Helheim <span class="hlt">glaciers</span> on the <span class="hlt">east</span> coast and the Jakobshavn Isbrae on the west coast. Each flows at such high speeds (on the order of a few kilometers per year) that their flow rates are difficult to assess by means of radar interferometry. The flexibility of the aircraft platform, however, allows for detailed measurements of the elevation and flow of these drainage areas, which are responsible for a significant portion of the ice discharge from the Greenland ice sheet. Velocity estimates for transects that span these <span class="hlt">glaciers</span> will be presented, and where the ice thickness values are available (provided by researchers from the University of Kansas) the fluxes will be calculated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017HESS...21.3249B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017HESS...21.3249B"><span>Assessing <span class="hlt">glacier</span> melt contribution to streamflow at Universidad <span class="hlt">Glacier</span>, central Andes of Chile</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bravo, Claudio; Loriaux, Thomas; Rivera, Andrés; Brock, Ben W.</p> <p>2017-07-01</p> <p><span class="hlt">Glacier</span> melt is an important source of water for high Andean rivers in central Chile, especially in dry years, when it can be an important contributor to flows during late summer and autumn. However, few studies have quantified <span class="hlt">glacier</span> melt contribution to streamflow in this region. To address this shortcoming, we present an analysis of meteorological conditions and ablation for Universidad <span class="hlt">Glacier</span>, one of the largest valley <span class="hlt">glaciers</span> in the central Andes of Chile at the head of the Tinguiririca River, for the 2009-2010 ablation season. We used meteorological measurements from two automatic weather stations installed on the <span class="hlt">glacier</span> to drive a distributed temperature-index and runoff routing model. The temperature-index model was calibrated at the lower weather station site and showed good agreement with melt estimates from an ablation stake and sonic ranger, and with a physically based energy balance model. Total modelled <span class="hlt">glacier</span> melt is compared with river flow measurements at three sites located between 0.5 and 50 km downstream. Universidad <span class="hlt">Glacier</span> shows extremely high melt rates over the ablation season which may exceed 10 m water equivalent in the lower ablation area, representing between 10 and 13 % of the mean monthly streamflow at the outlet of the Tinguiririca River Basin between December 2009 and March 2010. This contribution rises to a monthly maximum of almost 20 % in March 2010, demonstrating the importance of <span class="hlt">glacier</span> runoff to streamflow, particularly in dry years such as 2009-2010. The temperature-index approach benefits from the availability of on-<span class="hlt">glacier</span> meteorological data, enabling the calculation of the local hourly variable lapse rate, and is suited to high melt regimes, but would not be easily applicable to <span class="hlt">glaciers</span> further north in Chile where sublimation is more significant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://ngmdb.usgs.gov/Prodesc/proddesc_78283.htm','USGSPUBS'); return false;" href="http://ngmdb.usgs.gov/Prodesc/proddesc_78283.htm"><span>Geologic map of the Wrangell-Saint Elias National <span class="hlt">Park</span> and Reserve, Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Richter, Donald H.; Preller, Cindi C.; Labay, Keith A.; Shew, Nora B.</p> <p>2006-01-01</p> <p>Wrangell-Saint Elias National <span class="hlt">Park</span> and Preserve, the largest national <span class="hlt">park</span> within the U.S. National <span class="hlt">Park</span> Service system, extends from the northern Pacific Ocean to beyond the eastern Alaska Range into interior Alaska. It features impressively spectacular scenery such as high and craggy mountains, active and ancient volcanoes, expansive ice fields, immense tidewater <span class="hlt">glaciers</span>, and a myriad of alpine <span class="hlt">glaciers</span>. The <span class="hlt">park</span> also includes the famous Kennecott Mine, a world-class copper deposit that was mined from 1911 to 1938, and remnant ghost town, which is now a National Historic Landmark. Geologic investigations encompassing Wrangell-Saint Elias National <span class="hlt">Park</span> and Preserve began in 1796, with Dmitriv Tarkhanov, a Russian mining engineer, who unsuccessfully ventured up the Copper River in search of rumored copper. Lieutenant H.T. Allen (1897) of the U.S. Army made a successful epic summer journey with a limited military crew up the Copper River in 1885, across the Alaska Range, and down the Tanana and Yukon Rivers. Allen?s crew was supported by a prospector named John Bremner and local Eyak and Ahtna native guides whose tribes controlled access into the Copper River basin. Allen witnessed the Ahtnas? many uses of the native copper. His stories about the copper prompted prospectors to return to this area in search of the rich copper ore in the years following his journey. The region boasts a rich mining and exploration history prior to becoming a <span class="hlt">park</span> in 1980. Several U.S. Geological Survey geologists have conducted reconnaissance surveys in the area since Allen?s explorations. This map is the result of their work and is enhanced by more detailed investigations, which began in the late 1950s and are still continuing. For a better understanding of the processes that have shaped the geology of the <span class="hlt">park</span> and a history of the geologic investigations in the area, we recommend U.S. Geological Survey Professional Paper 1616, ?A Geologic Guide to Wrangell-Saint Elias National <span class="hlt">Park</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/p1616/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/p1616/"><span>A geologic guide to Wrangell-Saint Elias National <span class="hlt">Park</span> and Preserve, Alaska; a tectonic collage of northbound terranes</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Winkler, Gary R.; with contributions by MacKevett, E. M.; Plafker, George; Richter, D.H.; Rosenkrans, D.S.; Schmoll, H.R.</p> <p>2000-01-01</p> <p>Wrangell-Saint Elias National <span class="hlt">Park</span> and Preserve, the largest unit in the U.S. National <span class="hlt">Park</span> System, encompasses near 13.2 million acres of geological wonderments. This geologic guide presents history of exploration and Earth-science investigation; describes the complex geologic makeup; characterizes the vast college of accretion geologic terranes in this area of Alaska's continental margin; recapitulates the effects of earthquakes, volcanoes, and <span class="hlt">glaciers</span>; characterizes the copper and gold resources of the parklands; and describes outstanding locales within the <span class="hlt">park</span> and preserve area. A glossary of geologic terms and a categorized list of additional sources of information complete this report.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6594874-malaspina-glacier-modern-analog-laurentide-glacier-new-england','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6594874-malaspina-glacier-modern-analog-laurentide-glacier-new-england"><span>Malaspina <span class="hlt">Glacier</span>: a modern analog to the Laurentide <span class="hlt">Glacier</span> in New England</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>Gustavson, T.C.; Boothroyd, J.C.</p> <p>1985-01-01</p> <p>The land-based temperate Malaspina <span class="hlt">Glacier</span> is a partial analog to the late Wisconsinan Laurentide Ice Sheet that occupied New England and adjacent areas. The Malaspina occupies a bedrock basin similar to basins occupied by the margin of the Laurentide Ice Sheet. Ice lobes of the Malaspina are similar in size to end moraine lobes in southern New England and Long Island,New York. Estimated ice temperature, ablation rates, surface slopes and meltwater discharge per unit of surface area for the Laurentide Ice Sheet are similar to those for the Malaspina <span class="hlt">Glacier</span>. In a simple hydrologic-fluvial model for the Malaspina <span class="hlt">Glacier</span> meltwatermore » moves towards the <span class="hlt">glacier</span> bed and down-<span class="hlt">glacier</span> along intercrystalline pathways, crevasses and moulins, and a series of tunnels. Regolith and bedrock at the <span class="hlt">glacier</span> floor, which are eroded and transported by subglacial and englacial streams, are the sources of essentially all fluvio-lacustrine sediment on the Malaspina Foreland. Supraglacial eskers containing coarse gravels occur as much as 100 m above the <span class="hlt">glacier</span> bed and are evidence that bedload can be lifted hydraulically. Subordinant amounts of sediment are contributed to outwash by small surface streams draining the ice margin. By analogy a similar hydrologic-fluvial system existed along the southeastern margin of the Laurentide Ice Sheet. Subglacial regolith and bedrock eroded from beneath the Laurentide Ice Sheet by meltwater was also the source of most glaciofluvial and glaciolacustrine deposits in southern New England, not sediment carried to the surface of the ice sheet along shear planes and washed off the <span class="hlt">glacier</span> by meltwater.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NatGe..11..258K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NatGe..11..258K"><span>Net retreat of Antarctic <span class="hlt">glacier</span> grounding lines</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Konrad, Hannes; Shepherd, Andrew; Gilbert, Lin; Hogg, Anna E.; McMillan, Malcolm; Muir, Alan; Slater, Thomas</p> <p>2018-04-01</p> <p>Grounding lines are a key indicator of ice-sheet instability, because changes in their position reflect imbalance with the surrounding ocean and affect the flow of inland ice. Although the grounding lines of several Antarctic <span class="hlt">glaciers</span> have retreated rapidly due to ocean-driven melting, records are too scarce to assess the scale of the imbalance. Here, we combine satellite altimeter observations of ice-elevation change and measurements of ice geometry to track grounding-line movement around the entire continent, tripling the coverage of previous surveys. Between 2010 and 2016, 22%, 3% and 10% of surveyed grounding lines in West Antarctica, <span class="hlt">East</span> Antarctica and at the Antarctic Peninsula retreated at rates faster than 25 m yr-1 (the typical pace since the Last Glacial Maximum) and the continent has lost 1,463 km2 ± 791 km2 of grounded-ice area. Although by far the fastest rates of retreat occurred in the Amundsen Sea sector, we show that the Pine Island <span class="hlt">Glacier</span> grounding line has stabilized, probably as a consequence of abated ocean forcing. On average, Antarctica's fast-flowing ice streams retreat by 110 metres per metre of ice thinning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.C21B0731R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C21B0731R"><span><span class="hlt">Glacier</span> Surface Lowering and Stagnation in the Manaslu Region of Nepal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Robson, B. A.; Nuth, C.; Nielsen, P. R.; Hendrickx, M.; Dahl, S. O.</p> <p>2015-12-01</p> <p>Frequent and up-to-date <span class="hlt">glacier</span> outlines are needed for many applications of glaciology, not only <span class="hlt">glacier</span> area change analysis, but also for masks in volume or velocity analysis, for the estimation of water resources and as model input data. Remote sensing offers a good option for creating <span class="hlt">glacier</span> outlines over large areas, but manual correction is frequently necessary, especially in areas containing supraglacial debris. We show three different workflows for mapping clean ice and debris-covered ice within Object Based Image Analysis (OBIA). By working at the object level as opposed to the pixel level, OBIA facilitates using contextual, spatial and hierarchical information when assigning classes, and additionally permits the handling of multiple data sources. Our first example shows mapping debris-covered ice in the Manaslu Himalaya, Nepal. SAR Coherence data is used in combination with optical and topographic data to classify debris-covered ice, obtaining an accuracy of 91%. Our second example shows using a high-resolution LiDAR derived DEM over the Hohe Tauern National <span class="hlt">Park</span> in Austria. Breaks in surface morphology are used in creating image objects; debris-covered ice is then classified using a combination of spectral, thermal and topographic properties. Lastly, we show a completely automated workflow for mapping <span class="hlt">glacier</span> ice in Norway. The NDSI and NIR/SWIR band ratio are used to map clean ice over the entire country but the thresholds are calculated automatically based on a histogram of each image subset. This means that in theory any Landsat scene can be inputted and the clean ice can be automatically extracted. Debris-covered ice can be included semi-automatically using contextual and morphological information.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://geology.geoscienceworld.org/content/38/4/319','USGSPUBS'); return false;" href="http://geology.geoscienceworld.org/content/38/4/319"><span><span class="hlt">Glacier</span> microseismicity</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>West, Michael E.; Larsen, Christopher F.; Truffer, Martin; O'Neel, Shad; LeBlanc, Laura</p> <p>2010-01-01</p> <p>We present a framework for interpreting small <span class="hlt">glacier</span> seismic events based on data collected near the center of Bering <span class="hlt">Glacier</span>, Alaska, in spring 2007. We find extremely high microseismicity rates (as many as tens of events per minute) occurring largely within a few kilometers of the receivers. A high-frequency class of seismicity is distinguished by dominant frequencies of 20–35 Hz and impulsive arrivals. A low-frequency class has dominant frequencies of 6–15 Hz, emergent onsets, and longer, more monotonic codas. A bimodal distribution of 160,000 seismic events over two months demonstrates that the classes represent two distinct populations. This is further supported by the presence of hybrid waveforms that contain elements of both event types. The high-low-hybrid paradigm is well established in volcano seismology and is demonstrated by a comparison to earthquakes from Augustine Volcano. We build on these parallels to suggest that fluid-induced resonance is likely responsible for the low-frequency <span class="hlt">glacier</span> events and that the hybrid <span class="hlt">glacier</span> events may be caused by the rush of water into newly opening pathways.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1913170F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1913170F"><span>Development of Adygine <span class="hlt">glacier</span> complex (<span class="hlt">glacier</span> and proglacial lakes) and its link to outburst hazard</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Falatkova, Kristyna; Schöner, Wolfgang; Häusler, Hermann; Reisenhofer, Stefan; Neureiter, Anton; Sobr, Miroslav; Jansky, Bohumir</p> <p>2017-04-01</p> <p>Mountain <span class="hlt">glacier</span> retreat has a well-known impact on life of local population - besides anxiety over water supply for agriculture, industry, or households, it has proved to have a direct influence on <span class="hlt">glacier</span> hazard occurrence. The paper focuses on lake outburst hazard specifically, and aims to describe the previous and future development of Adygine <span class="hlt">glacier</span> complex and identify its relationship to the hazard. The observed <span class="hlt">glacier</span> is situated in the Northern Tien Shan, with an area of 4 km2 in northern exposition at an elevation range of 3,500-4,200 m a.s.l. The study <span class="hlt">glacier</span> ranks in the group of small-sized <span class="hlt">glaciers</span>, therefore we expect it to respond faster to changes of the climate compared to larger ones. Below the <span class="hlt">glacier</span> there is a three-level cascade of proglacial lakes at different stages of development. The site has been observed sporadically since 1960s, however, closer study has been carried out since 2007. Past development of the <span class="hlt">glacier</span>-lake complex is analyzed by combination of satellite imagery interpretations and on-site measurements (geodetic and bathymetric survey). A <span class="hlt">glacier</span> mass balance model is used to simulate future development of the <span class="hlt">glacier</span> resulting from climate scenarios. We used the simulated future <span class="hlt">glacier</span> extent and the <span class="hlt">glacier</span> base topography provided by GPR survey to assess potential for future lake formation. This enables us to assess the outburst hazard for the three selected lakes with an outlook for possible/probable hazard changes linked to further complex succession/progression (originating from climate change scenarios). Considering the proximity of the capital Bishkek, spreading settlements, and increased demand for tourism-related infrastructure within the main valley, it is of high importance to identify the present and possible future hazards that have a potential to affect this region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/0544d/pp544d_text.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/0544d/pp544d_text.pdf"><span>Effects of the March 1964 Alaska earthquake on <span class="hlt">glaciers</span>: Chapter D in The Alaska earthquake, March 27, 1964: effects on hydrologic regimen</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Post, Austin</p> <p>1967-01-01</p> <p>The 1964 Alaska earthquake occurred in a region where there are many hundreds of <span class="hlt">glaciers</span>, large and small. Aerial photographic investigations indicate that no snow and ice avalanches of large size occurred on <span class="hlt">glaciers</span> despite the violent shaking. Rockslide avalanches extended onto the <span class="hlt">glaciers</span> in many localities, seven very large ones occurring in the Copper River region 160 kilometers <span class="hlt">east</span> of the epicenter. Some of these avalanches traveled several kilometers at low gradients; compressed air may have provided a lubricating layer. If long-term changes in <span class="hlt">glaciers</span> due to tectonic changes in altitude and slope occur, they will probably be very small. No evidence of large-scale dynamic response of any <span class="hlt">glacier</span> to earthquake shaking or avalanche loading was found in either the Chugach or Kenai Mountains 16 months after the 1964 earthquake, nor was there any evidence of surges (rapid advances) as postulated by the Earthquake-Advance Theory of Tarr and Martin.</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('http://adsabs.harvard.edu/abs/2016AGUFM.C13C0848C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C13C0848C"><span>Large basal crevasses as a proxy for historic subglacial flooding events on Byrd <span class="hlt">Glacier</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Child, S. F.; Stearns, L. A.; van der Veen, C. J.; Hamilton, G. S.</p> <p>2016-12-01</p> <p>Active networks of subglacial lakes have recently been found beneath the Antarctic Ice Sheet. On Byrd <span class="hlt">Glacier</span>, <span class="hlt">East</span> Antarctica, a subglacial lake outburst event in 2005/07 led to a short-lived <span class="hlt">glacier</span> acceleration. Due to the sparse record of historical observations, it is unclear how frequently these outburst events occur, and the role they play in the dynamics of Antarctic outlet <span class="hlt">glaciers</span>. Crevasses form when the tensile stress is greater than the fracture strength of ice. High extensional strain rates often exist at the grounding line where grounded ice begins to float. We hypothesize that the formation of anomalously large basal crevasses coincides with the higher strain rates observed during flooding events. In this study, we use the location of large basal crevasses ( 330 m tall), located along the floating portion of the Byrd <span class="hlt">Glacier</span> flowline, to create a timeline of past flooding events. We first model crevasse formation to demonstrate that basal crevasses likely form at the grounding line. To do this, we use linear elastic fracture mechanics (LEFM) to estimate crevasse heights based on strain rates during known flood (300-350 m) and non-flood (100-150 m) time periods at Byrd <span class="hlt">Glacier</span>'s grounding line. Basal crevasse locations and heights are determined directly from radar echograms (2011/12 CReSIS radar data and 1974/75 SPRI NSF TUD radar data) along the Byrd <span class="hlt">Glacier</span> flowline. We also use the locations of large surface depressions to infer the presence of basal crevasses. When crevasses penetrate a threshold proportion of the ice column, the overlying ice is no longer supported and a surface depression forms. We identify 22 large basal crevasses through these combined methods; the oldest crevasse likely formed 600 years ago. This research provides a framework of Antarctic subglacial flooding frequency and the effects that subglacial water drainage events have on outlet <span class="hlt">glacier</span> dynamics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA02670.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA02670.html"><span>Patagonia <span class="hlt">Glacier</span>, Chile</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2001-07-21</p> <p>This ASTER image was acquired on May 2, 2000 over the North Patagonia Ice Sheet, Chile near latitude 47 degrees south, longitude 73 degrees west. The image covers 36 x 30 km. The false color composite displays vegetation in red. The image dramatically shows a single large <span class="hlt">glacier</span>, covered with crevasses. A semi-circular terminal moraine indicates that the <span class="hlt">glacier</span> was once more extensive than at present. ASTER data are being acquired over hundreds of <span class="hlt">glaciers</span> worldwide to measure their changes over time. Since <span class="hlt">glaciers</span> are sensitive indicators of warming or cooling, this program can provide global data set critical to understand climate change. This image is located at 46.5 degrees south latitude and 73.9 degrees west longitude. http://photojournal.jpl.nasa.gov/catalog/PIA02670</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C52A..07S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C52A..07S"><span>In-Situ and Remotely-Sensed <span class="hlt">Glacier</span> Monitoring in the Rwenzori Mountains, Uganda/D.R. Congo</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Samyn, D.; Uetake, J.; Kervyn, F.</p> <p>2017-12-01</p> <p>The tropics, often coined as the heat engine of the planet, are paramount for global climatology because they are the main driver of air and moisture circulation around the Earth. Despite the remarkable global homogeneity of the tropical atmosphere, both in time and in space, some regions in the tropics are characterized by high interannual variations in precipitation numbers, contributing to unstable response in high mountain regions with regard to <span class="hlt">glacier</span> mass balance. <span class="hlt">East</span> Africa, characterized in addition by a highly variable surface topography and spatially distinct climatic regimes, represents one of these sensitive regions. Despite the growing number in recent years of studies aiming at disentangling the complex interactions between the energetic conditions, the moisture circulation and the biogeosystems in the tropics, the response of tropical African climate, and more specifically of tropical African <span class="hlt">glaciers</span>, to current global change remains poorly understood. In this context, the Rwenzori mountains, with their steep topography peaking above 5000m elevation, their <span class="hlt">glaciers</span> straddling along the equator, and their location at the divide between Atlantic and Indian Ocean flow, represent a key region for gaining insight not only into tropical <span class="hlt">glacier</span> sensitivity to past, present and future climatic variations, but also into the respective roles of temperature and moisture in modulating ice mass and energy budgets. In the Rwenzori mountains, direct measurements of <span class="hlt">glacier</span> extent and mass balance are sparse due to the inaccessibility of <span class="hlt">glaciers</span> and the logistical constrains associated with maintaining and downloading continuous records. In addition, quasi-permanent cloud cover associated with small <span class="hlt">glacier</span> size severely hinder <span class="hlt">glacier</span> monitoring from space. In this research, we rely on multi-year field mapping and multi-decadal, multi-sensor satellite and aerial imagery to discuss the recession trend of Rwenzori <span class="hlt">glaciers</span>, with a view to provide an updated</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001485.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001485.html"><span><span class="hlt">Glaciers</span> and Sea Level Rise</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Calving front of the Perito Moreno <span class="hlt">Glacier</span> (Argentina). Contrary to the majority of the <span class="hlt">glaciers</span> from the southern Patagonian ice field, the Perito Moreno <span class="hlt">Glacier</span> is currently stable. It is also one of the most visited <span class="hlt">glaciers</span> in the world. To learn about the contributions of <span class="hlt">glaciers</span> to sea level rise, visit: www.nasa.gov/topics/earth/features/<span class="hlt">glacier</span>-sea-rise.html Credit: Etienne Berthier, Université de Toulouse 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA13382.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA13382.html"><span>Susitna <span class="hlt">Glacier</span>, Alaska</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2010-09-13</p> <p>Folds in the lower reaches of valley <span class="hlt">glaciers</span> can be caused by powerful surges of tributary ice streams. This phenomenon is spectacularly displayed by the Sustina <span class="hlt">Glacier</span> in the Alaska Range as seen by NASA Terra spacecraft.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C33D1234M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C33D1234M"><span>The Open Global <span class="hlt">Glacier</span> Model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marzeion, B.; Maussion, F.</p> <p>2017-12-01</p> <p>Mountain <span class="hlt">glaciers</span> are one of the few remaining sub-systems of the global climate system for which no globally applicable, open source, community-driven model exists. Notable examples from the ice sheet community include the Parallel Ice Sheet Model or Elmer/Ice. While the atmospheric modeling community has a long tradition of sharing models (e.g. the Weather Research and Forecasting model) or comparing them (e.g. the Coupled Model Intercomparison Project or CMIP), recent initiatives originating from the glaciological community show a new willingness to better coordinate global research efforts following the CMIP example (e.g. the <span class="hlt">Glacier</span> Model Intercomparison Project or the <span class="hlt">Glacier</span> Ice Thickness Estimation Working Group). In the recent past, great advances have been made in the global availability of data and methods relevant for <span class="hlt">glacier</span> modeling, spanning <span class="hlt">glacier</span> outlines, automatized <span class="hlt">glacier</span> centerline identification, bed rock inversion methods, and global topographic data sets. Taken together, these advances now allow the ice dynamics of <span class="hlt">glaciers</span> to be modeled on a global scale, provided that adequate modeling platforms are available. Here, we present the Open Global <span class="hlt">Glacier</span> Model (OGGM), developed to provide a global scale, modular, and open source numerical model framework for consistently simulating past and future global scale <span class="hlt">glacier</span> change. Global not only in the sense of leading to meaningful results for all <span class="hlt">glaciers</span> combined, but also for any small ensemble of <span class="hlt">glaciers</span>, e.g. at the headwater catchment scale. Modular to allow combinations of different approaches to the representation of ice flow and surface mass balance, enabling a new kind of model intercomparison. Open source so that the code can be read and used by anyone and so that new modules can be added and discussed by the community, following the principles of open governance. Consistent in order to provide uncertainty measures at all realizable scales.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/41291','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/41291"><span>Preliminary bathymetry of Aialik Bay and Neoglacial changes of Aialik and Pederson <span class="hlt">glaciers</span>, Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Post, Austin</p> <p>1980-01-01</p> <p>Preliminary bathymetry (at 1:20,000 scale) and scientific studies of Aialik Bay, Alaska, by the Research Vessel Growler in 1978 disclose that the head of the bay consists of a deep basin enclosed by a terminal-moraine shoal. A much smaller basin, into which Aialik <span class="hlt">Glacier</span> discharges icebergs, is located west of two islands and a submarine ridge. Comparison of 1978 soundings with U.S. Coast and Geodetic Survey (now National Oceanic and Atmospheric Administration) data obtained in 1912 shows shoaling of about 64 feet in the deepest part of the small basin nearest the <span class="hlt">glacier</span> and of about 40 feet in the large basin. The time of retreat of Aialik <span class="hlt">Glacier</span> from the moraine bar is unknown; a faint ' trimline ' is still visible in the forest on the <span class="hlt">east</span> side of the fiord, and a carbon-14 date suggests the retreat could have taken place as recently as 1800. The time of Aialik Glcier 's neoglacial advance to the moraine is unknown. Pederson <span class="hlt">Glacier</span>, which terminates in part in a tidal lagoon or lake, has retreated about 0.90 mile from a moraine judged by Grant and Higgins to have been in contact with the ice about 1896. (USGS)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.C43C0686H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.C43C0686H"><span>Surge of a Complex <span class="hlt">Glacier</span> System - The Current Surge of the Bering-Bagley <span class="hlt">Glacier</span> System, Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Herzfeld, U. C.; McDonald, B.; Trantow, T.; Hale, G.; Stachura, M.; Weltman, A.; Sears, T.</p> <p>2013-12-01</p> <p>Understanding fast <span class="hlt">glacier</span> flow and glacial accelerations is important for understanding changes in the cryosphere and ultimately in sea level. Surge-type <span class="hlt">glaciers</span> are one of four types of fast-flowing <span class="hlt">glaciers</span> --- the other three being continuously fast-flowing <span class="hlt">glaciers</span>, fjord <span class="hlt">glaciers</span> and ice streams --- and the one that has seen the least amount of research. The Bering-Bagley <span class="hlt">Glacier</span> System, Alaska, the largest <span class="hlt">glacier</span> system in North America, surged in 2011 and 2012. Velocities decreased towards the end of 2011, while the surge kinematics continued to expand. A new surge phase started in summer and fall 2012. In this paper, we report results from airborne observations collected in September 2011, June/July and September/October 2012 and in 2013. Airborne observations include simultaneously collected laser altimeter data, videographic data, GPS data and photographic data and are complemented by satellite data analysis. Methods range from classic interpretation of imagery to analysis and classification of laser altimeter data and connectionist (neural-net) geostatistical classification of concurrent airborne imagery. Results focus on the characteristics of surge progression in a large and complex <span class="hlt">glacier</span> system (as opposed to a small <span class="hlt">glacier</span> with relatively simple geometry). We evaluate changes in surface elevations including mass transfer and sudden drawdowns, crevasse types, accelerations and changes in the supra-glacial and englacial hydrologic system. Supraglacial water in Bering <span class="hlt">Glacier</span> during Surge, July 2012 Airborne laser altimeter profile across major rift in central Bering <span class="hlt">Glacier</span>, Sept 2011</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/38192','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/38192"><span>The relationship between whitebark pine health, cone production, and nutcracker occurrence across four National <span class="hlt">Parks</span></span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Lauren E. Barringer; Diana F. Tomback; Michael B. Wunder</p> <p>2011-01-01</p> <p>Whitebark pine (Pinus albicaulis) is declining in the central and northern Rocky Mountains from infection by the exotic pathogen Cronartium ribicola, which causes white pine blister rust, and from outbreaks of mountain pine beetle (Dendroctonus ponderosae). White pine blister rust has been present in <span class="hlt">Glacier</span> and Waterton Lakes National <span class="hlt">Parks</span> (NP) about two decades...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA03475.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA03475.html"><span>Malaspina <span class="hlt">Glacier</span>, Alaska</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2002-02-26</p> <p>This image from the Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) instrument on NASA's Terra satellite covers an area of 55 by 40 kilometers (34 by 25 miles) over the southwest part of the Malaspina <span class="hlt">Glacier</span> and Icy Bay in Alaska. The composite of infrared and visible bands results in the snow and ice appearing light blue, dense vegetation is yellow-orange and green, and less vegetated, gravelly areas are in orange. According to Dr. Dennis Trabant (U.S. Geological Survey, Fairbanks, Alaska), the Malaspina <span class="hlt">Glacier</span> is thinning. Its terminal moraine protects it from contact with the open ocean; without the moraine, or if sea level rises sufficiently to reconnect the <span class="hlt">glacier</span> with the ocean, the <span class="hlt">glacier</span> would start calving and retreat significantly. ASTER data are being used to help monitor the size and movement of some 15,000 tidal and piedmont <span class="hlt">glaciers</span> in Alaska. Evidence derived from ASTER and many other satellite and ground-based measurements suggests that only a few dozen Alaskan <span class="hlt">glaciers</span> are advancing. The overwhelming majority of them are retreating. This ASTER image was acquired on June 8, 2001. With its 14 spectral bands from the visible to the thermal infrared wavelength region, and its high spatial resolution of 15 to 90 meters (about 50 to 300 feet), ASTER will image Earth for the next six years to map and monitor the changing surface of our planet. http://photojournal.jpl.nasa.gov/catalog/PIA03475</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20000053507','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20000053507"><span>Southern Alaska as an Example of the Long-Term Consequences of Mountain Building Under the Influence of <span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Meigs, Andrew; Sauber, Jeanne</p> <p>2000-01-01</p> <p>Southern Alaska is a continent-scale region of ongoing crustal deformation within the Pacific-North American plate boundary zone. <span class="hlt">Glaciers</span> and glacial erosion have dictated patterns of denudation in the orogen over the last approx. 5 My. The orogen comprises three discrete topographic domains from south to north, respectively: (1) the Chugach/St. Elias Range; (2) the Wrangell Mountains; and (3) the eastern Alaska Range. Although present deformation is distributed across the orogen, much of the shortening and uplift are concentrated in the Chugach/St. Elias Range. A systematic increase in topographic wavelength of the range from <span class="hlt">east</span> to west reflects <span class="hlt">east</span>-to-west increases in the width of a shallowly-dipping segment of the plate interface, separation of major upper plate structures, and a decrease in the obliquity of plate motion relative to the plate boundary. Mean elevation decays exponentially from approx. 2500 m to approx. 1100 m from <span class="hlt">east</span> to west, respectively. Topographic control on the present and past distribution of <span class="hlt">glaciers</span> is indicated by close correspondence along the range between mean elevation and the modern equilibrium line altitude of <span class="hlt">glaciers</span> (ELA) and differences in the modern ELA, mean annual precipitation and temperature across the range between the windward, southern and leeward, northern flanks. Net, range- scale erosion is the sum of: (1) primary bedrock erosion by <span class="hlt">glaciers</span> and (2) erosion in areas of the landscape that are ice-marginal and are deglaciated at glacial minima. Oscillations between glacial and interglacial climates controls ice height and distribution, which, in turn, modulates the locus and mode of erosion in the landscape. Mean topography and the mean position of the ELA are coupled because of the competition between rock uplift, which tends to raise the ELA, and enhanced orographic precipitation accompanying mountain building, which tends to lower the ELA. Mean topography is controlled both by the 60 deg latitude and maritime</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020677','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020677"><span>Water flow through temperate <span class="hlt">glaciers</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>Fountain, A.G.; Walder, J.S.</p> <p>1998-01-01</p> <p>Understanding water movement through a <span class="hlt">glacier</span> is fundamental to several critical issues in glaciology, including <span class="hlt">glacier</span> dynamics, <span class="hlt">glacier</span>-induced floods, and the prediction of runoff from <span class="hlt">glacierized</span> drainage basins. to this end we have synthesized a conceptual model os water movement through a temperate <span class="hlt">glacier</span> from the surface to the outlet stream. Processes that regulate the rate and distribution of water input at the <span class="hlt">glacier</span> surface and that regulate water movement from the surface to the bed play important but commonly neglected roles in <span class="hlt">glacier</span> hydrology. Where a <span class="hlt">glacier</span> is covered by a layer of porous, permeable firn (the accumulation zone), the flux of water to the <span class="hlt">glacier</span> interior varies slowly because the firn temporarily stores water and thereby smooths out variations in the supply rate. In the firn-free ablation zone, in contrast, the flux of water into the <span class="hlt">glacier</span> depends directly on the rate of surface melt or rainfall and therefore varies greatly in time. Water moves from the surface to the bed through an upward branching arborescent network consisting of both steeply inclined conduits, formed by the enlargement of intergranular veins, and gently inclined conduits, sprqwned by water flow along the bottoms of near-surface fractures (crevasses). Englacial drainage conduits deliver water to the <span class="hlt">glacier</span> bed at a linited number of points, probably a long distance downglacier of where water enters the <span class="hlt">glacier</span>. Englacial conduits supplied from the accumulation zone are quasi steady state features that convey the slowly varying water flux delivered via the firn. their size adjusts so that they are usually full of water and flow is pressurized. In contrast, water flow in englacial conduits supplied from the ablation area is pressurized only near times of peak daily flow or during rainstorms; flow is otherwise in an open-channel configuration. The subglacial drainage system typically consists of several elements that are distinct both morpphologically and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/dc0640.photos.036893p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/dc0640.photos.036893p/"><span>1. STONE BRIDGE, LOOKING <span class="hlt">EAST</span> DOWNSTREAM Photocopy of photograph, summer ...</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. STONE BRIDGE, LOOKING <span class="hlt">EAST</span> DOWNSTREAM Photocopy of photograph, summer 1932 National <span class="hlt">Park</span> Service, National Capital Region files - Dumbarton Oaks <span class="hlt">Park</span>, Thirty-second & R Streets Northwest, Washington, District of Columbia, DC</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C53C0742S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C53C0742S"><span>Glacimarine Sedimentary Processes and Deposits at Fjord-Terminating Tidewater <span class="hlt">Glacier</span> Margins</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Streuff, K.; O'Cofaigh, C.; Lloyd, J. M.; Noormets, R.; Nielsen, T.; Kuijpers, A.</p> <p>2016-12-01</p> <p>Many fjords along Arctic coasts are influenced by tidewater <span class="hlt">glaciers</span>, some of them fast-flowing ice sheet outlets. Such <span class="hlt">glaciers</span> provide important links between terrestrial and marine environments, and, due to their susceptibility to climatic and oceanographic changes, have undergone a complex history of advance and retreat since the last glacial maximum (LGM). Although a growing body of evidence has led to a better understanding of the deglacial dynamics of individual <span class="hlt">glaciers</span> since the LGM, their overall Holocene glacimarine processes and associated sedimentary and geomorphological products often remain poorly understood. This study addresses this through a detailed analysis of sediment cores, swath bathymetric and sub-bottom profiler data collected from seven fjords in Spitsbergen and west Greenland. The sediment cores preserve a complex set of lithofacies, which include laminated and massive muds in ice-proximal, and bioturbated mud in more ice-distal settings, diamicton in iceberg-dominated areas and massive sand occurring as lenses, laminae and thick beds. These facies record the interplay of three main glacimarine processes, suspension settling, iceberg rafting and sediment gravity flows, and collectively emphasise the dominance of glacial meltwater delivery to sedimentation in high Arctic fjords. The seafloor geomorphology in the fjords shows a range of landforms that include glacial lineations associated with fast ice-flow, terminal moraines and debris lobes marking former maximum <span class="hlt">glacier</span> extents, and small transverse moraines formed during deglaciation by glaciotectonic deformation at the grounding line and crevasse-squeezing. Additional landforms such as iceberg ploughmarks, submarine channels, pockmarks, and debris lobes formed during or after deglaciation by iceberg calving, erosion by meltwater, and sediment reworking. We present here a new model for sedimentary and geomorphological processes in front of contemporary tidewater <span class="hlt">glaciers</span>, which</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.C22B..05B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.C22B..05B"><span>New constraints on the structure and dynamics of the <span class="hlt">East</span> Antarctic Ice Sheet from the joint IPY/Ice Bridge ICECAP aerogeophysical project</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blankenship, D. D.; Young, D. A.; Siegert, M. J.; van Ommen, T. D.; Roberts, J. L.; Wright, A.; Warner, R. C.; Holt, J. W.; Young, N. W.; Le Meur, E.; Legresy, B.; Cavitte, M.; Icecap Team</p> <p>2010-12-01</p> <p>Ice within marine basins of <span class="hlt">East</span> Antarctica, and their outlets, represent the ultimate limit on sea level change. The region of <span class="hlt">East</span> Antarctica between the Ross Sea and Wilkes Land hosts a number of major basin, but has been poorly understood. Long range aerogeophysics from US, Australian and French stations, with significant British and IceBridge support, has, under the banner of the ICECAP project, greatly improved our knowledge of ice thickness, surface elevation, and crustal structure of the Wilkes and Aurora Subglacial Basins, as well as the Totten <span class="hlt">Glacier</span>, Cook Ice Shelf, and Byrd <span class="hlt">Glacier</span>. We will discuss the evolution of the Wilkes and Aurora Subglacial Basins, new constraints on the geometry of the major outlet <span class="hlt">glaciers</span>, as well as our results from surface elevation change measurements over dynamic regions of the ice sheet. We will discuss the implications of our data for the presence of mid Pleistocene ice in central <span class="hlt">East</span> Antarctica. Future directions for ICECAP will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3759454','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3759454"><span>The Significance of Shifts in Precipitation Patterns: Modelling the Impacts of Climate Change and <span class="hlt">Glacier</span> Retreat on Extreme Flood Events in Denali National <span class="hlt">Park</span>, Alaska</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Crossman, Jill; Futter, Martyn N.; Whitehead, Paul G.</p> <p>2013-01-01</p> <p>In <span class="hlt">glacier</span>-fed systems climate change may have various effects over a range of time scales, including increasing river discharge, flood frequency and magnitude. This study uses a combination of empirical monitoring and modelling to project the impacts of climate change on the glacial-fed Middle Fork Toklat River, Denali National <span class="hlt">Park</span>, Alaska. We use a regional calibration of the model HBV to account for a paucity of long term observed flow data, validating a local application using glacial mass balance data and summer flow records. Two Global Climate Models (HADCM3 and CGCM2) and two IPCC scenarios (A2 and B2) are used to ascertain potential changes in meteorological conditions, river discharge, flood frequency and flood magnitude. Using remote sensing methods this study refines existing estimates of glacial recession rates, finding that since 2000, rates have increased from 24m per year to 68.5m per year, with associated increases in ablation zone ice loss. GCM projections indicate that over the 21st century these rates will increase still further, most extensively under the CGCM2 model, and A2 scenarios. Due to greater winter precipitation and ice and snow accumulation, <span class="hlt">glaciers</span> release increasing meltwater quantities throughout the 21st century. Despite increases in glacial melt, results indicate that it is predominantly precipitation that affects river discharge. Three of the four IPCC scenarios project increases in flood frequency and magnitude, events which were primarily associated with changing precipitation patterns, rather than extreme temperature increases or meltwater release. Results suggest that although increasing temperatures will significantly increase glacial melt and winter baseflow, meltwater alone does not pose a significant flood hazard to the Toklat River catchment. Projected changes in precipitation are the primary concern, both through changing snow volumes available for melt, and more directly through increasing catchment runoff. PMID</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24023925','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24023925"><span>The significance of shifts in precipitation patterns: modelling the impacts of climate change and <span class="hlt">glacier</span> retreat on extreme flood events in Denali National <span class="hlt">Park</span>, Alaska.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Crossman, Jill; Futter, Martyn N; Whitehead, Paul G</p> <p>2013-01-01</p> <p>In <span class="hlt">glacier</span>-fed systems climate change may have various effects over a range of time scales, including increasing river discharge, flood frequency and magnitude. This study uses a combination of empirical monitoring and modelling to project the impacts of climate change on the glacial-fed Middle Fork Toklat River, Denali National <span class="hlt">Park</span>, Alaska. We use a regional calibration of the model HBV to account for a paucity of long term observed flow data, validating a local application using glacial mass balance data and summer flow records. Two Global Climate Models (HADCM3 and CGCM2) and two IPCC scenarios (A2 and B2) are used to ascertain potential changes in meteorological conditions, river discharge, flood frequency and flood magnitude. Using remote sensing methods this study refines existing estimates of glacial recession rates, finding that since 2000, rates have increased from 24 m per year to 68.5m per year, with associated increases in ablation zone ice loss. GCM projections indicate that over the 21(st) century these rates will increase still further, most extensively under the CGCM2 model, and A2 scenarios. Due to greater winter precipitation and ice and snow accumulation, <span class="hlt">glaciers</span> release increasing meltwater quantities throughout the 21(st) century. Despite increases in glacial melt, results indicate that it is predominantly precipitation that affects river discharge. Three of the four IPCC scenarios project increases in flood frequency and magnitude, events which were primarily associated with changing precipitation patterns, rather than extreme temperature increases or meltwater release. Results suggest that although increasing temperatures will significantly increase glacial melt and winter baseflow, meltwater alone does not pose a significant flood hazard to the Toklat River catchment. Projected changes in precipitation are the primary concern, both through changing snow volumes available for melt, and more directly through increasing catchment runoff.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e002161.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e002161.html"><span>Icefall, Lambert <span class="hlt">Glacier</span>, Antarctica</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Image taken 12/2/2000: The Lambert <span class="hlt">Glacier</span> in Antarctica, is the world's largest <span class="hlt">glacier</span>. The focal point of this image is an icefall that feeds into the Lambert <span class="hlt">glacier</span> from the vast ice sheet covering the polar plateau. Ice flows like water, albeit much more slowly. Cracks can be seen in this icefall as it bends and twists on its slow-motion descent 1300 feet (400 meters) to the <span class="hlt">glacier</span> below. This Icefall can be found on Landsat 7 WRS Path 42 Row 133/134/135, center: -70.92, 69.15. To learn more about the Landsat satellite go to: landsat.gsfc.nasa.gov/</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.B54B..05H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.B54B..05H"><span>Differences in dissolved organic matter lability between alpine <span class="hlt">glaciers</span> and alpine rock <span class="hlt">glaciers</span> of the American West</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hall, E.; Fegel, T. S., II; Baron, J.; Boot, C. M.</p> <p>2015-12-01</p> <p>While alpine <span class="hlt">glaciers</span> in montane regions represent the largest flux of dissolved organic matter (DOM) from global ice melt no research has examined the bioavailability of DOM melted out of glacial ice in the western continental United States. Furthermore, rock <span class="hlt">glaciers</span> are an order of magnitude more abundant than ice <span class="hlt">glaciers</span> in U.S., yet are not included in budgets for perennial ice carbon stores. Our research aims to understand differences in the bioavailability of carbon from ice <span class="hlt">glaciers</span> and rock <span class="hlt">glaciers</span> along the Central Rocky Mountains of Colorado. Identical microbial communities were fed standardized amounts of DOM from four different ice <span class="hlt">glacier</span>-rock <span class="hlt">glaciers</span> pairs. Using laboratory incubations, paired with mass spectrometry based metabolomics and 16S gene sequencing; we were able to examine functional definitions of DOM lability in glacial ice. We hypothesized that even though DOM quantities are similar in the outputs of both glacial types in our study area, ice glacial DOM would be more bioavailable than DOM from rock <span class="hlt">glaciers</span> due to higher proportions of byproducts from microbial metabolism than rock <span class="hlt">glacier</span> DOM, which has higher amounts of "recalcitrant" plant material. Our results show that DOM from ice <span class="hlt">glaciers</span> is more labile than DOM from geologically and geographically similar paired rock <span class="hlt">glaciers</span>. Ice <span class="hlt">glacier</span> DOM represents an important pool of labile carbon to headwater ecosystems of the Rocky Mountains. Metabolomic analysis shows numerous compounds from varying metabolite pathways, including byproducts of nitrification before and after incubation, meaning that, similar to large maritime <span class="hlt">glaciers</span> in Alaska and Europe, subglacial environments in the mountain ranges of the United States are hotspots for biological activity and processing of organic carbon.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA11419.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA11419.html"><span>Byrd <span class="hlt">Glacier</span>, Antarctica</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2008-11-17</p> <p>Byrd <span class="hlt">Glacier</span> is a major <span class="hlt">glacier</span> in Antarctica; it drains an extensive area of the polar plateau and flows eastward between the Britannia Range and the Churchill Mountains to discharge into the Ross Ice Shelf. This image is from NASA Terra satellite.</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://www.osti.gov/biblio/5175178-columbia-glacier-disintegration-underway','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5175178-columbia-glacier-disintegration-underway"><span>Columbia <span class="hlt">Glacier</span> in 1984: disintegration underway</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>Meier, M.F.; Rasmussen, L.A.; Miller, D.S.</p> <p>1985-01-01</p> <p>Columbia <span class="hlt">Glacier</span> is a large, iceberg-calving <span class="hlt">glacier</span> near Valdez, Alaska. The terminus of this <span class="hlt">glacier</span> was relatively stable from the time of the first scientific studies in 1899 until 1978. During this period the <span class="hlt">glacier</span> terminated partly on Heather Island and partly on a submerged moraine shoal. In December, 1978, the <span class="hlt">glacier</span> terminus retreated from Heather Island, and retreat has accelerated each year since then, except during a period of anomalously low calving in 1980. Although the <span class="hlt">glacier</span> has not terminated on Heather Island since 1978, a portion of the terminus remained on the crest of the moraine shoal untilmore » the fall of 1983. By December 8, 1983, that feature had receded more than 300 m from the crest of the shoal, and by December 14, 1984, had disappeared completely, leaving most of the terminus more than 2000 meters behind the crest of the shoal. Recession of the <span class="hlt">glacier</span> from the shoal has placed the terminus in deeper water, although the <span class="hlt">glacier</span> does not float. The active calving face of the <span class="hlt">glacier</span> now terminates in seawater that is about 300 meters deep at the <span class="hlt">glacier</span> centerline. Rapid calving appears to be associated with buoyancy effects due to deep water at the terminus and subglacial runoff. 12 refs., 10 figs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1714078M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1714078M"><span>Annual and seasonal mass balances of Chhota Shigri <span class="hlt">Glacier</span> (benchmark <span class="hlt">glacier</span>, Western Himalaya), India</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mandal, Arindan; Ramanathan, Alagappan; Farooq Azam, Mohd; Wagnon, Patrick; Vincent, Christian; Linda, Anurag; Sharma, Parmanand; Angchuk, Thupstan; Bahadur Singh, Virendra; Pottakkal, Jose George; Kumar, Naveen; Soheb, Mohd</p> <p>2015-04-01</p> <p>Several studies on Himalayan <span class="hlt">glaciers</span> have been recently initiated as they are of particular interest in terms of future water supply, regional climate change and sea-level rise. In 2002, a long-term monitoring program was initiated on Chhota Shigri <span class="hlt">Glacier</span> (15.7 square km, 9 km long, 6263-4050 m a.s.l.) located in Lahaul and Spiti Valley, Himachal Pradesh, India. This <span class="hlt">glacier</span> lies in the monsoon-arid transition zone (western Himalaya) and is a representative <span class="hlt">glacier</span> in Lahaul and Spiti Valley. While annual mass balances have been measured continuously since 2002 using the glaciological method, seasonal scale observations began in 2009. The annual and seasonal mass balances were then analyzed along with meteorological conditions in order to understand the role of winter and summer balances on annual <span class="hlt">glacier</span>-wide mass balance of Chhota Shigri <span class="hlt">glacier</span>. During the period 2002-2013, the <span class="hlt">glacier</span> experienced a negative <span class="hlt">glacier</span>-wide mass balance of -0.59±0.40 m w.e. a-1 with a cumulative glaciological mass balance of -6.45 m w.e. Annual <span class="hlt">glacier</span>-wide mass balances were negative except for four years (2004/05, 2008/09, 2009/10 and 2010/11) where it was generally close to balanced conditions. Equilibrium line altitude (ELA) for steady state condition is calculated as 4950 m a.s.l. corresponding to an accumulation area ratio (AAR) of 62% using annual <span class="hlt">glacier</span>-wide mass balance, ELA and AAR data between 2002 and 2013. The winter <span class="hlt">glacier</span>-wide mass balance between 2009 and 2013 ranges from a maximum value of 1.38 m w.e. in 2009/10 to a minimum value of 0.89 in 2012/13 year whereas the summer <span class="hlt">glacier</span>-wide mass balance varies from the highest value of -0.95 m w.e. in 2010/11 to the lowest value of -1.72 m w.e. in 2011/12 year. The mean vertical mass balance gradient between 2002 and 2013 was 0.66 m w.e. (100 m)-1 quite similar to Alps, Nepalese Himalayas etc. Over debris covered area, the gradients are highly variable with a negative mean value of -2.15 m w.e. (100 m)-1 over 2002</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.C11E..01R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.C11E..01R"><span>GLIMS <span class="hlt">Glacier</span> Database: Status and Challenges</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raup, B. H.; Racoviteanu, A.; Khalsa, S. S.; Armstrong, R.</p> <p>2008-12-01</p> <p>GLIMS (Global Land Ice Measurements from Space) is an international initiative to map the world's <span class="hlt">glaciers</span> and to build a GIS database that is usable via the World Wide Web. The GLIMS programme includes 70 institutions, and 25 Regional Centers (RCs), who analyze satellite imagery to map <span class="hlt">glaciers</span> in their regions of expertise. The analysis results are collected at the National Snow and Ice Data Center (NSIDC) and ingested into the GLIMS <span class="hlt">Glacier</span> Database. The database contains approximately 80 000 <span class="hlt">glacier</span> outlines, half the estimated total on Earth. In addition, the database contains metadata on approximately 200 000 ASTER images acquired over <span class="hlt">glacierized</span> terrain. <span class="hlt">Glacier</span> data and the ASTER metadata can be viewed and searched via interactive maps at http://glims.org/. As <span class="hlt">glacier</span> mapping with GLIMS has progressed, various hurdles have arisen that have required solutions. For example, the GLIMS community has formulated definitions for how to delineate <span class="hlt">glaciers</span> with different complicated morphologies and how to deal with debris cover. Experiments have been carried out to assess the consistency of the database, and protocols have been defined for the RCs to follow in their mapping. Hurdles still remain. In June 2008, a workshop was convened in Boulder, Colorado to address issues such as mapping debris-covered <span class="hlt">glaciers</span>, mapping ice divides, and performing change analysis using two different <span class="hlt">glacier</span> inventories. This contribution summarizes the status of the GLIMS <span class="hlt">Glacier</span> Database and steps taken to ensure high data quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/dc0640.photos.036914p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/dc0640.photos.036914p/"><span>22. MEADOW, LOOKING <span class="hlt">EAST</span> WITH STREAM ARBOR ON RIGHT Photocopy ...</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. MEADOW, LOOKING <span class="hlt">EAST</span> WITH STREAM ARBOR ON RIGHT Photocopy of photograph, 1930s National <span class="hlt">Park</span> Service, National Capital Region files - Dumbarton Oaks <span class="hlt">Park</span>, Thirty-second & R Streets Northwest, Washington, District of Columbia, DC</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1910656I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1910656I"><span>How well do we really know the timing and extent of <span class="hlt">glaciers</span> during the Last Glacial Maximum in the Alps?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ivy-Ochs, Susan; Braakhekke, Jochem; Monegato, Giovanni; Gianotti, Franco; Forno, Gabriella; Hippe, Kristina; Christl, Marcus; Akçar, Naki; Schluechter, Christian</p> <p>2017-04-01</p> <p>The Last Glacial Maximum (LGM) in the Alps saw much of the mountains inundated by ice. Several main accumulation areas comprising local ice caps and plateau icefields fit into a picture of transection <span class="hlt">glaciers</span> flowing into huge valley <span class="hlt">glaciers</span>. In the north the valley <span class="hlt">glaciers</span> covered long distances (hundreds of kilometers) to reach the forelands where they spread out in fan-shaped piedmont lobes tens of kilometers across, e.g. the Rhine <span class="hlt">glacier</span>. In the south travel distances to the mountain front were often shorter, the pathway steeper. Nevertheless, not all <span class="hlt">glaciers</span> even reached beyond the front, as the temperatures were notably warmer in the south. For example at Orta the <span class="hlt">glacier</span> snout remained within the mountains. Where <span class="hlt">glaciers</span> reached the forelands they stopped abruptly and the moraine amphitheaters were constructed, e.g. at Ivrea and Rivoli-Avigliana. Sets of stacked moraines built-up as <span class="hlt">glacier</span> advance was directly confined by the older moraines. We may temporally and spatially identify the culmination of the last glacial cycle by pinpointing the outermost moraines that date to the LGM (generally about 26-24 ka). On the other hand, the timing of abandonment of foreland positions is given by ages of the innermost, often lake-bounding, moraines (about 19-18 ka). Between the two, <span class="hlt">glacier</span> fluctuations left the stadial moraines. In the Linth-Rhine system three stadials have been recognized: Killwangen, Schlieren and Zurich. Nevertheless, already in the Swiss sector correlation of the LGM stadials among the several foreland lobes is not unambiguous. Across the Alps, not only north to south but also west to <span class="hlt">east</span>, how do the timing and extent of <span class="hlt">glaciers</span> during the LGM vary? Recent <span class="hlt">glacier</span> modelling by Seguinot et al. (2017) informs and suggests the possibility of differences in timing for reaching of the maximum extent and for the number of oscillations of individual lobes during the LGM. At present few sites in the Alps have detailed enough geomorphological</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28733603','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28733603"><span>Sediment transport drives tidewater <span class="hlt">glacier</span> periodicity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brinkerhoff, Douglas; Truffer, Martin; Aschwanden, Andy</p> <p>2017-07-21</p> <p>Most of Earth's <span class="hlt">glaciers</span> are retreating, but some tidewater <span class="hlt">glaciers</span> are advancing despite increasing temperatures and contrary to their neighbors. This can be explained by the coupling of ice and sediment dynamics: a shoal forms at the <span class="hlt">glacier</span> terminus, reducing ice discharge and causing advance towards an unstable configuration followed by abrupt retreat, in a process known as the tidewater <span class="hlt">glacier</span> cycle. Here we use a numerical model calibrated with observations to show that interactions between ice flow, glacial erosion, and sediment transport drive these cycles, which occur independent of climate variations. Water availability controls cycle period and amplitude, and enhanced melt from future warming could trigger advance even in <span class="hlt">glaciers</span> that are steady or retreating, complicating interpretations of <span class="hlt">glacier</span> response to climate change. The resulting shifts in sediment and meltwater delivery from changes in <span class="hlt">glacier</span> configuration may impact interpretations of marine sediments, fjord geochemistry, and marine ecosystems.The reason some of the Earth's tidewater <span class="hlt">glaciers</span> are advancing despite increasing temperatures is not entirely clear. Here, using a numerical model that simulates both ice and sediment dynamics, the authors show that internal dynamics drive <span class="hlt">glacier</span> variability independent of climate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..14.1644O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..14.1644O"><span>Attribution of <span class="hlt">glacier</span> fluctuations to climate change</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Oerlemans, J.</p> <p>2012-04-01</p> <p><span class="hlt">Glacier</span> retreat is a worlwide phenomenon, which started around the middle of the 19th century. During the period 1800-1850 the number of retreating and advancing <span class="hlt">glaciers</span> was roughly equal (based on 42 records from different continents). During the period 1850-1900 about 92% of all mountain <span class="hlt">glaciers</span> became shorter (based on 65 records). After this, the percentage of shrinking <span class="hlt">glaciers</span> has been around 90% until the present time. The <span class="hlt">glacier</span> signal is rather coherent over the globe, especially when surging and calving <span class="hlt">glaciers</span> are not considered (for such <span class="hlt">glaciers</span> the response to climate change is often masked by length changes related to internal dynamics). From theoretical studies as well as extensive meteorological work on <span class="hlt">glaciers</span>, the processes that control the response of <span class="hlt">glaciers</span> to climate change are now basically understood. It is useful to make a difference between geometric factors (e.g. slope, altitudinal range, hypsometry) and climatic setting (e.g. seasonal cycle, precipitation). The most sensitive <span class="hlt">glaciers</span> appear to be flat <span class="hlt">glaciers</span> in a maritime climate. Characterizing the dynamic properties of a <span class="hlt">glacier</span> requires at least two quantities: the climate sensitivity, expressing how the equilibrium <span class="hlt">glacier</span> state depends on the climatic conditions, and the response time, indicating how fast a <span class="hlt">glacier</span> approaches a new equilibrium state after a stepwise change in the climatic forcing. These quantities can be estimated from relatively simple theory, showing that differences among <span class="hlt">glaciers</span> are substantial. For larger <span class="hlt">glaciers</span>, climate sensitivities (in terms of <span class="hlt">glacier</span> length) vary from 1 to 8 km per 100 m change in the equilibrium-line altitude. Response times are mainly in the range of 20 to 200 years, with most values between 30 and 80 years. Changes in the equilibrium-line altitude or net mass balance of a <span class="hlt">glacier</span> are mainly driven by fluctuations in air temperature, precipitation, and global radiation. Energy-balance modelling for many <span class="hlt">glaciers</span> shows that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170008844','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170008844"><span>Optimizing <span class="hlt">Parking</span> Orbits for Roundtrip Mars Missions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Qu, Min; Merill, Raymond G.; Chai, Patrick; Komar, David R.</p> <p>2017-01-01</p> <p>A roundtrip Mars mission presents many challenges to the design of a transportation system and requires a series of orbital maneuvers within Mars vicinity to capture, reorient, and then return the spacecraft back to Earth. The selection of a Mars <span class="hlt">parking</span> orbit is crucial to the mission design; not only can the <span class="hlt">parking</span> or-bit choice drastically impact the ?V requirements of these maneuvers but also it must be properly aligned to target desired surface or orbital destinations. This paper presents a method that can optimize the Mars <span class="hlt">parking</span> orbits given the arrival and departure conditions from heliocentric trajectories, and it can also en-force constraints on the <span class="hlt">parking</span> orbits to satisfy other architecture design requirements such as co-planar subperiapsis descent to planned landing sites, due <span class="hlt">east</span> or co-planar ascent back to the <span class="hlt">parking</span> orbit, or low cost transfers to and from Phobos and Deimos.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1910602R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1910602R"><span>Recent Advances in the GLIMS <span class="hlt">Glacier</span> Database</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raup, Bruce; Cogley, Graham; Zemp, Michael; Glaus, Ladina</p> <p>2017-04-01</p> <p><span class="hlt">Glaciers</span> are shrinking almost without exception. <span class="hlt">Glacier</span> losses have impacts on local water availability and hazards, and contribute to sea level rise. To understand these impacts and the processes behind them, it is crucial to monitor <span class="hlt">glaciers</span> through time by mapping their areal extent, changes in volume, elevation distribution, snow lines, ice flow velocities, and changes to associated water bodies. The <span class="hlt">glacier</span> database of the Global Land Ice Measurements from Space (GLIMS) initiative is the only multi-temporal <span class="hlt">glacier</span> database capable of tracking all these <span class="hlt">glacier</span> measurements and providing them to the scientific community and broader public. Here we present recent results in 1) expansion of the geographic and temporal coverage of the GLIMS <span class="hlt">Glacier</span> Database by drawing on the Randolph <span class="hlt">Glacier</span> Inventory (RGI) and other new data sets; 2) improved tools for visualizing and downloading GLIMS data in a choice of formats and data models; and 3) a new data model for handling multiple <span class="hlt">glacier</span> records through time while avoiding double-counting of <span class="hlt">glacier</span> number or area. The result of this work is a more complete <span class="hlt">glacier</span> data repository that shows not only the current state of <span class="hlt">glaciers</span> on Earth, but how they have changed in recent decades. The database is useful for tracking changes in water resources, hazards, and mass budgets of the world's <span class="hlt">glaciers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001909.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001909.html"><span>Gyldenlove <span class="hlt">Glacier</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-04-11</p> <p>On April 11, 2011, IceBridge finally got the clear weather necessary to fly over <span class="hlt">glaciers</span> in southeast Greenland. But with clear skies came winds of up to 70 knots, which made for a bumpy ride over the calving front of <span class="hlt">glaciers</span> like Gyldenlove. Operation IceBridge, now in its third year, makes annual campaigns in the Arctic and Antarctic where science flights monitor <span class="hlt">glaciers</span>, ice sheets and sea ice. Credit: NASA/GSFC/Michael Studinger To learn more about Ice Bridge go to: www.nasa.gov/mission_pages/icebridge/news/spr11/index.html 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 Join us on Facebook</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca2171.photos.383343p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca2171.photos.383343p/"><span>7. ORIGINAL SOUTH SIDE, <span class="hlt">EAST</span> PART, ALSO SHOWING PATH 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>7. ORIGINAL SOUTH SIDE, <span class="hlt">EAST</span> PART, ALSO SHOWING PATH FROM <span class="hlt">EAST</span> FRONT LEADING TO CENTRAL CAMPUS. - U.S. Geological Survey, Rock Magnetics Laboratory, 345 Middlefield Road, Menlo <span class="hlt">Park</span>, San Mateo County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70073496','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70073496"><span>Application of snow models to snow removal operations on the Going-to-the-Sun Road, <span class="hlt">Glacier</span> National <span class="hlt">Park</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>Fagre, Daniel B.; Klasner, Frederick L.</p> <p>2000-01-01</p> <p>Snow removal, and the attendant avalanche risk for road crews, is a major issue on mountain highways worldwide. The Going-to-the-Sun Road is the only road that crosses <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana. This 80-km highway ascends over 1200m along the wall of a glaciated basin and crosses the continental divide. The annual opening of the road is critical to the regional economy and there is public pressure to open the road as early as possible. Despite the 67-year history of snow removal activities, few stat on snow conditions at upper elevations were available to guide annual planning for the raod opening. We examined statistical relationships between the opening date and nearby SNOTEL data on snow water equivalence (WE) for 30 years. Early spring SWE (first Monday in April) accounted for only 33% of the variance in road opening dates. Because avalanche spotters, used to warn heavy equipment operators of danger, are ineffective during spring storms or low-visibility conditions, we incorporated the percentage of days with precipitation during plowing as a proxy for visibility. This improved the model's predictive power to 69%/ A mountain snow simulator (MTSNOW) was used to calculate the depth and density of snow at various points along the road and field data were collected for comparison. MTSNOW underestimated the observed snow conditions, in part because it does not yet account for wind redistribution of snow. The severe topography of the upper reaches of the road are subjected to extensive wind redistribution of snow as evidence by the formation of "The Big Drift" on the lee side of Logan Pass.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015TCry....9.2215B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015TCry....9.2215B"><span>Brief communication: Getting Greenland's <span class="hlt">glaciers</span> right - a new data set of all official Greenlandic <span class="hlt">glacier</span> names</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bjørk, A. A.; Kruse, L. M.; Michaelsen, P. B.</p> <p>2015-12-01</p> <p>Place names in Greenland can be difficult to get right, as they are a mix of Greenlandic, Danish, and other foreign languages. In addition, orthographies have changed over time. With this new data set, we give the researcher working with Greenlandic <span class="hlt">glaciers</span> the proper tool to find the correct name for <span class="hlt">glaciers</span> and ice caps in Greenland and to locate <span class="hlt">glaciers</span> described in the historic literature with the old Greenlandic orthography. The data set contains information on the names of 733 <span class="hlt">glaciers</span>, 285 originating from the Greenland Ice Sheet (GrIS) and 448 from local <span class="hlt">glaciers</span> and ice caps (LGICs).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70179841','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70179841"><span>Chemistry of selected high-elevation lakes in seven national <span class="hlt">parks</span> in the western United States</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Clow, David W.; Striegl, Robert G.; Nanus, Leora; Mast, M. Alisa; Campbell, Donald H.; Krabbenhoft, David P.</p> <p>2002-01-01</p> <p>A chemical survey of 69 high-altitude lakes in seven national <span class="hlt">parks</span> in the western United States was conducted during the fallof 1999; the lakes were previously sampled during the fall of 1985, as part of the Western Lake Survey. Lakes in <span class="hlt">parks</span> in the Sierra/southern Cascades (Lassen Volcanic, Yosemite, Sequoia/Kings Canyon National <span class="hlt">Parks</span>) and in the southern RockyMountains (Rocky Mountain National <span class="hlt">Park</span>) were very dilute; medianspecific conductance ranged from 4.4 to 12.2 μS cm-1 andmedian alkalinity concentrations ranged from 32.2 to 72.9 μeqL-1. Specific conductances and alkalinity concentrations were substantially higher in lakes in the central and northernRocky Mountains <span class="hlt">parks</span> (Grand Teton, Yellowstone, and <span class="hlt">Glacier</span>National <span class="hlt">Parks</span>), probably due to the prevalence of more reactivebedrock types. Regional patterns in lake concentrations of NO3 and SO4 were similar to regional patterns in NO3 and SO4 concentrations in precipitation, suggestingthat the lakes are showing a response to atmospheric deposition.Concentrations of NO3 were particularly high in Rocky Mountain National <span class="hlt">Park</span>, where some ecosystems appear to be undergoing nitrogen saturation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMNH51A1228F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMNH51A1228F"><span>Fuzzy Cognitive Maps for <span class="hlt">Glacier</span> Hazards Assessment: Application to Predicting the Potential for <span class="hlt">Glacier</span> Lake Outbursts</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Furfaro, R.; Kargel, J. S.; Fink, W.; Bishop, M. P.</p> <p>2010-12-01</p> <p><span class="hlt">Glaciers</span> and ice sheets are among the largest unstable parts of the solid Earth. Generally, <span class="hlt">glaciers</span> are devoid of resources (other than water), are dangerous, are unstable and no infrastructure is normally built directly on their surfaces. Areas down valley from large alpine <span class="hlt">glaciers</span> are also commonly unstable due to landslide potential of moraines, debris flows, snow avalanches, outburst floods from <span class="hlt">glacier</span> lakes, and other dynamical alpine processes; yet there exists much development and human occupation of some disaster-prone areas. Satellite remote sensing can be extremely effective in providing cost-effective and time- critical information. Space-based imagery can be used to monitor <span class="hlt">glacier</span> outlines and their lakes, including processes such as iceberg calving and debris accumulation, as well as changing thicknesses and flow speeds. Such images can also be used to make preliminary identifications of specific hazardous spots and allows preliminary assessment of possible modes of future disaster occurrence. Autonomous assessment of <span class="hlt">glacier</span> conditions and their potential for hazards would present a major advance and permit systematized analysis of more data than humans can assess. This technical leap will require the design and implementation of Artificial Intelligence (AI) algorithms specifically designed to mimic <span class="hlt">glacier</span> experts’ reasoning. Here, we introduce the theory of Fuzzy Cognitive Maps (FCM) as an AI tool for predicting and assessing natural hazards in alpine <span class="hlt">glacier</span> environments. FCM techniques are employed to represent expert knowledge of <span class="hlt">glaciers</span> physical processes. A cognitive model embedded in a fuzzy logic framework is constructed via the synergistic interaction between glaciologists and AI experts. To verify the effectiveness of the proposed AI methodology as applied to predicting hazards in <span class="hlt">glacier</span> environments, we designed and implemented a FCM that addresses the challenging problem of autonomously assessing the <span class="hlt">Glacier</span> Lake Outburst Flow</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ISPAr42.3.1285M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ISPAr42.3.1285M"><span>Measuring Surface Deformation in <span class="hlt">Glacier</span> Retreated Areas Based on Ps-Insar - Geladandong <span class="hlt">Glacier</span> as a Case Study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mohamadi, B.; Balz, T.</p> <p>2018-04-01</p> <p><span class="hlt">Glaciers</span> are retreating in many parts of the world as a result of global warming. Many researchers consider Qinghai-Tibetan Plateau as a reference for climate change by measuring <span class="hlt">glaciers</span> retreat on the plateau. This retreat resulted in some topographic changes in retreated areas, and in some cases can lead to geohazards as landslides, and rock avalanches, which is known in <span class="hlt">glacier</span> retreated areas as paraglacial slope failure (PSF). In this study, Geladandong biggest and main <span class="hlt">glacier</span> mass was selected to estimate surface deformation on its <span class="hlt">glacier</span> retreated areas and define potential future PSF based on PS-InSAR technique. 56 ascending and 49 descending images were used to fulfill this aim. Geladandong <span class="hlt">glacier</span> retreated areas were defined based on the maximum extent of the <span class="hlt">glacier</span> in the little ice age. Results revealed a general uplift in the <span class="hlt">glacier</span> retreated areas with velocity less than 5mm/year. Obvious surface motion was revealed in seven parts surround <span class="hlt">glacier</span> retreated areas with high relative velocity reached ±60mm/year in some parts. Four parts were considered as PSF potential motion, and two of them showed potential damage for the main road in the study area in case of rock avalanche into recent <span class="hlt">glacier</span> lakes that could result in <span class="hlt">glacier</span> lake outburst flooding heading directly to the road. Finally, further analysis and field investigations are needed to define the main reasons for different types of deformation and estimate future risks of these types of surface motion in the Qinghai-Tibetan Plateau.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1615188B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1615188B"><span>"Dynamic Geodiversity" of glacial environments: new techniques for monitoring landscape variations on Alpine areas. Examples from the Gran Paradiso National <span class="hlt">Park</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bertotto, Stefania; Giardino, Marco; Perotti, Luigi; Mortara, Giovanni; Baroni, Carlo</p> <p>2014-05-01</p> <p>The importance of keeping memory of past morpho-climatic events is particular evident in recently deglaciated areas. The survival of <span class="hlt">glaciers</span> is now very uncertain, due to climate changes and related effects occurring in the last decades. In the Western Alps, many <span class="hlt">glaciers</span> are now extinct or show a dramatic reduction of area and thickness. Permafrost and periglacial areas are also responding promptly to climate changes as <span class="hlt">glaciers</span> do, but they are not good "visual" indicators of climate changes, because they are not easily recognizable. Indeed, Italian glacial elements are constantly monitored by the Italian Glaciological Committe (CGI) in the last two centuries. The volunteers of CGI constantly monitor variations of <span class="hlt">glacier</span> snout position of a great majority of Italian <span class="hlt">glaciers</span>. CGI is not only a very important source of historical documentation and information, but also a very important scientific reference of the studies conducted in glacial areas. Particularly, thanks to CGI, it was created an inventory of Italian <span class="hlt">glaciers</span> was created. Anyway, due to recent rapid changes, it is difficult to quickly update the inventory, also considering the difficulty of reaching alpine high mountain areas. The recent use of Geomatics in geological and geomorphological studies can be applied to evaluate landform changes in glacial and periglacial areas. The combination of remote sensing and on field techniques (i.e. aerial photogrammetry, GPS, Terrestrial photogrammetry, satellite images and LiDaR) provides constant monitoring of landform changes and updating inventories. The Gran Paradiso National <span class="hlt">Park</span> (Piemonte and Valle d'Aosta Regions, Western Italian Alps) represents an excellent example of conservation of geodiversity. Many key-elements of the high mountain landscape are present here: <span class="hlt">glaciers</span>, glacial cirques, rock <span class="hlt">glaciers</span>, moraines (not only from Holocene, but also from Little Ice Age, of XVI-XIX centuries), steepled peaks, rock walls, roche moutonnée, ravines, debris</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Geomo.293..272C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Geomo.293..272C"><span>Glacial conditioning of stream position and flooding in the braid plain of the Exit <span class="hlt">Glacier</span> foreland, Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Curran, Janet H.; Loso, Michael G.; Williams, Haley B.</p> <p>2017-09-01</p> <p>Flow spilling out of an active braid plain often signals the onset of channel migration or avulsion to previously occupied areas. In a recently deglaciated environment, distinguishing between shifts in active braid plain location, considered reversible by fluvial processes at short timescales, and more permanent <span class="hlt">glacier</span>-conditioned changes in stream position can be critical to understanding flood hazards. Between 2009 and 2014, increased spilling from the Exit Creek braid plain in Kenai Fjords National <span class="hlt">Park</span>, Alaska, repeatedly overtopped the only access road to the popular Exit <span class="hlt">Glacier</span> visitor facilities and trails. To understand the likely cause of road flooding, we consider recent processes and the interplay between <span class="hlt">glacier</span> and fluvial system dynamics since the maximum advance of the Little Ice Age, around 1815. Patterns of temperature and precipitation, the variables that drive high streamflow via snowmelt, <span class="hlt">glacier</span> meltwater runoff, and rainfall, could not fully explain the timing of road floods. Comparison of high-resolution topographic data between 2008 and 2012 showed a strong pattern of braid plain aggradation along 3 km of <span class="hlt">glacier</span> foreland, not unexpected at the base of mountainous <span class="hlt">glaciers</span> and likely an impetus for channel migration. Historically, a dynamic zone follows the retreating <span class="hlt">glacier</span> in which channel positions shift rapidly in response to changes in the <span class="hlt">glacier</span> margin and fresh morainal deposits. This period of paraglacial adjustment lasts one to several decades at Exit <span class="hlt">Glacier</span>. Subsequently, as moraine breaches consolidate and lock the channel into position, and as the stream regains the lower-elevation valley center, upper-elevation surfaces are abandoned as terraces inaccessible by fluvial processes for timescales of decades to centuries. Where not constrained by these terraces and moraines, the channel is free to migrate, which in this aggradational setting generates an alluvial fan at the breach of the final prominent moraine. The position of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70170558','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70170558"><span>Magnetotelluric investigation of the Vestfold Hills and Rauer Group, <span class="hlt">East</span> Antarctica</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Peacock, Jared R.; Selway, Katherine</p> <p>2016-01-01</p> <p>The Vestfold Hills and Rauer Group in <span class="hlt">East</span> Antarctica have contrasting Archean to Neoproterozoic geological histories and are believed to be juxtaposed along a suture zone that now lies beneath the Sørsdal <span class="hlt">Glacier</span>. Exact location and age of this suture zone are unknown, as is its relationship to regional deformation associated with the amalgamation of <span class="hlt">East</span> Gondwana. To image the suture zone, magnetotelluric (MT) data were collected in Prydz Bay, <span class="hlt">East</span> Antarctica, mainly along a profile crossing the Sørsdal <span class="hlt">Glacier</span> and regions inland of the Vestfold Hills and Rauer Group islands. Time-frequency analysis of the MT time series yielded three important observations: (1) Wind speeds in excess of ∼8 m/s reduce coherence between electric and magnetic fields due to charged wind-blown particles of ice and snow. (2) Estimation of the MT transfer function is best between 1000 and 1400 UT when ionospheric Hall currents enhance the magnetic source field. (3) Nonplanar source field effects were minimal but detectable and removed from estimation of the MT transfer function. Inversions of MT data in 2-D and 3-D produce similar resistivity models, where structures in the preferred 3-D resistivity model correlate strongly with regional magnetic data. The electrically conductive Rauer Group is separated from the less conductive Vestfold Hills by a resistive zone under the Sørsdal <span class="hlt">Glacier</span>, which is interpreted to be caused by oxidation during suturing. Though a suture zone has been imaged, no time constrains on suturing can be made from the MT data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020502','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020502"><span>Isotopic composition of ice cores and meltwater from upper fremont <span class="hlt">glacier</span> and Galena Creek rock <span class="hlt">glacier</span>, Wyoming</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>DeWayne, Cecil L.; Green, J.R.; Vogt, S.; Michel, R.; Cottrell, G.</p> <p>1998-01-01</p> <p>Meltwater runoff from <span class="hlt">glaciers</span> can result from various sources, including recent precipitation and melted glacial ice. Determining the origin of the meltwater from <span class="hlt">glaciers</span> through isotopic analysis can provide information about such things as the character and distribution of ablation on <span class="hlt">glaciers</span>. A 9.4 m ice core and meltwater were collected in 1995 and 1996 at the glacigenic Galena Creek rock <span class="hlt">glacier</span> in Wyoming's Absaroka Mountains. Measurements of chlorine-36 (36Cl), tritium (3H), sulphur-35 (35S), and delta oxygen-18 (??18O) were compared to similar measurements from an ice core taken from the Upper Fremont <span class="hlt">Glacier</span> in the Wind River Range of Wyoming collected in 1991-95. Meltwater samples from three sites on the rock <span class="hlt">glacier</span> yielded 36Cl concentrations that ranged from 2.1 ?? 1.0 X 106 to 5.8??0.3 X 106 atoms/l. The ice-core 36Cl concentrations from Galena Creek ranged from 3.4??0.3 X 105 to 1.0??0.1 X 106 atoms/l. Analysis of an ice core from the Upper Fremont <span class="hlt">Glacier</span> yielded 36Cl concentrations of 1.2??0.2 X 106 and 5.2??0.2 X 106 atoms/l for pre- 1940 ice and between 2 X 106 and 3 X 106 atoms/l for post-1980 ice. Purdue's PRIME Lab analyzed the ice from the Upper Fremont <span class="hlt">Glacier</span>. The highest concentration of 36Cl in the ice was 77 ?? 2 X 106 atoms/l and was deposited during the peak of atmospheric nuclear weapons testing in the late 1950s. This is an order of magnitude greater than the largest measured concentration from both the Upper Fremont <span class="hlt">Glacier</span> ice core that was not affected by weapons testing fallout and the ice core collected from the Galena Creek rock <span class="hlt">glacier</span>. Tritium concentrations from the rock <span class="hlt">glacier</span> ranged from 9.2??0.6 to 13.2??0.8 tritium units (TU) in the meltwater to -1.3??1.3 TU in the ice core. Concentrations of 3H in the Upper Fremont <span class="hlt">Glacier</span> ice core ranged from 0 TU in the ice older than 50 years to 6-12 TU in the ice deposited in the last 10 years. The maximum 3H concentration in ice from the Upper Fremont <span class="hlt">Glacier</span> deposited in the</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/2016AGUFM.C41C0685M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C41C0685M"><span>A Worldwide <span class="hlt">Glacier</span> Information System to go</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mölg, N.; Steinmann, M.; Zemp, M.</p> <p>2016-12-01</p> <p>In the forefront of the Paris Climate Conference COP21 in December 2015, the WGMS and UNESCO jointly launched a <span class="hlt">glacier</span> application for mobile devices. This new information system aims at bringing scientifically sound facts and figures on worldwide <span class="hlt">glacier</span> changes to decision makers at governmental and intergovernmental levels as well as reaching out to the interested public. The wgms <span class="hlt">Glacier</span> App provides a map interface based on satellite images that display all the observed <span class="hlt">glaciers</span> in the user's proximity. Basic information is provided for each <span class="hlt">glacier</span>, including photographs and general information on size and elevation. Graphs with observation data illustrate the <span class="hlt">glacier</span>'s development, along with information on latest principal investigators and their sponsoring agencies as well as detailed explanations of the measurement types. A text search allows the user to filter the <span class="hlt">glacier</span> by name, country, region, measurement type and the current "health" status, i.e. if the <span class="hlt">glacier</span> has gained or lost ice over the past decade. A compass shows the closest observed <span class="hlt">glaciers</span> in all directions from the user's current position. Finally, the card game allows the user to compete against the computer on the best monitored <span class="hlt">glaciers</span> in the world. Our poster provides a visual entrance point to the wgms <span class="hlt">Glacier</span> App and, hence, provides access to fluctuation series of more than 3'700 <span class="hlt">glaciers</span> around the world.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.8348W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.8348W"><span>Reduced melt on debris-covered <span class="hlt">glaciers</span>: investigations from Changri Nup <span class="hlt">Glacier</span>, Nepal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wagnon, Patrick; Vincent, Christian; Shea, Joseph M.; Immerzeel, Walter W.; Kraaijenbrink, Philip; Shrestha, Dibas; Soruco, Alvaro; Arnaud, Yves; Brun, Fanny; Berthier, Etienne; Futi Sherpa, Sonam</p> <p>2017-04-01</p> <p>Approximately 25% of the <span class="hlt">glacierized</span> area in the Everest region is covered by debris, yet the surface mass balance of debris-covered portions of these <span class="hlt">glaciers</span> has not been measured directly. In this study, ground-based measurements of surface elevation and ice depth are combined with terrestrial photogrammetry, unmanned aerial vehicle (UAV) and satellite elevation models to derive the surface mass balance of the debris-covered tongue of Changri Nup <span class="hlt">Glacier</span>, located in the Everest region. Over the debris-covered tongue, the mean elevation change between 2011 and 2015 is -0.93 m year-1 or -0.84 m water equivalent per year (w.e. a-1). The mean emergence velocity over this region, estimated from the total ice flux through a cross section immediately above the debris-covered zone, is +0.37mw.e. a-1. The debris-covered portion of the <span class="hlt">glacier</span> thus has an area averaged mass balance of -1.21+/-0.2mw.e. a-1 between 5240 and 5525 m above sea level (m a.s.l.). Surface mass balances observed on nearby debris-free <span class="hlt">glaciers</span> suggest that the ablation is strongly reduced (by ca. 1.8mw.e. a-1) by the debris cover. The insulating effect of the debris cover has a larger effect on total mass loss than the enhanced ice ablation due to supraglacial ponds and exposed ice cliffs. This finding contradicts earlier geodetic studies and should be considered for modelling the future evolution of debris-covered <span class="hlt">glaciers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhDT.......251T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhDT.......251T"><span>Flow instabilities of Alaskan <span class="hlt">glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Turrin, James Bradley</p> <p></p> <p>Over 300 of the largest <span class="hlt">glaciers</span> in southern Alaska have been identified as either surge-type or pulse-type, making <span class="hlt">glaciers</span> with flow instabilities the norm among large <span class="hlt">glaciers</span> in that region. Consequently, the bulk of mass loss due to climate change will come from these unstable <span class="hlt">glaciers</span> in the future, yet their response to future climate warming is unknown because their dynamics are still poorly understood. To help broaden our understanding of unstable <span class="hlt">glacier</span> flow, the decadal-scale ice dynamics of 1 surging and 9 pulsing <span class="hlt">glaciers</span> are investigated. Bering <span class="hlt">Glacier</span> had a kinematic wave moving down its ablation zone at 4.4 +/- 2.0 km/yr from 2002 to 2009, which then accelerated to 13.9 +/- 2.0 km/yr as it traversed the piedmont lobe. The wave first appeared in 2001 near the confluence with Bagley Ice Valley and it took 10 years to travel ~64 km. A surge was triggered in 2008 after the wave activated an ice reservoir in the midablation zone, and it climaxed in 2011 while the terminus advanced several km into Vitus Lake. Ruth <span class="hlt">Glacier</span> pulsed five times between 1973 and 2012, with peak velocities in 1981, 1989, 1997, 2003, and 2010; approximately every 7 years. A typical pulse increased ice velocity 300%, from roughly 40 m/yr to 160 m/yr in the midablation zone, and involved acceleration and deceleration of the ice en masse; no kinematic wave was evident. The pulses are theorized to be due to deformation of a subglacial till causing enhanced basal motion. Eight additional pulsing <span class="hlt">glaciers</span> are identified based on the spatiotemporal pattern of their velocity fields. These <span class="hlt">glaciers</span> pulsed where they were either constricted laterally or joined by a tributary, and their surface slopes are 1-2°. These traits are consistent with an overdeepening. This observation leads to a theory of ice motion in overdeepenings that explains the cyclical behavior of pulsing <span class="hlt">glaciers</span>. It is based on the concept of glaciohydraulic supercooling, and includes sediment transport and erosion</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-ED04-0056-101.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-ED04-0056-101.html"><span><span class="hlt">Glacier</span> Grey view from Lago Grey (Grey Lake), photographed during NASA's AirSAR 2004 campaign in Chile</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2004-03-11</p> <p><span class="hlt">Glacier</span> Grey view from Lago Grey (Grey Lake), photographed during NASA's AirSAR 2004 campaign in Chile. Land visible in this photo was covered by <span class="hlt">glacier</span> just 6 years earlier. AirSAR 2004 is a three-week expedition in Central and South America by an international team of scientists that is using an all-weather imaging tool, called the Airborne Synthetic Aperture Radar (AirSAR), located onboard NASA's DC-8 airborne laboratory. Scientists from many parts of the world are combining ground research with NASA's AirSAR technology to improve and expand on the quality of research they are able to conduct. Founded in 1959, Torres del Paine National <span class="hlt">Park</span> encompasses 450,000 acres in the Patagonia region of Chile. This region is being studied by NASA using a DC-8 equipped with an Airborne Synthetic Aperture Radar (AirSAR) developed by scientists from NASA’s Jet Propulsion Laboratory. This is a very sensitive region that is important to scientists because the temperature has been consistently rising causing a subsequent melting of the region’s <span class="hlt">glaciers</span>. AirSAR will provide a baseline model and unprecedented mapping of the region. This data will make it possible to determine whether the warming trend is slowing, continuing or accelerating. AirSAR will also provide reliable information on ice shelf thickness to measure the contribution of the <span class="hlt">glaciers</span> to sea level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70155990','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70155990"><span>Surface melt dominates Alaska <span class="hlt">glacier</span> mass balance</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Larsen Chris F,; Burgess, E; Arendt, A.A.; O'Neel, Shad; Johnson, A.J.; Kienholz, C.</p> <p>2015-01-01</p> <p>Mountain <span class="hlt">glaciers</span> comprise a small and widely distributed fraction of the world's terrestrial ice, yet their rapid losses presently drive a large percentage of the cryosphere's contribution to sea level rise. Regional mass balance assessments are challenging over large <span class="hlt">glacier</span> populations due to remote and rugged geography, variable response of individual <span class="hlt">glaciers</span> to climate change, and episodic calving losses from tidewater <span class="hlt">glaciers</span>. In Alaska, we use airborne altimetry from 116 <span class="hlt">glaciers</span> to estimate a regional mass balance of −75 ± 11 Gt yr−1 (1994–2013). Our <span class="hlt">glacier</span> sample is spatially well distributed, yet pervasive variability in mass balances obscures geospatial and climatic relationships. However, for the first time, these data allow the partitioning of regional mass balance by <span class="hlt">glacier</span> type. We find that tidewater <span class="hlt">glaciers</span> are losing mass at substantially slower rates than other <span class="hlt">glaciers</span> in Alaska and collectively contribute to only 6% of the regional mass loss.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C41B1217V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C41B1217V"><span>Comparison of Glaciological and Gravimetric <span class="hlt">Glacier</span> Mass Balance Measurements of Taku and Lemon Creek <span class="hlt">Glaciers</span>, Southeast Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vogler, K.; McNeil, C.; Bond, M.; Getraer, B.; Huxley-Reicher, B.; McNamara, G.; Reinhardt-Ertman, T.; Silverwood, J.; Kienholz, C.; Beedle, M. J.</p> <p>2017-12-01</p> <p><span class="hlt">Glacier</span>-wide annual mass balances (Ba) have been calculated for Taku (726 km2) and Lemon Creek <span class="hlt">glaciers</span> (10.2 km2) since 1946 and 1953 respectively. These are the longest mass balance records in North America, and the only Ba time-series available for Southeast Alaska, making them particularly valuable for the global <span class="hlt">glacier</span> mass balance monitoring network. We compared Ba time-series from Taku and Lemon Creek <span class="hlt">glaciers</span> to Gravity Recovery and Climate Experiment (GRACE) mascon solutions (1352 and 1353) during the 2004-2015 period to assess how well these gravimetric solutions reflect individual glaciological records. Lemon Creek <span class="hlt">Glacier</span> is a challenging candidate for this comparison because it is small compared to the 12,100 km2 GRACE mascon solutions. Taku <span class="hlt">Glacier</span> is equally challenging because its mass balance is stable compared to the negative balances dominating its neighboring <span class="hlt">glaciers</span>. Challenges notwithstanding, a high correlation between the glaciological and gravimetrically-derived Ba for Taku and Lemon Creek <span class="hlt">glaciers</span> encourage future use of GRACE to measure <span class="hlt">glacier</span> mass balance. Additionally, we employed high frequency ground penetrating radar (GPR) to measure the variability of accumulation around glaciological sites to assess uncertainty in our glaciological measurements, and the resulting impact to Ba. Finally, we synthesize this comparison of glaciological and gravimetric mass balance solutions with a discussion of potential sources of error in both methods and their combined utility for measuring regional <span class="hlt">glacier</span> change during the 21st century.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001486.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001486.html"><span><span class="hlt">Glaciers</span> and Sea Level Rise</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Calving front of the Upsala <span class="hlt">Glacier</span> (Argentina). This <span class="hlt">glacier</span> has been thinning and retreating at a rapid rate during the last decades – from 2006 to 2010, it receded 43.7 yards (40 meters) per year. During summer 2012, large calving events prevented boat access to the <span class="hlt">glacier</span>. To learn about the contributions of <span class="hlt">glaciers</span> to sea level rise, visit: www.nasa.gov/topics/earth/features/<span class="hlt">glacier</span>-sea-rise.html Credit: Etienne Berthier, Université de Toulouse 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA20745.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA20745.html"><span>Glorious <span class="hlt">Glacier</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2016-07-15</p> <p>This image has low-sun lighting that accentuates the many transverse ridges on this slope, extending from Euripus Mons (mountains). These flow-like structures were previously called "lobate debris aprons," but the Shallow Radar (SHARAD) instrument on MRO has shown that they are actually debris-covered flows of ice, or <span class="hlt">glaciers</span>. There is no evidence for present-day flow of these <span class="hlt">glaciers</span>, so they appear to be remnants of past climates. http://photojournal.jpl.nasa.gov/catalog/PIA20745</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFMPP44A..01L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFMPP44A..01L"><span>Sediments Exposed by Drainage of a Collapsing <span class="hlt">Glacier</span>-Dammed Lake Show That Contemporary Summer Temperatures and <span class="hlt">Glacier</span> Retreat Exceed the Medieval Warm Period in Southern Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loso, M. G.; Anderson, R. S.; Anderson, S. P.; Reimer, P. J.</p> <p>2007-12-01</p> <p>In the mountains of southcentral Alaska, recent and widespread <span class="hlt">glacier</span> retreat is well-documented, but few instrumental or proxy records of temperature are available to place recent changes in a long-term context. The Medieval Warm Period in particular, is poorly documented because subsequent Little Ice Age <span class="hlt">glacier</span> advances destroyed much of the existing sedimentary record. In a rare exception, sudden and unexpected catastrophic drainage of a previously stable <span class="hlt">glacier</span>-dammed lake recently revealed lacustrine stratigraphy that spans over 1500 years. Located near the Bagley Icefield in Wrangell-St. Elias National <span class="hlt">Park</span> and Preserve, Iceberg Lake first drained in A.D. 1999 and has not regained a stable shoreline since that time. Rapid incision of the exposed lakebed provided subaerial exposure of annual laminations (varves, confirmed by radiogenic evidence) that record continuous sediment deposition from A.D. 442 to A.D. 1998. We present a recalculated master chronology of varve thickness that combines measurements from several sites within the former lake. Varve thickness in this chronology is positively correlated with northern hemisphere temperature trends and also with a local, ~600 year long tree ring width chronology. Varve thickness increases in warm summers because of higher melt, runoff, and sediment transport, and also because shrinkage of the <span class="hlt">glacier</span> dam allows shoreline regression that concentrates sediment in the smaller lake. Relative to the entire record, varve thicknesses and implied summer temperatures were lowest around A.D. 600, high between A.D. 1000 and A.D. 1300, low between A.D. 1500 and A.D 1850, and highest in the late 20th century. Combined with stratigraphic evidence that contemporary jokulhlaups are unprecedented since at least A.D. 442, this record suggests that late 20th century warming was more intense, and accompanied by more extensive <span class="hlt">glacier</span> retreat, than the Medieval Warm Period or any other time in the last 1500 years. We emphasize</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1710688F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1710688F"><span>Testing geographical and climatic controls on <span class="hlt">glacier</span> retreat</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Freudiger, Daphné; Stahl, Kerstin; Weiler, Markus</p> <p>2015-04-01</p> <p><span class="hlt">Glacier</span> melt provides an important part of the summer discharge in many mountainous basins. The understanding of the processes behind the <span class="hlt">glacier</span> mass losses and <span class="hlt">glacier</span> retreats observed during the last century is therefore relevant for a sustainable management of the water resources and reliable models for the prediction of future changes. The changes in <span class="hlt">glacier</span> area of 49 sub-basins of the Rhine River in the Alps were analyzed for the time period 1900-2010 by comparing the <span class="hlt">glacier</span> areas of Siegfried maps for the years 1900 and 1940 with satellite derived <span class="hlt">glacier</span> areas for the years 1973, 2003 and 2010. The aim was to empirically investigate the controls of <span class="hlt">glacier</span> retreat and its regional differences. All <span class="hlt">glaciers</span> in the <span class="hlt">glacierized</span> basins retreated over the last 110 years with some variations in the sub-periods. However, the relative changes in <span class="hlt">glacier</span> area compared to 1900 differed for every sub-basin and some <span class="hlt">glaciers</span> decreased much faster than others. These observed differences were related to a variety of different potential controls derived from different sources, including mean annual solar radiation on the <span class="hlt">glacier</span> surface, average slope, mean <span class="hlt">glacier</span> elevation, initial <span class="hlt">glacier</span> area, average precipitation (summer and winter), and the precipitation catchment area of the <span class="hlt">glacier</span>. We fitted a generalized linear model (GLM) and selected predictors that were significant to assess the individual effects of the potential controls. The fitted model explains more than 60% of the observed variance of the relative change in <span class="hlt">glacier</span> area with the initial area alone only explaining a small proportion. Some interesting patterns emerge with higher average elevation resulting in higher area changes, but steeper slopes or solar radiation resulting in lower relative <span class="hlt">glacier</span> area changes. Further controls that will be tested include snow transport by wind or avalanches as they play an important role for the <span class="hlt">glacier</span> mass balance and potentially reduce the changes in <span class="hlt">glacier</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C41B1199R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C41B1199R"><span>Recent Developments of the GLIMS <span class="hlt">Glacier</span> Database</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raup, B. H.; Berthier, E.; Bolch, T.; Kargel, J. S.; Paul, F.; Racoviteanu, A.</p> <p>2017-12-01</p> <p>Earth's <span class="hlt">glaciers</span> are shrinking almost without exception, leading to changes in water resources, timing of runoff, sea level, and hazard potential. Repeat mapping of <span class="hlt">glacier</span> outlines, lakes, and <span class="hlt">glacier</span> topography, along with glacial processes, is critically needed to understand how <span class="hlt">glaciers</span> will react to a changing climate, and how those changes will impact humans. To understand the impacts and processes behind the observed changes, it is crucial to monitor <span class="hlt">glaciers</span> through time by mapping their areal extent, snow lines, ice flow velocities, associated water bodies, and thickness changes. The <span class="hlt">glacier</span> database of the Global Land Ice Measurements from Space (GLIMS) initiative is the only multi-temporal <span class="hlt">glacier</span> database capable of tracking all these <span class="hlt">glacier</span> measurements and providing them to the scientific community and broader public.Recent developments in GLIMS include improvements in the database and web applications and new activities in the international GLIMS community. The coverage of the GLIMS database has recently grown geographically and temporally by drawing on the Randolph <span class="hlt">Glacier</span> Inventory (RGI) and other new data sets. The GLIMS database is globally complete, and approximately one third of <span class="hlt">glaciers</span> have outlines from more than one time. New tools for visualizing and downloading GLIMS data in a choice of formats and data models have been developed, and a new data model for handling multiple <span class="hlt">glacier</span> records through time while avoiding double-counting of <span class="hlt">glacier</span> number or area is nearing completion. A GLIMS workshop was held in Boulder, Colorado this year to facilitate two-way communication with the greater community on future needs.The result of this work is a more complete and accurate <span class="hlt">glacier</span> data repository that shows both the current state of <span class="hlt">glaciers</span> on Earth and how they have changed in recent decades. Needs for future scientific and technical developments were identified and prioritized at the GLIMS Workshop, and are reported here.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMEP33E..08K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMEP33E..08K"><span><span class="hlt">Glacier</span>, Glacial Lake, and Ecological Response Dynamics of the Imja <span class="hlt">Glacier</span>-Lake-Moraine System, Nepal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kargel, J. S.; Shugar, D. H.; Leonard, G. J.; Haritashya, U. K.; Harrison, S.; Shrestha, A. B.; Mool, P. K.; Karki, A.; Regmi, D.</p> <p>2016-12-01</p> <p><span class="hlt">Glacier</span> response dynamics—involving a host of processes—produce a sequence of short- to long-term delayed responses to any step-wise, oscillating, or continuous trending climatic perturbation. We present analysis of Imja Lake, Nepal and examine its thinning and retreat and a sequence of the detachment of tributaries; the inception and growth of Imja Lake and concomitant <span class="hlt">glacier</span> retreat, thinning, and stagnation, and relationships to lake dynamics; the response dynamics of the ice-cored moraine; the development of the local ecosystem; prediction of short-term dynamical responses to lake lowering (<span class="hlt">glacier</span> lake outburst flood—GLOF—mitigation); and prospects for coming decades. The evolution of this <span class="hlt">glacier</span> system provides a case study by which the global record of GLOFs can be assessed in terms of climate change attribution. We define three response times: <span class="hlt">glacier</span> dynamical response time (for <span class="hlt">glacier</span> retreat, thinning, and slowing of ice flow), limnological response time (lake growth), and GLOF trigger time (for a variety of hazardous trigger events). Lake lowering (to be completed in August 2016; see AGU abstract by D. Regmi et al.) will reduce hazards, but we expect that the elongation of the lake and retreat of the <span class="hlt">glacier</span> will continue for decades after a pause in 2016-2017. The narrowing of the moraine dam due to thaw degradation of the ice-cored end moraine means that the hazard due to Imja Lake will soon again increase. We examine both long-term response dynamics, and two aspects of Himalayan <span class="hlt">glaciers</span> that have very rapid responses: the area of Imja Lake fluctuates seasonally and even with subseasonal weather variations in response to changes in lake temperature and <span class="hlt">glacier</span> meltback; and as known from other studies, <span class="hlt">glacier</span> flow speed can vary between years and even on shorter timescales. The long-term development and stabilization of glacial moraines and small lacustrine plains in drained lake basins impacts the development of local ecosystems</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70019284','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70019284"><span><span class="hlt">Glacier</span> generated floods</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Walder, J.S.; Fountain, A.G.; ,</p> <p>1997-01-01</p> <p>Destructive floods result from drainage of <span class="hlt">glacier</span>-dammed lakes and sudden release of water stored within <span class="hlt">glaciers</span>. There is a good basis - both empirical and theoretical - for predicting the magnitude of floods from ice-dammed lakes, although some aspects of flood initiation need to be better understood. In contrast, an understanding of floods resulting from release of internally stored water remains elusive, owing to lack of knowledge of how and where water is stored and to inadequate understanding of the complex physics of the temporally and spatially variable subglacial drainage system.Destructive floods result from drainage of <span class="hlt">glacier</span>-dammed lakes and sudden release of water stored within <span class="hlt">glaciers</span>. There is a good basis - both empirical and theoretical - for predicting the magnitude of floods from ice-dammed lakes, although some aspects of flood initiation need to be better understood. In contrast, an understanding of floods resulting from release of internally stored water remains elusive, owing to lack of knowledge of how and where water is stored and to inadequate understanding of the complex physics of the temporally and spatially variable subglacial drainage system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..16.4842H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16.4842H"><span>Internationally coordinated <span class="hlt">glacier</span> monitoring: strategy and datasets</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hoelzle, Martin; Armstrong, Richard; Fetterer, Florence; Gärtner-Roer, Isabelle; Haeberli, Wilfried; Kääb, Andreas; Kargel, Jeff; Nussbaumer, Samuel; Paul, Frank; Raup, Bruce; Zemp, Michael</p> <p>2014-05-01</p> <p>Internationally coordinated monitoring of long-term <span class="hlt">glacier</span> changes provide key indicator data about global climate change and began in the year 1894 as an internationally coordinated effort to establish standardized observations. Today, world-wide monitoring of <span class="hlt">glaciers</span> and ice caps is embedded within the Global Climate Observing System (GCOS) in support of the United Nations Framework Convention on Climate Change (UNFCCC) as an important Essential Climate Variable (ECV). The Global Terrestrial Network for <span class="hlt">Glaciers</span> (GTN-G) was established in 1999 with the task of coordinating measurements and to ensure the continuous development and adaptation of the international strategies to the long-term needs of users in science and policy. The basic monitoring principles must be relevant, feasible, comprehensive and understandable to a wider scientific community as well as to policy makers and the general public. Data access has to be free and unrestricted, the quality of the standardized and calibrated data must be high and a combination of detailed process studies at selected field sites with global coverage by satellite remote sensing is envisaged. Recently a GTN-G Steering Committee was established to guide and advise the operational bodies responsible for the international <span class="hlt">glacier</span> monitoring, which are the World <span class="hlt">Glacier</span> Monitoring Service (WGMS), the US National Snow and Ice Data Center (NSIDC), and the Global Land Ice Measurements from Space (GLIMS) initiative. Several online databases containing a wealth of diverse data types having different levels of detail and global coverage provide fast access to continuously updated information on <span class="hlt">glacier</span> fluctuation and inventory data. For world-wide inventories, data are now available through (a) the World <span class="hlt">Glacier</span> Inventory containing tabular information of about 130,000 <span class="hlt">glaciers</span> covering an area of around 240,000 km2, (b) the GLIMS-database containing digital outlines of around 118,000 <span class="hlt">glaciers</span> with different time stamps and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sir/2007/5147/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sir/2007/5147/"><span>Assessment of Historical Water-Quality Data for National <span class="hlt">Park</span> Units in the Rocky Mountain Network, Colorado and Montana, through 2004</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Mast, M. Alisa</p> <p>2007-01-01</p> <p>This report summarizes historical water-quality data for six National <span class="hlt">Park</span> units that compose the Rocky Mountain Network. The <span class="hlt">park</span> units in Colorado are Florissant Fossil Beds National Monument, Great Sand Dunes National <span class="hlt">Park</span> and Preserve, and Rocky Mountain National <span class="hlt">Park</span>; and in Montana, they are <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Grant-Kohrs Ranch National Historic Site, and Little Bighorn Battlefield National Monument. This study was conducted in cooperation with the Inventory and Monitoring Program of the National <span class="hlt">Park</span> Service to aid in the design of an effective and efficient water-quality monitoring plan for each <span class="hlt">park</span>. Data were retrieved from a number of sources for the period of record through 2004 and compiled into a relational database. Descriptions of the environmental setting of each <span class="hlt">park</span> and an overview of the <span class="hlt">park</span>'s water resources are presented. Statistical summaries of water-quality constituents are presented and compared to aquatic-life and drinking-water standards. Spatial, seasonal, and temporal patterns in constituent concentrations also are described and suggestions for future water-quality monitoring are provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C42B..05M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C42B..05M"><span>Quantifying Tropical <span class="hlt">Glacier</span> Mass Balance Sensitivity to Climate Change Through Regional-Scale Modeling and The Randolph <span class="hlt">Glacier</span> Inventory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malone, A.</p> <p>2017-12-01</p> <p>Quantifying mass balance sensitivity to climate change is essential for forecasting <span class="hlt">glacier</span> evolution and deciphering climate signals embedded in archives of past <span class="hlt">glacier</span> changes. Ideally, these quantifications result from decades of field measurement, remote sensing, and a hierarchy modeling approach, but in data-sparse regions, such as the Himalayas and tropical Andes, regional-scale modeling rooted in first principles provides a first-order picture. Previous regional-scaling modeling studies have applied a surface energy and mass balance approach in order to quantify equilibrium line altitude sensitivity to climate change. In this study, an expanded regional-scale surface energy and mass balance model is implemented to quantify <span class="hlt">glacier</span>-wide mass balance sensitivity to climate change for tropical Andean <span class="hlt">glaciers</span>. Data from the Randolph <span class="hlt">Glacier</span> Inventory are incorporated, and additional physical processes are included, such as a dynamic albedo and cloud-dependent atmospheric emissivity. The model output agrees well with the limited mass balance records for tropical Andean <span class="hlt">glaciers</span>. The dominant climate variables driving interannual mass balance variability differ depending on the climate setting. For wet tropical <span class="hlt">glaciers</span> (annual precipitation >0.75 m y-1), temperature is the dominant climate variable. Different hypotheses for the processes linking wet tropical <span class="hlt">glacier</span> mass balance variability to temperature are evaluated. The results support the hypothesis that <span class="hlt">glacier</span>-wide mass balance on wet tropical <span class="hlt">glaciers</span> is largely dominated by processes at the lowest elevation where temperature plays a leading role in energy exchanges. This research also highlights the transient nature of wet tropical <span class="hlt">glaciers</span> - the vast majority of tropical <span class="hlt">glaciers</span> and a vital regional water resource - in an anthropogenic warming world.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/prof/p1386c/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/prof/p1386c/"><span><span class="hlt">Glaciers</span> of Greenland</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Williams, Richard S.; Ferrigno, Jane G.</p> <p>1995-01-01</p> <p>Landsat imagery, combined with aerial photography, sketch maps, and diagrams, is used as the basis for a description of the geography, climatology, and glaciology, including mass balance, variation, and hazards, of the Greenland ice sheet and local ice caps and <span class="hlt">glaciers</span>. The Greenland ice sheet, with an estimated area of 1,736,095+/-100 km2 and volume of 2,600,000 km3, is the second largest <span class="hlt">glacier</span> on the planet and the largest relict of the Ice Age in the Northern Hemisphere. Greenland also has 48,599+/-100 km2 of local ice caps and other types of <span class="hlt">glaciers</span> in coastal areas and islands beyond the margin of the ice sheet.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AGUFM.C31C0454Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AGUFM.C31C0454Y"><span>The recent <span class="hlt">glacier</span> changes in Mongolian Altai Mountains</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yabuki, H.; Ohata, T.</p> <p>2009-12-01</p> <p>In the 4th IPCC report (AR-4) is reported that global warming in recent years is a clear thing. Shrinkage of the mountain <span class="hlt">glacier</span> and two poles is reporting as an observation fact as the actual condition of the cryosphere by warming. There are mass balance reports of the <span class="hlt">glacier</span> of 80 of world by WGMS (World <span class="hlt">Glacier</span> Monitoring Service) as a report of the actual condition of <span class="hlt">glacier</span> mass balance change, and the actual condition of the <span class="hlt">glacier</span> mass change in world is clarified. In the report of WGMS, after 1980’s the <span class="hlt">glacier</span> mass balance, in the Europe Alps and the Alaska region are decreases, and in Scandinavia region are increases. On the other hand, the <span class="hlt">glacier</span> mass balance in the Russia Altai Mountains located in Central Asia has the little change after 1980’s. These are research using the long-term observational data of Russian region of western part of Altai Mountains. The Altai Mountains including Russia, China, and Mongolia Kazakhstan, and there are description to a World <span class="hlt">Glacier</span> Inventory (WGI) about the <span class="hlt">glaciers</span> of Russia, China and Kazakhstan area, but the <span class="hlt">glaciers</span> of a Mongolian area, there are no description to the WGI. There is almost no information on the <span class="hlt">glacier</span> of a Mongolian Altai region, and there are many unknown points about <span class="hlt">glacier</span> change of the whole Altai Mountain region. In this research, while research clarified the present condition of <span class="hlt">glacier</span> distribution of the Mongolia Altai region, the actual condition of a <span class="hlt">glacier</span> change in recent years was clarified by comparison with the past topographical map. In this research, the <span class="hlt">glacier</span> area was distinguished based on the satellite image of the Mongolian <span class="hlt">glacier</span> regions. The used satellite image were 17 Landsat 7 ETM+ in 1999 to 2002. The <span class="hlt">glacier</span> distinguishes using NDSI (Normalized Difference Snow Index) indexusing Band5 and Band2. The topographical map of the Mongolian area was got based on the distribution information on this satellite <span class="hlt">glacier</span> area. The topographical map is 1/100,000 which</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11878639','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11878639"><span>Assessment of climate change effects on Canada's National <span class="hlt">Park</span> system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Suffling, Roger; Scott, Daniel</p> <p>2002-03-01</p> <p>To estimate the magnitude of climate change anticipated for Canada's 38 National <span class="hlt">Parks</span> (NPs) and <span class="hlt">Park</span> Reserves, seasonal temperature and precipitation scenarios were constructed for 2050 and 2090 using the Canadian Centre for Climate Modelling and Analysis (CCCma) coupled model (CGCM1). For each <span class="hlt">park</span>, we assessed impacts on physical systems, species, ecosystems and people. Important, widespread changes relate to marine and freshwater hydrology, glacial balance, waning permafrost, increased natural disturbance, shorter ice season, northern and upward altitudinal species and biome shifts, and changed visitation patterns. Other changes are regional (e.g., combined <span class="hlt">East</span> coast subsidence and sea level rise increase coastal erosion and deposition, whereas, on the Pacific coast, tectonic uplift negates sea level rise). Further predictions concern individual <span class="hlt">parks</span> (e.g., Unique fens of Bruce Peninsular NP will migrate lakewards with lowered water levels, but structural regulation of Lake Huron for navigation and power generation would destroy the fens). Knowledge gaps are the most important findings. For example: we could not form conclusions about glacial mass balance, or its effects on rivers and fjords. Likewise, for the <span class="hlt">East</span> Coast Labrador Current we could neither estimate temperature and salinity effects of extra iceberg formation, nor the further effects on marine food chains, and breeding <span class="hlt">park</span> seabirds. We recommend 1) Research on specific large knowledge gaps; 2) Climate change information exchange with protected area agencies in other northern countries; and 3) incorporating climate uncertainty into <span class="hlt">park</span> plans and management. We discuss options for a new <span class="hlt">park</span> management philosophy in the face of massive change and uncertainty.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-sts068-247-061.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-sts068-247-061.html"><span>Yellowstone Lake/National <span class="hlt">Park</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1994-09-30</p> <p>STS068-247-061 (30 September-11 October 1994) --- Photographed through the Space Shuttle Endeavour's flight windows, this 70mm frame centers on Yellowstone Lake in the Yellowstone National <span class="hlt">Park</span>. North will be at the top if picture is oriented with series of sun glinted creeks and river branches at top center. The lake, at 2,320 meters (7,732 feet) above sea level, is the largest high altitude lake in North America. <span class="hlt">East</span> of the <span class="hlt">park</span> part of the Absaroka Range can be traced by following its north to south line of snow capped peaks. Jackson Lake is southeast of Yellowstone <span class="hlt">Park</span>, and the connected Snake River can be seen in the lower left corner. Yellowstone, established in 1872 is the world's oldest national <span class="hlt">park</span>. It covers an area of 9,000 kilometers (3,500 square miles), lying mainly on a broad plateau of the Rocky Mountains on the Continental Divide. It's average altitude is 2,440 meters (8,000 feet) above sea level. The plateau is surrounded by mountains exceeding 3,600 meters (12,000 feet) in height. Most of the plateau was formed from once-molten lava flows, the last of which is said to have occurred 100,000 years ago. Early volcanic activity is still evident in the region by nearly 10,000 hot springs, 200 geysers and numerous vents found throughout the <span class="hlt">park</span>.</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('https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1326.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1326.pdf"><span>36 CFR 13.1326 - Snowmachines.</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>2014-07-01</p> <p>... NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National <span class="hlt">Park</span> Exit <span class="hlt">Glacier</span> Developed...) On Exit <span class="hlt">Glacier</span> Road; (b) In <span class="hlt">parking</span> areas; (c) On a designated route through the Exit <span class="hlt">Glacier</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title36-vol1/pdf/CFR-2011-title36-vol1-sec13-1326.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title36-vol1/pdf/CFR-2011-title36-vol1-sec13-1326.pdf"><span>36 CFR 13.1326 - Snowmachines.</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>... NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National <span class="hlt">Park</span> Exit <span class="hlt">Glacier</span> Developed...) On Exit <span class="hlt">Glacier</span> Road; (b) In <span class="hlt">parking</span> areas; (c) On a designated route through the Exit <span class="hlt">Glacier</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title36-vol1/pdf/CFR-2013-title36-vol1-sec13-1326.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title36-vol1/pdf/CFR-2013-title36-vol1-sec13-1326.pdf"><span>36 CFR 13.1326 - Snowmachines.</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>2013-07-01</p> <p>... NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National <span class="hlt">Park</span> Exit <span class="hlt">Glacier</span> Developed...) On Exit <span class="hlt">Glacier</span> Road; (b) In <span class="hlt">parking</span> areas; (c) On a designated route through the Exit <span class="hlt">Glacier</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2010-title36-vol1/pdf/CFR-2010-title36-vol1-sec13-1326.pdf','CFR'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2010-title36-vol1/pdf/CFR-2010-title36-vol1-sec13-1326.pdf"><span>36 CFR 13.1326 - Snowmachines.</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>... NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National <span class="hlt">Park</span> Exit <span class="hlt">Glacier</span> Developed...) On Exit <span class="hlt">Glacier</span> Road; (b) In <span class="hlt">parking</span> areas; (c) On a designated route through the Exit <span class="hlt">Glacier</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUFMIN41A0066R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUFMIN41A0066R"><span>The GLIMS <span class="hlt">Glacier</span> Database</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raup, B. H.; Khalsa, S. S.; Armstrong, R.</p> <p>2007-12-01</p> <p>The Global Land Ice Measurements from Space (GLIMS) project has built a geospatial and temporal database of <span class="hlt">glacier</span> data, composed of <span class="hlt">glacier</span> outlines and various scalar attributes. These data are being derived primarily from satellite imagery, such as from ASTER and Landsat. Each "snapshot" of a <span class="hlt">glacier</span> is from a specific time, and the database is designed to store multiple snapshots representative of different times. We have implemented two web-based interfaces to the database; one enables exploration of the data via interactive maps (web map server), while the other allows searches based on text-field constraints. The web map server is an Open Geospatial Consortium (OGC) compliant Web Map Server (WMS) and Web Feature Server (WFS). This means that other web sites can display <span class="hlt">glacier</span> layers from our site over the Internet, or retrieve <span class="hlt">glacier</span> features in vector format. All components of the system are implemented using Open Source software: Linux, PostgreSQL, PostGIS (geospatial extensions to the database), MapServer (WMS and WFS), and several supporting components such as Proj.4 (a geographic projection library) and PHP. These tools are robust and provide a flexible and powerful framework for web mapping applications. As a service to the GLIMS community, the database contains metadata on all ASTER imagery acquired over <span class="hlt">glacierized</span> terrain. Reduced-resolution of the images (browse imagery) can be viewed either as a layer in the MapServer application, or overlaid on the virtual globe within Google Earth. The interactive map application allows the user to constrain by time what data appear on the map. For example, ASTER or <span class="hlt">glacier</span> outlines from 2002 only, or from Autumn in any year, can be displayed. The system allows users to download their selected <span class="hlt">glacier</span> data in a choice of formats. The results of a query based on spatial selection (using a mouse) or text-field constraints can be downloaded in any of these formats: ESRI shapefiles, KML (Google Earth), Map</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1813607V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1813607V"><span>Detecting <span class="hlt">glacier</span>-bed overdeepenings for <span class="hlt">glaciers</span> in the Western Italian Alps using the GlabTop2 model: the test site of the Rutor <span class="hlt">Glacier</span>, Aosta Valley</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Viani, Cristina; Machguth, Horst; Huggel, Christian; Perotti, Luigi; Giardino, Marco</p> <p>2016-04-01</p> <p>It is expected that the rapid retreat of <span class="hlt">glaciers</span>, observed in the European Alps and other mountain regions of the world, will continue in the future. One of the most evident and relevant consequences of this phenomenon is the formation of new <span class="hlt">glacier</span> lakes in recently deglaciated areas. During <span class="hlt">glacier</span> retreat overdeepened parts of the <span class="hlt">glacier</span> bed become exposed and, in some cases, filled with water. It is important to understand where these new lakes can appear because of the associated potential risks (i.e. lake outburst and consequent flood) and opportunities (tourism, hydroelectricity, water reservoir, etc.) especially in densely populated areas such as the European Alps. GlabTop2 (<span class="hlt">Glacier</span> Bed Topography model version 2) allows to model <span class="hlt">glacier</span> bed topography over large glaciated areas combining digital terrain information and slope-related estimates of <span class="hlt">glacier</span> thickness. The model requires a minimum set of input data: <span class="hlt">glaciers</span> outlines and a surface digital elevation model (DEM). In this work we tested the model on the Rutor <span class="hlt">Glacier</span> (8,1 km2) located in the Aosta Valley. The <span class="hlt">glacier</span> has a well-known history of a series of <span class="hlt">glacier</span> lake outburst floods between 1430 AD and 1864 AD due to front fluctuations. After the last advance occurred during the 70s of the previous century, <span class="hlt">glacier</span> shrinkage has been continuous and new lakes have formed in newly exposed overdeepenings. We applied GlabTop2 to DEMs derived from historical data (topographic maps and aerial photos pair) representing conditions before the proglacial lake formation. The results obtained have been compared with the present situation and existing lakes. Successively we used the model also on present-day DEMs, which are of higher resolution than the historical derived ones, and compared the modeled bed topography with an existing bedrock map obtained by in-situ geophysical investigations (GPR surveys). Preliminary results, obtained with the 1991 surface model, confirm the robustness of GlabTop2 in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C13D0872J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C13D0872J"><span>Monitoring Jakobshavn <span class="hlt">Glacier</span> using Sequential Landsat Images</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jian, Z.; Zhuoqi, C.; Cheng, X.</p> <p>2016-12-01</p> <p>Jakobshavn <span class="hlt">Glacier</span> is the fastest (19 m per day) and one of the most active <span class="hlt">glaciers</span> around the world. Discharging more than 35km3 of ice every year, its mass loss surpasses anyone else outside the Antarctic. From Landsat 8 OLI Images on August 14, 2015, we find a huge iceberg about 5 km2 calved from resulting in the front shrinking for 1060.8m. NSIDC ice velocity data and weather station data on Jakobshavn <span class="hlt">glacier</span> are used to analyze the cause of calving. On one hand, upstream <span class="hlt">glacier</span> push forward the Jakobshavn <span class="hlt">glacier</span> westward continually, many cracks were formed over the <span class="hlt">glacier</span> surface. Surface melting water flow into the interior of <span class="hlt">glaciers</span> to accelerate calving. On the other hand with the gradually rising temperature, the bottom of <span class="hlt">glaciers</span> accelerate ablation. When <span class="hlt">glaciers</span> move into the ocean and the thin bottom can not provide strong enough support, calving occurs. Before this incident, we trace sequential Landsat data during 1986 to 2015. In 2010, it had another large-scale calving. We draw from our data that Jakobshavn retreated intensely in the past 30 years although in the last 10 years it appears more stable. The speed of <span class="hlt">glacier</span> shrinking during 1996 to 2006 is three times as fast as past 10 years.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EOSTr..93..212K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EOSTr..93..212K"><span><span class="hlt">Glaciers</span> in Patagonia: Controversy and prospects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kargel, J. S.; Alho, P.; Buytaert, W.; Célleri, R.; Cogley, J. G.; Dussaillant, A.; Guido, Z.; Haeberli, W.; Harrison, S.; Leonard, G.; Maxwell, A.; Meier, C.; Poveda, G.; Reid, B.; Reynolds, J.; Rodríguez, C. A. Portocarrero; Romero, H.; Schneider, J.</p> <p>2012-05-01</p> <p>Lately, <span class="hlt">glaciers</span> have been subjects of unceasing controversy. Current debate about planned hydroelectric facilities—a US7- to 10-billion megaproject—in a pristine <span class="hlt">glacierized</span> area of Patagonia, Chile [Romero Toledo et al., 2009; Vince, 2010], has raised anew the matter of how glaciologists and global change experts can contribute their knowledge to civic debates on important issues. There has been greater respect for science in this controversy than in some previous debates over projects that pertain to <span class="hlt">glaciers</span>, although valid economic motivations again could trump science and drive a solution to the energy supply problem before the associated safety and environmental problems are understood. The connection between <span class="hlt">glaciers</span> and climate change—both anthropogenic and natural—is fundamental to glaciology and to <span class="hlt">glaciers</span>' practical importance for water and hydropower resources, agriculture, tourism, mining, natural hazards, ecosystem conservation, and sea level [Buytaert et al., 2010; Glasser et al., 2011]. The conflict between conservation and development can be sharper in <span class="hlt">glacierized</span> regions than almost anywhere else. <span class="hlt">Glaciers</span> occur in spectacular natural landscapes, but they also supply prodigious exploitable meltwater.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7184033-glacier-recession-iceland-austria','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/7184033-glacier-recession-iceland-austria"><span><span class="hlt">Glacier</span> recession in Iceland and Austria</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>Hall, D.K.; Williams, R.S. Jr.; Bayr, K.J.</p> <p>1992-03-01</p> <p>It has been possible to measure <span class="hlt">glacier</span> recession on the basis of Landsat data, in conjunction with comparisons of the magnitude of recession of a <span class="hlt">glacier</span> margin with in situ measurements at fixed points along the same margin. Attention is presently given to the cases of Vatnajokull ice cap, in Iceland, and the Pasterze <span class="hlt">Glacier</span>, in Austria, on the basis of satellite data from 1973-1987 and 1984-1990, respectively. Indications of a trend toward negative mass balance are noted. Nevertheless, while most of the world's small <span class="hlt">glaciers</span> have been receding, some are advancing either due to local climate or the tidewatermore » <span class="hlt">glacier</span> cycle. 21 refs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70036629','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70036629"><span>Assessment of lake sensitivity to acidic deposition in national <span class="hlt">parks</span> of the Rocky Mountains</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Nanus, L.; Williams, M.W.; Campbell, D.H.; Tonnessen, K.A.; Blett, T.; Clow, D.W.</p> <p>2009-01-01</p> <p>The sensitivity of high-elevation lakes to acidic deposition was evaluated in five national <span class="hlt">parks</span> of the Rocky Mountains based on statistical relations between lake acid-neutralizing capacity concentrations and basin characteristics. Acid-neutralizing capacity (ANC) of 151 lakes sampled during synoptic surveys and basin-characteristic information derived from geographic information system (GIS) data sets were used to calibrate the statistical models. The explanatory basin variables that were considered included topographic parameters, bedrock type, and vegetation type. A logistic regression model was developed, and modeling results were cross-validated through lake sampling during fall 2004 at 58 lakes. The model was applied to lake basins greater than 1 ha in area in <span class="hlt">Glacier</span> National <span class="hlt">Park</span> (n = 244 lakes), Grand Teton National <span class="hlt">Park</span> (n = 106 lakes), Great Sand Dunes National <span class="hlt">Park</span> and Preserve (n = 11 lakes), Rocky Mountain National <span class="hlt">Park</span> (n = 114 lakes), and Yellowstone National <span class="hlt">Park</span> (n = 294 lakes). Lakes that had a high probability of having an ANC concentration 3000 m, with 80% of the catchment bedrock having low buffering capacity. The modeling results indicate that the most sensitive lakes are located in Rocky Mountain National <span class="hlt">Park</span> and Grand Teton National <span class="hlt">Park</span>. This technique for evaluating the lake sensitivity to acidic deposition is useful for designing long-term monitoring plans and is potentially transferable to other remote mountain areas of the United States and the world.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://link.springer.com/article/10.1007/s10584-013-1042-7/fulltext.html','USGSPUBS'); return false;" href="http://link.springer.com/article/10.1007/s10584-013-1042-7/fulltext.html"><span>Assessing streamflow sensitivity to variations in <span class="hlt">glacier</span> mass balance</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>O'Neel, Shad; Hood, Eran; Arendt, Anthony; Sass, Louis</p> <p>2014-01-01</p> <p>The purpose of this paper is to evaluate relationships among seasonal and annual <span class="hlt">glacier</span> mass balances, <span class="hlt">glacier</span> runoff and streamflow in two <span class="hlt">glacierized</span> basins in different climate settings. We use long-term <span class="hlt">glacier</span> mass balance and streamflow datasets from the United States Geological Survey (USGS) Alaska Benchmark <span class="hlt">Glacier</span> Program to compare and contrast <span class="hlt">glacier</span>-streamflow interactions in a maritime climate (Wolverine <span class="hlt">Glacier</span>) with those in a continental climate (Gulkana <span class="hlt">Glacier</span>). Our overall goal is to improve our understanding of how <span class="hlt">glacier</span> mass balance processes impact streamflow, ultimately improving our conceptual understanding of the future evolution of <span class="hlt">glacier</span> runoff in continental and maritime climates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeoRL..4312466K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeoRL..4312466K"><span>Contemporary <span class="hlt">glacier</span> retreat triggers a rapid landslide response, Great Aletsch <span class="hlt">Glacier</span>, Switzerland</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kos, Andrew; Amann, Florian; Strozzi, Tazio; Delaloye, Reynald; Ruette, Jonas; Springman, Sarah</p> <p>2016-12-01</p> <p>The destabilization and catastrophic failure of landslides triggered by retreating <span class="hlt">glaciers</span> is an expected outcome of global climate change and poses a significant threat to inhabitants of glaciated mountain valleys around the globe. Of particular importance are the formation of landslide-dammed lakes, outburst floods, and related sediment entrainment. Based on field observations and remote sensing of a deep-seated landslide, located at the present-day terminus of the Great Aletsch <span class="hlt">Glacier</span>, we show that the spatiotemporal response of the landslide to <span class="hlt">glacier</span> retreat is rapid, occurring within a decade. Our observations uniquely capture the critical period of increase in slope deformations, onset of failure, and show that measured displacements at the crown and toe regions of the landslide demonstrate a feedback mechanism between <span class="hlt">glacier</span> ice reduction and response of the entire landslide body. These observations shed new light on the geomorphological processes of landslide response in paraglacial environments, which were previously understood to occur over significantly longer time periods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFMED21C0594C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFMED21C0594C"><span>Rapidly Deglaciating and Uplifting Landscapes in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve GBNPP Provide Alaskan High School Students with Summer Field Research Experiences in Paleoclimate Disciplines and Exposure to Active Researchers in Synergistic Sciences</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Connor, C. L.; Hekkers, M.; Schaller, S.; Parks, R.</p> <p>2011-12-01</p> <p>During summer 2011, ten Alaska high school students enrolled in a summer college introductory field science course through the Design Discover Research Program (DDR) at University Alaska Southeast, with support from the Juneau Economic Development Council, the University Alaska Fairbanks Alaska Summer Research Academy (ASRA), the National <span class="hlt">Park</span> Service, and the National Oceanic and Atmospheric Administration (NOAA)-Interdisciplinary Scientific Environmental Technology (ISET) Cooperative Science Center (CSC). They conducted field surveys in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and across the adjacent Gustavus <span class="hlt">glacier</span> forefield landscape. This was the 4th summer for this program. Following a 2-1/2 day orientation on the UAS campus, students traveled by a newly established (2011) Alaska Marine Highway service from Juneau to Gustavus. They utilized bicycles and hiking to access the Dude Creek Critical Habitat, Good Creek watershed, and the Nagoonberry Trail built by the Nature Conservancy along the emerging shoreline of Icy Strait. This region is currently experiencing uplift rates of 20-28 mm/year as a result of Little Ice Age deglaciation and isostatic rebound, rates much higher than eustatic sea level rise. North of the Gustavus ferry dock, roughly 19 acres of emergent land is now a 9-hole golf course, raised from the sea since the 1950s. DDR-students interacted with wildlife biologists, ornithologists, quaternary geologists, glaciologists, and botanists to integrate their understanding of the response of plants and animals to this dynamic landscape. Radio-collared moose populations, which migrated into this area from Haines since the mind-1960s, are being studied to asses their impacts to local vegetation and the behavior of local predators (bears, wolves, and coyotes). Sitka spruce forest expansion onto Icy Strait uplifting salt marshes, now threaten Sandhill Crane habitat and their flyover stops along the Pacific Flyway. Intertidal areas in Bartlett Cove in GBNPP feature</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001479.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001479.html"><span><span class="hlt">Glaciers</span> and Sea Level Rise</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>This ice cave in Belcher <span class="hlt">Glacier</span> (Devon Island, Canada) was formed by melt water flowing within the <span class="hlt">glacier</span> ice. To learn about the contributions of <span class="hlt">glaciers</span> to sea level rise, visit: www.nasa.gov/topics/earth/features/<span class="hlt">glacier</span>-sea-rise.html Credit: Angus Duncan, University of Saskatchewan 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70191887','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70191887"><span>Increasing rock-avalanche size and mobility in <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, Alaska detected from 1984 to 2016 Landsat imagery</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Coe, Jeffrey A.; Bessette-Kirton, Erin; Geertsema, Marten</p> <p>2018-01-01</p> <p>In the USA, climate change is expected to have an adverse impact on slope stability in Alaska. However, to date, there has been limited work done in Alaska to assess if changes in slope stability are occurring. To address this issue, we used 30-m Landsat imagery acquired from 1984 to 2016 to establish an inventory of 24 rock avalanches in a 5000-km2 area of <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve in southeast Alaska. A search of available earthquake catalogs revealed that none of the avalanches were triggered by earthquakes. Analyses of rock-avalanche magnitude, mobility, and frequency reveal a cluster of large (areas ranging from 5.5 to 22.2 km2), highly mobile (height/length < 0.3) rock avalanches that occurred from June 2012 through June 2016 (near the end of the 33-year period of record). These rock avalanches began about 2  years after the long-term trend in mean annual maximum air temperature may have exceeded 0 °C. Possibly more important, most of these rock avalanches occurred during a multiple-year period of record-breaking warm winter and spring air temperatures. These observations suggested to us that rock avalanches in the study area may be becoming larger because of rock-permafrost degradation. However, other factors, such as accumulating elastic strain, glacial thinning, and increased precipitation, may also play an important role in preconditioning slopes for failure during periods of warm temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2013-title36-vol1/pdf/CFR-2013-title36-vol1-sec13-1318.pdf','CFR2013'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2013-title36-vol1/pdf/CFR-2013-title36-vol1-sec13-1318.pdf"><span>36 CFR 13.1318 - Location of the EGDA.</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>2013-07-01</p> <p>... NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National <span class="hlt">Park</span> Exit <span class="hlt">Glacier</span> Developed... boundary to Exit <span class="hlt">Glacier</span> Campground Entrance Road, all <span class="hlt">park</span> areas within 350 meters (383 yards) of the centerline of the Exit <span class="hlt">Glacier</span> Road; (2) From Exit <span class="hlt">Glacier</span> Campground Entrance Road to the end of the main...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title36-vol1/pdf/CFR-2011-title36-vol1-sec13-1318.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title36-vol1/pdf/CFR-2011-title36-vol1-sec13-1318.pdf"><span>36 CFR 13.1318 - Location of the EGDA.</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>... NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National <span class="hlt">Park</span> Exit <span class="hlt">Glacier</span> Developed... boundary to Exit <span class="hlt">Glacier</span> Campground Entrance Road, all <span class="hlt">park</span> areas within 350 meters (383 yards) of the centerline of the Exit <span class="hlt">Glacier</span> Road; (2) From Exit <span class="hlt">Glacier</span> Campground Entrance Road to the end of the main...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1318.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1318.pdf"><span>36 CFR 13.1318 - Location of the EGDA.</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>2014-07-01</p> <p>... NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-Kenai Fjords National <span class="hlt">Park</span> Exit <span class="hlt">Glacier</span> Developed... boundary to Exit <span class="hlt">Glacier</span> Campground Entrance Road, all <span class="hlt">park</span> areas within 350 meters (383 yards) of the centerline of the Exit <span class="hlt">Glacier</span> Road; (2) From Exit <span class="hlt">Glacier</span> Campground Entrance Road to the end of the main...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15..930L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15..930L"><span>Radio-echo sounding of Caucasus <span class="hlt">glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lavrentiev, Ivan; Kutuzov, Stanislav; Vasilenko, Evgeny; Macheret, Yuri</p> <p>2013-04-01</p> <p>Accurate <span class="hlt">glacier</span> volume and ice-thickness estimations are highly important for many glaciological applications. Recent <span class="hlt">glacier</span> reduction is affecting local river discharge and contributes to the global sea level rise. However, direct measurements of ice thickness are very sparse due to its high cost and laboriousness. One of the <span class="hlt">glacierized</span> mountain regions with a lack of direct ice-thickness measurements is Caucasus. So far data for several seismic and GPR profiles have been reported for only 3 <span class="hlt">glaciers</span> from more than 1.7 thousands located in Caucasus. In 2010-2012 a number of ground base and airborne radio-echo sounding surveys have been accomplished in Caucasus Mountains using 20 MHz monopulse radar VIRL-6. Special aerial version of this ground penetrating radar was designed for helicopter-born measurements. The radar has a relatively long (10 m) receiving and transmitting antennas, which together with receiving, recording and transmitting devices can be mounted on a special girder, being suspended from a helicopter. VIRL-6 radar is light weight and can be quickly transformed into ground version. Equipment has been used on 16 <span class="hlt">glaciers</span> including biggest <span class="hlt">glacier</span> in Caucasus - Bezengi (36 km2) most of which have a highly crevassed surfaces and heterogeneous internal structure. Independent data were obtained also for two <span class="hlt">glaciers</span> using ground version of the same VIRL-6 radar. In total more than 120 km of radar profiles were obtained. Results showed good agreement between ground and aerial measurements. Ice-thickness values exceeded 420 m for some of the Central Caucasus <span class="hlt">glaciers</span>. Successful use of VIRL-6 radar in Caucasus opens up the possibility of using such equipment on different types of <span class="hlt">glaciers</span> in polar and mountain regions, including temperate, polythermal and surging <span class="hlt">glaciers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://arc.lib.montana.edu/snow-science/item/460','USGSPUBS'); return false;" href="http://arc.lib.montana.edu/snow-science/item/460"><span>Characterizing wet slab and glide slab avalanche occurrence along the Going-to-the-Sun Road, <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Peitzsch, Erich H.; Hendrikx, Jordy; Fagre, Daniel B.; Reardon, Blase</p> <p>2010-01-01</p> <p>Wet slab and glide slab snow avalanches are dangerous and yet can be particularly difficult to predict. Both wet slab and glide slab avalanches are thought to depend upon free water moving through the snowpack but are driven by different processes. In <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, both types of avalanches can occur in the same year and affect the Going-to-the-Sun Road (GTSR). Both wet slab and glide slab avalanches along the GTSR from 2003-2010 are investigated. Meteorological data from two high-elevation weather stations and one SNOTEL site are used in conjunction with an avalanche database and snowpit profiles. These data were used to characterize years when only glide slab avalanches occurred and those years when both glide slab and wet slab avalanches occurred. Results of 168 glide slab and 57 wet slab avalanches along the GTSR suggest both types of avalanche occurrence depend on sustained warming periods with intense solar radiation (or rain on snow) to produce free water in the snowpack. Differences in temperature and net radiation metrics between wet slab and glide slab avalanches emerge as one moves from one day to seven days prior to avalanche occurrence. On average, a more rapid warming precedes wet slab avalanche occurrence. Glide slab and wet slab avalanches require a similar amount of net radiation. Wet slab avalanches do not occur every year, while glide slab avalanches occur annually. These results aim to enhance understanding of the required meteorological conditions for wet slab and glide slab avalanches and aid in improved wet snow avalanche forecasting.</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://images.nasa.gov/#/details-GSFC_20171208_Archive_e001874.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001874.html"><span>Malaspina <span class="hlt">Glacier</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>NASA image captured August 31, 2000 The tongue of the Malaspina <span class="hlt">Glacier</span>, the largest <span class="hlt">glacier</span> in Alaska, fills most of this image. The Malaspina lies west of Yakutat Bay and covers 1,500 sq. MI (3,880 sq. km). Credit: NASA/Landsat 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 Join us on Facebook</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3339.photos.194866p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3339.photos.194866p/"><span>3. VIEW TO WEST SHOWING <span class="hlt">EAST</span> END OF OIL HOUSE ...</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. VIEW TO WEST SHOWING <span class="hlt">EAST</span> END OF OIL HOUSE (LEFT) AND <span class="hlt">EAST</span> SIDE OF CRATING SHED AND ASSEMBLY BUILDING (RIGHT BACKGROUND). - Rosie the Riveter National Historical <span class="hlt">Park</span>, Ford Assembly Plant, 1400 Harbour Way South, Richmond, Contra Costa County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11543521','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11543521"><span>Sedimentology and geochemistry of a perennially ice-covered epishelf lake in Bunger Hills Oasis, <span class="hlt">East</span> Antarctica.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Doran, P T; Wharton, R A; Lyons, W B; Des Marais, D J; Andersen, D T</p> <p>2000-01-01</p> <p>A process-oriented study was carried out in White Smoke lake, Bunger Hills, <span class="hlt">East</span> Antarctica, a perennially ice-covered (1.8 to 2.8 m thick) epishelf (tidally-forced) lake. The lake water has a low conductivity and is relatively well mixed. Sediments are transferred from the adjacent <span class="hlt">glacier</span> to the lake when <span class="hlt">glacier</span> ice surrounding the sediment is sublimated at the surface and replaced by accumulating ice from below. The lake bottom at the west end of the lake is mostly rocky with a scant sediment cover. The <span class="hlt">east</span> end contains a thick sediment profile. Grain size and delta 13C increase with sediment depth, indicating a more proximal <span class="hlt">glacier</span> in the past. Sedimentary 210Pb and 137Cs signals are exceptionally strong, probably a result of the focusing effect of the large glacial catchment area. The post-bomb and pre-bomb radiocarbon reservoirs are c. 725 14C yr and c. 1950 14C yr, respectively. Radiocarbon dating indicates that the <span class="hlt">east</span> end of the lake is >3 ka BP, while photographic evidence and the absence of sediment cover indicate that the west end has formed only over the last century. Our results indicate that the southern ice edge of Bunger Hills has been relatively stable with only minor fluctuations (on the scale of hundreds of metres) over the last 3000 years.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17749022','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17749022"><span>Quantifying global warming from the retreat of <span class="hlt">glaciers</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Oerlemans, J</p> <p>1994-04-08</p> <p>Records of <span class="hlt">glacier</span> fluctuations compiled by the World <span class="hlt">Glacier</span> Monitoring Service can be used to derive an independent estimate of global warming during the last 100 years. Records of different <span class="hlt">glaciers</span> are made comparable by a two-step scaling procedure: one allowing for differences in <span class="hlt">glacier</span> geometry, the other for differences in climate sensitivity. The retreat of <span class="hlt">glaciers</span> during the last 100 years appears to be coherent over the globe. On the basis of modeling of the climate sensitivity of <span class="hlt">glaciers</span>, the observed <span class="hlt">glacier</span> retreat can be explained by a linear warming trend of 0.66 kelvin per century.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.H13L1589M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.H13L1589M"><span>Hydrochemical Signatures of <span class="hlt">Glacier</span> Melt and Groundwater Storage on Volcán Chimborazo, Ecuador</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McLaughlin, R.; Ng, G. H. C.; La Frenierre, J.; Wickert, A. D.; Baraer, M.</p> <p>2016-12-01</p> <p>With ever-growing water demands for hydroelectricity, agriculture, and domestic use, the accelerated retreat of tropical <span class="hlt">glaciers</span> is raising concerns about future water supply sustainability. In the tropical Andes, where precipitation is seasonal and spatially heterogeneous, <span class="hlt">glaciers</span> are particularly important as their storage and slow release of water helps to modulate stream discharge on daily to yearly time scales. Predicting the effect their shrinkage will have on water resources is not straightforward as little is known about the connections in these glaciated volcanic catchments between meltwater, groundwater, precipitation and surficial discharge. Here, stable isotope and major ion analyses inform a hydrochemical mixing model in order to identify water sources and their relative contributions to stream and spring discharge on Volcán Chimborazo, a stratovolcano located in the Ecuadorian Andes. Moisture in this region generally arrives from the Amazon basin to the <span class="hlt">east</span>, resulting in a steep northeast-southwest precipitation gradient that produces wet and dry sides of the mountain. Dry season water samples were collected on both sides from major streams and springs at varying elevations and distances from the <span class="hlt">glacier</span> tongues, along with samples of precipitation (when possible) and <span class="hlt">glacier</span> ice. Data on specific conductivity, pH, and temperature were collected in situ for each sample. The paired catchment study allows us to isolate a primarily glacial melt signature on the dry side and compare it to data on the wet side, where glacial melt and precipitation both contribute to groundwater and surface-water discharge.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17739514','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17739514"><span><span class="hlt">Glacier</span> Geophysics: Dynamic response of <span class="hlt">glaciers</span> to changing climate may shed light on processes in the earth's interior.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kamb, B</p> <p>1964-10-16</p> <p>From physical measurements on <span class="hlt">glaciers</span> and experimental studies of ice properties a framework of concept and theory is being built which bids fair to place <span class="hlt">glaciers</span> among the more quantitatively understandable phenomena in the earth sciences. Measurements of flow velocity, deformation and stress, ice thickness and channel configuration, temperature, internal structure of theice, mass and energy balance, and response to meteorological variables all contribute to this understanding, as do still other measurements hardly discussed here, such as electrical properties, radioactive age measurements, and detailed studies of chemical and isotopic composition. The obvious goals of this work-the interpretation of past and present <span class="hlt">glacier</span> fluctuations in terms of changes in world climate, and the prediction of <span class="hlt">glacier</span> behavior-remain elusive, even though a good conceptual groundwork has been laid for dealing with the more tractable aspects of these problems. Intriguing recent discoveries have been made about such matters as the way in which <span class="hlt">glaciers</span> react dynamically to changing conditions, the inter-relations between thermal regime and ice motion, the structural mechanisms of <span class="hlt">glacier</span> flow, and the changes produced in ice by flow. One can recognize in these developments the possibility that concepts derived from the study of <span class="hlt">glacier</span> flow may be applicable to phenomena of solid deformation deep in the earth. In this way <span class="hlt">glacier</span> geophysics may have a useful impact beyond the study of <span class="hlt">glaciers</span> themselves.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1912332S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1912332S"><span>Geodetic <span class="hlt">glacier</span> mass balancing on ice caps - inseparably connected to firn modelling?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saß, Björn L.; Sauter, Tobias; Seehaus, Thorsten; Braun, Matthias H.</p> <p>2017-04-01</p> <p>Observed melting of <span class="hlt">glaciers</span> and ice caps in the polar regions contribute to the ongoing global sea level rise (SLR). A rising sea level and its consequences are one of the major challenges for coastal societies in the next decades to centuries. Gaining knowledge about the main drivers of SLR and bringing it together is one recent key-challenge for environmental science. The high arctic Svalbard archipelago faced a strong climatic change in the last decades, associated with a change in the cryosphere. Vestfonna, a major Arctic ice cap in the north <span class="hlt">east</span> of Svalbard, harbors land and marine terminating <span class="hlt">glaciers</span>, which expose a variability of behavior. We use high resolution remote sensing data from space-borne radar (TanDEM-X, TerraSAR-X, Sentinel-1a), acquired between 2009 and 2015, to estimate <span class="hlt">glacier</span> velocity and high accurate surface elevation changes. For DEM registration we use space-borne laser altimetry (ICESat) and an existing in-situ data archive (IPY Kinnvika). In order to separate individual <span class="hlt">glacier</span> basin changes for a detailed mass balance study and for further SLR contribution estimates, we use <span class="hlt">glacier</span> outlines from the Global Land Ice Measurements from Space (GLIMS) project. Remaining challenges of space-borne observations are the reduction of measurement uncertainties, in the case of Synthetic Aperture Radar most notably signal penetration into the <span class="hlt">glacier</span> surface. Furthermore, in order to convert volume to mass change one has to use the density of the changed mass (conversion factor) and one has to account for the mass conservation processes in the firn package (firn compaction). Both, the conversion factor and the firn compaction are not (yet) measurable for extensive ice bodies. They have to be modelled by coupling point measurements and regional gridded climate data. Results indicate a slight interior thickening contrasted with wide spread thinning in the ablation zone of the marine terminating outlets. While one <span class="hlt">glacier</span> system draining to the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.8585V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.8585V"><span>Future streamflow droughts in <span class="hlt">glacierized</span> catchments: the impact of dynamic <span class="hlt">glacier</span> modelling and changing thresholds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Van Tiel, Marit; Van Loon, Anne; Wanders, Niko; Vis, Marc; Teuling, Ryan; Stahl, Kerstin</p> <p>2017-04-01</p> <p>In <span class="hlt">glacierized</span> catchments, snowpack and <span class="hlt">glaciers</span> function as an important storage of water and hydrographs of highly <span class="hlt">glacierized</span> catchments in mid- and high latitudes thus show a clear seasonality with low flows in winter and high flows in summer. Due to the ongoing climate change we expect this type of storage capacity to decrease with resultant consequences for the discharge regime. In this study we focus on streamflow droughts, here defined as below average water availability specifically in the high flow season, and which methods are most suitable to characterize future streamflow droughts as regimes change. Two <span class="hlt">glacierized</span> catchments, Nigardsbreen (Norway) and Wolverine (Alaska), are used as case study and streamflow droughts are compared between two periods, 1975-2004 and 2071-2100. Streamflow is simulated with the HBV light model, calibrated on observed discharge and seasonal <span class="hlt">glacier</span> mass balances, for two climate change scenarios (RCP 4.5 & RCP 8.5). In studies on future streamflow drought often the same variable threshold of the past has been applied to the future, but in regions where a regime shift is expected this method gives severe "droughts" in the historic high-flow period. We applied the new alternative transient variable threshold, a threshold that adapts to the changing hydrological regime and is thus better able to cope with this issue, but has never been thoroughly tested in <span class="hlt">glacierized</span> catchments. As the <span class="hlt">glacier</span> area representation in the hydrological modelling can also influence the modelled discharge and the derived streamflow droughts, we evaluated in this study both the difference between the historical variable threshold (HVT) and transient variable threshold (TVT) and two different <span class="hlt">glacier</span> area conceptualisations (constant area (C) and dynamical area (D)), resulting in four scenarios: HVT-C, HVT-D, TVT-C and TVT-D. Results show a drastic decrease in the number of droughts in the HVT-C scenario due to increased <span class="hlt">glacier</span> melt. The deficit</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1918586S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1918586S"><span>Ocean impact on Nioghalvfjerdsfjorden <span class="hlt">Glacier</span>, Northeast Greenland</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schaffer, Janin; Kanzow, Torsten; von Appen, Wilken-Jon; Mayer, Christoph</p> <p>2017-04-01</p> <p>The ocean plays an important role in modulating the mass balance of the Greenland Ice Sheet by delivering heat to the marine-terminating outlet <span class="hlt">glaciers</span> around Greenland. The largest of three outlet <span class="hlt">glaciers</span> draining the Northeast Greenland Ice Stream is Nioghalvfjerdsfjorden <span class="hlt">Glacier</span> (also referred to as 79 North <span class="hlt">Glacier</span>). Historic observations showed that warm waters of Atlantic origin are present in the subglacial cavity below the 80 km long floating ice tongue of the Nioghalvfjerdsfjorden <span class="hlt">Glacier</span> and cause strong basal melt at the grounding line, but to date it has been unknown how those warm water enter the cavity. In order to understand how Atlantic origin waters carry heat into the subglacial cavity beneath Nioghalvfjerdsfjorden <span class="hlt">Glacier</span>, we performed bathymetric, hydrographic, and velocity observations in the vicinity of the main <span class="hlt">glacier</span> calving front aboard RV Polarstern in summer 2016. The bathymetric multibeam data shows a 500 m deep and 2 km narrow passage downstream of a 310 m deep sill. This turned out to be the only location deep enough for an exchange of Atlantic waters between the <span class="hlt">glacier</span> cavity and the continental shelf. Hydrographic and velocity measurements revealed a density driven plume in the vicinity of the <span class="hlt">glacier</span> calving front causing a rapid flow of waters of Atlantic origin warmer 1°C into the subglacial cavity through the 500 m deep passage. In addition, glacially modified waters flow out of the <span class="hlt">glacier</span> cavity below the 80 m deep ice base. In the vicinity of the <span class="hlt">glacier</span>, the glacially modified waters form a distinct mixed layer situated above the Atlantic waters and below the ambient Polar water. At greater distances from the <span class="hlt">glacier</span> this layer is eroded by lateral mixing with ambient water. Based on our observations we will present an estimate of the ocean heat transport into the subglacial cavity. In comparison with historic observations we find an increase in Atlantic water temperatures throughout the last 20 years. The resulting</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.C21B0597A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.C21B0597A"><span>Bathymetry of Patagonia <span class="hlt">glacier</span> fjords and <span class="hlt">glacier</span> ice thickness from high-resolution airborne gravity combined with other data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>An, L.; Rignot, E.; Rivera, A.; Bunetta, M.</p> <p>2012-12-01</p> <p>The North and South Patagonia Ice fields are the largest ice masses outside Antarctica in the Southern Hemisphere. During the period 1995-2000, these <span class="hlt">glaciers</span> lost ice at a rate equivalent to a sea level rise of 0.105 ± 0.001 mm/yr. In more recent years, the <span class="hlt">glaciers</span> have been thinning more quickly than can be explained by warmer air temperatures and decreased precipitation. A possible cause is an increase in flow speed due to enhanced ablation of the submerged <span class="hlt">glacier</span> fronts. To understand the dynamics of these <span class="hlt">glaciers</span> and how they change with time, it is critical to have a detailed view of their ice thickness, the depth of the <span class="hlt">glacier</span> bed below sea or lake level, how far inland these <span class="hlt">glaciers</span> remain below sea or lake level, and whether bumps or hollows in the bed may slow down or accelerate their retreat. A grid of free-air gravity data over the Patagonia <span class="hlt">Glaciers</span> was collected in May 2012 and October 2012, funded by the Gordon and Betty Moore Foundation (GBMF) to measure ice thickness and sea floor bathymetry. This survey combines the Sander Geophysics Limited (SGL) AIRGrav system, SGL laser altimetry and Chilean CECS/UCI ANDREA-2 radar. To obtain high-resolution and high-precision gravity data, the helicopter operates at 50 knots (25.7 m/s) with a grid spacing of 400m and collects gravity data at sub mGal level (1 Gal =1 Galileo = 1 cm/s2) near <span class="hlt">glacier</span> fronts. We use data from the May 2012 survey to derive preliminarily high-resolution, high-precision thickness estimates and bathymetry maps of Jorge Montt <span class="hlt">Glacier</span> and San Rafael <span class="hlt">Glacier</span>. Boat bathymetry data is used to optimize the inversion of gravity over water and radar-derived thickness over <span class="hlt">glacier</span> ice. The bathymetry maps will provide a breakthrough in our knowledge of the ice fields and enable a new era of <span class="hlt">glacier</span> modeling and understanding that is not possible at present because ice thickness is not known.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1810891N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1810891N"><span>Internationally coordinated <span class="hlt">glacier</span> monitoring - a timeline since 1894</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nussbaumer, Samuel U.; Armstrong, Richard; Fetterer, Florence; Gärtner-Roer, Isabelle; Hoelzle, Martin; Machguth, Horst; Mölg, Nico; Paul, Frank; Raup, Bruce H.; Zemp, Michael</p> <p>2016-04-01</p> <p>Changes in <span class="hlt">glaciers</span> and ice caps provide some of the clearest evidence of climate change, with impacts on sea-level variations, regional hydrological cycles, and natural hazard situations. Therefore, <span class="hlt">glaciers</span> have been recognized as an Essential Climate Variable (ECV). Internationally coordinated collection and distribution of standardized information about the state and change of <span class="hlt">glaciers</span> and ice caps was initiated in 1894 and is today organized within the Global Terrestrial Network for <span class="hlt">Glaciers</span> (GTN-G). GTN-G ensures the continuous development and adaptation of the international strategies to the long-term needs of users in science and policy. A GTN-G Steering Committee coordinates, supports and advices the operational bodies responsible for the international <span class="hlt">glacier</span> monitoring, which are the World <span class="hlt">Glacier</span> Monitoring Service (WGMS), the US National Snow and Ice Data Center (NSIDC), and the Global Land Ice Measurements from Space (GLIMS) initiative. In this presentation, we trace the development of the internationally coordinated <span class="hlt">glacier</span> monitoring since its beginning in the 19th century. Today, several online databases containing a wealth of diverse data types with different levels of detail and global coverage provide fast access to continuously updated information on <span class="hlt">glacier</span> fluctuation and inventory data. All <span class="hlt">glacier</span> datasets are made freely available through the respective operational bodies within GTN-G, and can be accessed through the GTN-G Global <span class="hlt">Glacier</span> Browser (http://www.gtn-g.org/data_browser.html). <span class="hlt">Glacier</span> inventory data (e.g., digital outlines) are available for about 180,000 <span class="hlt">glaciers</span> (GLIMS database, RGI - Randolph <span class="hlt">Glacier</span> Inventory, WGI - World <span class="hlt">Glacier</span> Inventory). <span class="hlt">Glacier</span> front variations with about 45,000 entries since the 17th century and about 6,200 glaciological and geodetic mass (volume) change observations dating back to the 19th century are available in the Fluctuations of <span class="hlt">Glaciers</span> (FoG) database. These datasets reveal clear evidence that</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1052939-contribution-glacier-melt-streamflow','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1052939-contribution-glacier-melt-streamflow"><span>The contribution of <span class="hlt">glacier</span> melt to streamflow</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>Schaner, Neil; Voisin, Nathalie; Nijssen, Bart</p> <p>2012-09-13</p> <p>Ongoing and projected future changes in <span class="hlt">glacier</span> extent and water storage globally have lead to concerns about the implications for water supplies. However, the current magnitude of <span class="hlt">glacier</span> contributions to river runoff is not well known, nor is the population at risk to future <span class="hlt">glacier</span> changes. We estimate an upper bound on <span class="hlt">glacier</span> melt contribution to seasonal streamflow by computing the energy balance of <span class="hlt">glaciers</span> globally. Melt water quantities are computed as a fraction of total streamflow simulated using a hydrology model and the melt fraction is tracked down the stream network. In general, our estimates of the <span class="hlt">glacier</span> meltmore » contribution to streamflow are lower than previously published values. Nonetheless, we find that globally an estimated 225 (36) million people live in river basins where maximum seasonal <span class="hlt">glacier</span> melt contributes at least 10% (25%) of streamflow, mostly in the High Asia region.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMGC21A0852S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMGC21A0852S"><span><span class="hlt">Glacier</span> Sensitivity Across the Andes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sagredo, E. A.; Lowell, T. V.; Rupper, S.</p> <p>2010-12-01</p> <p>Most of the research on causes driving former glacial fluctuations, and the climatic signals involved, has focused on the comparisons of sequences of glacial events in separate regions of the world and their temporal-phasing relationship with terrestrial or extraterrestrial climate-forcing mechanisms. Nevertheless the climatic signals related with these glacial advances are still under debate. This impossibility to resolve these questions satisfactorily have been generally attributed to the insufficiently precise chronologies and unevenly distributed records. However, behind these ideas lies the implicit assumption that <span class="hlt">glaciers</span> situated in different climate regimes respond uniformly to similar climatic perturbations. This ongoing research is aimed to explore the climate-<span class="hlt">glacier</span> relationship at regional scale, through the analysis of the spatial variability of <span class="hlt">glacier</span> sensitivity to climatic change. By applying a Surface Energy Mass Balance model (SEMB) developed by Rupper and Roe (2008) to <span class="hlt">glaciers</span> located in different climatic regimes, we analyzed the spatial variability of mass balance changes under different baseline conditions and under different scenarios of climatic change. For the sake of this research, the analysis is being focused on the Andes, which in its 9,000 km along the western margin of South America offers an unparalleled climatic diversity. Preliminary results suggest that above some threshold of climate change (a hypothetical uniform perturbation), all the <span class="hlt">glaciers</span> across the Andes would respond in the “same direction” (advancing or retreating). Below that threshold, <span class="hlt">glaciers</span> located in some climatic regimes may be insensitive to the specific perturbation. On the other hand, <span class="hlt">glaciers</span> located in different climatic regimes may exhibit a “different magnitude” of change under a uniform climatic perturbation. Thus, <span class="hlt">glaciers</span> located in the dry Andes of Perú, Chile and Argentina are more sensitive to precipitation changes than variations in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.4807P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.4807P"><span><span class="hlt">Glacier</span>-specific elevation changes in western Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Paul, Frank; Le Bris, Raymond</p> <p>2013-04-01</p> <p>Deriving <span class="hlt">glacier</span>-specific elevation changes from DEM differencing and digital <span class="hlt">glacier</span> outlines is rather straight-forward if the required datasets are available. Calculating such changes over large regions and including <span class="hlt">glaciers</span> selected for mass balance measurements in the field, provides a possibility to determine the representativeness of the changes observed at these <span class="hlt">glaciers</span> for the entire region. The related comparison of DEM-derived values for these <span class="hlt">glaciers</span> with the overall mean avoids the rather error-prone conversion of volume to mass changes (e.g. due to unknown densities) and gives unit-less correction factors for upscaling the field measurements to a larger region. However, several issues have to be carefully considered, such as proper co-registration of the two DEMs, date and accuracy of the datasets compared, as well as source data used for DEM creation and potential artefacts (e.g. voids). In this contribution we present an assessment of the representativeness of the two mass balance <span class="hlt">glaciers</span> Gulkana and Wolverine for the overall changes of nearly 3200 <span class="hlt">glaciers</span> in western Alaska over a ca. 50-year time period. We use an elevation change dataset from a study by Berthier et al. (2010) that was derived from the USGS DEM of the 1960s (NED) and a more recent DEM derived from SPOT5 data for the SPIRIT project. Additionally, the ASTER GDEM was used as a more recent DEM. Historic <span class="hlt">glacier</span> outlines were taken from the USGS digital line graph (DLG) dataset, corrected with the digital raster graph (DRG) maps from USGS. Mean <span class="hlt">glacier</span> specific elevation changes were derived based on drainage divides from a recently created inventory. Land-terminating, lake-calving and tidewater <span class="hlt">glaciers</span> were marked in the attribute table to determine their changes separately. We also investigated the impact of handling potential DEM artifacts in three different ways and compared elevation changes with altitude. The mean elevation changes of Gulkana and Wolverine <span class="hlt">glaciers</span> (about -0</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/nj1293.photos.038137p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/nj1293.photos.038137p/"><span>2. COTTAGES, NORTH SIDE OF OCEAN PATHWAY <span class="hlt">EAST</span> OF BEACH ...</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. COTTAGES, NORTH SIDE OF OCEAN PATHWAY <span class="hlt">EAST</span> OF BEACH AVENUE, (NOS. 17, 15, 13, 11, 7 AND 5), GENERAL VIEW LOOKING NORTH - Town of Ocean Grove, <span class="hlt">East</span> terminus of State Route 33, south of Asbury <span class="hlt">Park</span>, Ocean Grove, Monmouth County, NJ</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/949959','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/949959"><span>Fiscal Year 2008 Phased Construction Completion Report for EU Z2-33 in Zone 2, <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee</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>Bechtel Jacobs</p> <p></p> <p>The Record of Decision for Soil, Buried Waste, and Subsurface Structure Actions in Zone 2, <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee (DOE/OR/01-2161&D2) (Zone 2 ROD) acknowledged that most of the 800 acres in Zone 2 were contaminated, but that sufficient data to confirm the levels of contamination were lacking. The Zone 2 ROD further specified that a sampling strategy for filling the data gaps would be developed. The Remedial Design Report/Remedial Action Work Plan for Zone 2 Soils, Slabs, and Subsurface Structures, <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee (DOE/OR/01-2224&D3) (Zone 2 RDR/RAWP) defined the sampling strategy as themore » Dynamic Verification Strategy (DVS), generally following the approach used for characterization of the Zone 1 exposure units (EUs). The Zone 2 ROD divided the Zone 2 area into seven geographic areas and 44 EUs. To facilitate the data quality objectives (DQOs) of the DVS process, the Zone 2 RDR/RAWP regrouped the 44 EUs into 12 DQO scoping EU groups. These groups facilitated the DQO process by placing similar facilities and their support facilities together and allowing identification of data gaps. The EU groups were no longer pertinent after DQO planning was completed and characterization was conducted as areas became accessible. As the opportunity to complete characterization became available, the planned DVS program and remedial actions (RAs) were completed for EU Z2-33. Remedial action was also performed at two additional areas in adjacent EU Z2-42 because of their close proximity and similar nature to a small surface soil RA in EU Z2-33. Remedial actions for building slabs performed in EU Z2-33 during fiscal year (FY) 2007 were reported in the Fiscal Year 2007 Phased Construction Completion Report for the Zone 2 Soils, Slabs, and Subsurface Structures at <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee (DOE/OR/01-2723&D1). Recommended RAs for EU Z2-42 were described in the Fiscal Year 2006 Phased</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003EAEJA.....2616O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003EAEJA.....2616O"><span>Glaciochemical investigation of an ice core from Belukha <span class="hlt">Glacier</span>,Siberian Altai</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Olivier, S.; Schwikowski, S.; Gäggeler, H. W.; Lüthi, M.; Eyrik, S.; Blaser, C.; Saurer, M.; Schotterer, U.</p> <p>2003-04-01</p> <p>Little is known about climatic change and paleo-atmospheric composition in Siberia. The Altai is the only alpine region in this area covered by <span class="hlt">glaciers</span> that might serve as archives for such studies. Moreover, it is located close to air pollution sources in <span class="hlt">East</span> Kazakhstan and South Siberia (heavy metal mining, metallurgy) as well as to the nuclear test site of Semipalatinsk (release of radionuclides into the atmosphere). In order to reconstruct air pollution levels in the Altai region, a 140-meter ice core down to bedrock was recovered from the Belukha <span class="hlt">glacier</span> (N49^o48'26", E86^o34'43", 4062 m asl) in July 2001. This site was selected based on the results of an exploratory study conducted in 2000. So far, the concentrations of major ionic species and the stable isotope ratio δ18O were determined in the approx. 90 topmost meters of the ice core covering about 200 years. Dating of the upper part of the ice core was performed by a combination of methods that include e.g. nuclear techniques and annual-layer counting. The annual net accumulation amounts to about 0.53 m weq. and indicates that snow at the Belukha <span class="hlt">glacier</span> might be partly eroded by wind, a situation that is often observed for a <span class="hlt">glacier</span> saddle. The borehole temperature (-16 ^oC at 80 m depth), the discernible fluctuations of the stable isotope and chemistry records as well as the linearity of the decrease of the log. 210Pb activities with depth indicate that the glaciochemical record is well preserved and not significantly altered by melting processes. In pre-industrial ice concentrations of carboxylic acids and ammonium are high, suggesting the surrounding forest as source of biogenic emissions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ACP....18.6441N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ACP....18.6441N"><span>Seasonal variation and light absorption property of carbonaceous aerosol in a typical <span class="hlt">glacier</span> region of the southeastern Tibetan Plateau</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Niu, Hewen; Kang, Shichang; Wang, Hailong; Zhang, Rudong; Lu, Xixi; Qian, Yun; Paudyal, Rukumesh; Wang, Shijin; Shi, Xiaofei; Yan, Xingguo</p> <p>2018-05-01</p> <p>Deposition and accumulation of light-absorbing carbonaceous aerosol on <span class="hlt">glacier</span> surfaces can alter the energy balance of <span class="hlt">glaciers</span>. In this study, 2 years (December 2014 to December 2016) of continuous observations of carbonaceous aerosols in the <span class="hlt">glacierized</span> region of the Mt. Yulong and Ganhaizi (GHZ) basin are analyzed. The average elemental carbon (EC) and organic carbon (OC) concentrations were 1.51±0.93 and 2.57±1.32 µg m-3, respectively. Although the annual mean OC / EC ratio was 2.45±1.96, monthly mean EC concentrations during the post-monsoon season were even higher than OC in the high altitudes (approximately 5000 m a. s. l. ) of Mt. Yulong. Strong photochemical reactions and local tourism activities were likely the main factors inducing high OC / EC ratios in the Mt. Yulong region during the monsoon season. The mean mass absorption efficiency (MAE) of EC, measured for the first time in Mt. Yulong, at 632 nm with a thermal-optical carbon analyzer using the filter-based method, was 6.82±0.73 m2 g-1, comparable with the results from other studies. Strong seasonal and spatial variations of EC MAE were largely related to the OC abundance. Source attribution analysis using a global aerosol-climate model, equipped with a black carbon (BC) source tagging technique, suggests that <span class="hlt">East</span> Asia emissions, including local sources, have the dominant contribution (over 50 %) to annual mean near-surface BC in the Mt. Yulong area. There is also a strong seasonal variation in the regional source apportionment. South Asia has the largest contribution to near-surface BC during the pre-monsoon season, while <span class="hlt">East</span> Asia dominates the monsoon season and post-monsoon season. Results in this study have great implications for accurately evaluating the influences of carbonaceous matter on glacial melting and water resource supply in <span class="hlt">glacierization</span> areas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3721114','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3721114"><span>Microbial biodiversity in <span class="hlt">glacier</span>-fed streams</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wilhelm, Linda; Singer, Gabriel A; Fasching, Christina; Battin, Tom J; Besemer, Katharina</p> <p>2013-01-01</p> <p>While <span class="hlt">glaciers</span> become increasingly recognised as a habitat for diverse and active microbial communities, effects of their climate change-induced retreat on the microbial ecology of <span class="hlt">glacier</span>-fed streams remain elusive. Understanding the effect of climate change on microorganisms in these ecosystems is crucial given that microbial biofilms control numerous stream ecosystem processes with potential implications for downstream biodiversity and biogeochemistry. Here, using a space-for-time substitution approach across 26 Alpine <span class="hlt">glaciers</span>, we show how microbial community composition and diversity, based on 454-pyrosequencing of the 16S rRNA gene, in biofilms of <span class="hlt">glacier</span>-fed streams may change as <span class="hlt">glaciers</span> recede. Variations in streamwater geochemistry correlated with biofilm community composition, even at the phylum level. The most dominant phyla detected in glacial habitats were Proteobacteria, Bacteroidetes, Actinobacteria and Cyanobacteria/chloroplasts. Microorganisms from ice had the lowest α diversity and contributed marginally to biofilm and streamwater community composition. Rather, streamwater apparently collected microorganisms from various glacial and non-glacial sources forming the upstream metacommunity, thereby achieving the highest α diversity. Biofilms in the <span class="hlt">glacier</span>-fed streams had intermediate α diversity and species sorting by local environmental conditions likely shaped their community composition. α diversity of streamwater and biofilm communities decreased with elevation, possibly reflecting less diverse sources of microorganisms upstream in the catchment. In contrast, β diversity of biofilms decreased with increasing streamwater temperature, suggesting that <span class="hlt">glacier</span> retreat may contribute to the homogenisation of microbial communities among <span class="hlt">glacier</span>-fed streams. PMID:23486246</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002AGUSM.U22A..06K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002AGUSM.U22A..06K"><span><span class="hlt">Glaciers</span> in 21st Century Himalayan Geopolitics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kargel, J. S.; Wessels, R.; Kieffer, H. H.</p> <p>2002-05-01</p> <p><span class="hlt">Glaciers</span> are ablating rapidly the world over. Nowhere are the rates of retreat and downwasting greater than in the Hindu Kush-Himalaya (HKH) region. It is estimated that over the next century, 40,000 square kilometers of present <span class="hlt">glacier</span> area in the HKH region will become ice free. Most of this area is in major valleys and the lowest glaciated mountain passes. The existence and characteristics of <span class="hlt">glaciers</span> have security impacts, and rapidly changing HKH <span class="hlt">glaciers</span> have broad strategic implications: (1) <span class="hlt">Glaciers</span> supply much of the fresh water and hydroelectric power in South and Central Asia, and so <span class="hlt">glaciers</span> are valuable resources. (2) Shared economic interests in water, hydroelectricity, flood hazards, and habitat preservation are a force for common cause and reasoned international relations. (3) <span class="hlt">Glaciers</span> and their high mountains generally pose a natural barrier tending to isolate people. Historically, they have hindered trade and intercultural exchanges and have protected against aggression. This has further promoted an independent spirit of the region's many ethnic groups. (4) Although <span class="hlt">glaciers</span> are generally incompatible with human development and habitation, many of the HKH region's <span class="hlt">glaciers</span> and their mountains have become sanctuaries and transit routes for militants. Siachen <span class="hlt">Glacier</span> in Kashmir has for 17 years been "the world's highest battlefield," with tens of thousands of troops deployed on both sides of the India/Pakistan line of control. In 1999, that conflict threatened to trigger all-out warfare, and perhaps nuclear warfare. Other recent terrorist and military action has taken place on <span class="hlt">glaciers</span> in Kyrgyzstan and Tajikistan. As terrorists are forced from easily controlled territories, many may tend to migrate toward the highest ground, where definitive encounters may take place in severe alpine glacial environments. This should be a major concern in Nepali security planning, where an Army offensive is attempting to reign in an increasingly robust and brutal</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('http://adsabs.harvard.edu/abs/2017EGUGA..1915896D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1915896D"><span>A fjord-<span class="hlt">glacier</span> coupled system model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>de Andrés, Eva; Otero, Jaime; Navarro, Francisco; Prominska, Agnieszka; Lapazaran, Javier; Walczowski, Waldemar</p> <p>2017-04-01</p> <p>With the aim of studying the processes occurring at the front of marine-terminating <span class="hlt">glaciers</span>, we couple a fjord circulation model with a flowline <span class="hlt">glacier</span> dynamics model, with subglacial discharge and calving, which allows the calculation of submarine melt and its influence on calving processes. For ocean modelling, we use a general circulation model, MITgcm, to simulate water circulation driven by both fjord conditions and subglacial discharge, and for calculating submarine melt rates at the <span class="hlt">glacier</span> front. To constrain freshwater input to the fjord, we use estimations from European Arctic Reanalysis (EAR). To determine the optimal values for each run period, we perform a sensitivity analysis of the model to subglacial discharge variability, aimed to get the best fit of model results to observed temperature and salinity profiles in the fjord for each of these periods. Then, we establish initial and boundary fjord conditions, which we vary weekly-fortnightly, and calculate the submarine melt rate as a function of depth at the calving front. These data are entered into the <span class="hlt">glacier</span>-flow model, Elmer/Ice, which has been added a crevasse-depth calving model, to estimate the <span class="hlt">glacier</span> terminus position at a weekly time resolution. We focus our study on the Hansbreen <span class="hlt">Glacier</span>-Hansbukta Fjord system, in Southern Spitsbergen, Svalbard, where a large set of data are available for both <span class="hlt">glacier</span> and fjord. The bathymetry of the entire system has been determined from ground penetrating radar and sonar data. In the fjord we have got temperature and salinity data from CTDs (May to September, 2010-2014) and from a mooring (September to May, 2011-2012). For Hansbreen, we use <span class="hlt">glacier</span> surface topography data from the SPIRIT DEM, surface mass balance from EAR, centre line <span class="hlt">glacier</span> velocities from stake measurements (May 2005-April 2011), weekly terminus positions from time-lapse photos (Sept. 2009-Sept. 2011), and sea-ice concentrations from time-lapse photos and Nimbus-7 SMMR and DMSP SSM</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.C22A..02N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.C22A..02N"><span>Response of major Greenland outlet <span class="hlt">glaciers</span> to oceanic and atmospheric forcing: Results from numerical modeling on Petermann, Jakobshavn and Helheim <span class="hlt">Glacier</span>.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nick, F. M.; Vieli, A.; Pattyn, F.; Van de Wal, R.</p> <p>2011-12-01</p> <p>Oceanic forcing has been suggested as a major trigger for dynamic changes of Greenland outlet <span class="hlt">glaciers</span>. Significant melting near their calving front or beneath the floating tongue and reduced support from sea ice or ice melange in front of their calving front can result in retreat of the terminus or the grounding line, and an increase in calving activities. Depending on the geometry and basal topography of the <span class="hlt">glacier</span>, these oceanic forcing can affect the <span class="hlt">glacier</span> dynamic differently. Here, we carry out a comparison study between three major outlet <span class="hlt">glaciers</span> in Greenland and investigate the impact of a warmer ocean on <span class="hlt">glacier</span> dynamics and ice discharge. We present results from a numerical ice-flow model applied to Petermann <span class="hlt">Glacier</span> in the north, Jakobshavn <span class="hlt">Glacier</span> in the west, and Helheim <span class="hlt">Glacier</span> in the southeast of Greenland.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001482.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001482.html"><span><span class="hlt">Glaciers</span> and Sea Level Rise</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>The Aletsch <span class="hlt">Glacier</span> in Switzerland is the largest valley <span class="hlt">glacier</span> in the Alps. Its volume loss since the middle of the 19th century is well-visible from the trimlines to the right of the image. To learn about the contributions of <span class="hlt">glaciers</span> to sea level rise, visit: www.nasa.gov/topics/earth/features/<span class="hlt">glacier</span>-sea-rise.html Credit: Frank Paul, University of Zurich 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1919373E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1919373E"><span>Tracer-based identification of rock <span class="hlt">glacier</span> thawing in a <span class="hlt">glacierized</span> Alpine catchment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Engel, Michael; Penna, Daniele; Tirler, Werner; Comiti, Francesco</p> <p>2017-04-01</p> <p>Current warming in high mountains leads to increased melting of snow, <span class="hlt">glacier</span> ice and permafrost. In particular rock <span class="hlt">glaciers</span>, as a creeping form of mountain permafrost, may release contaminants such as heavy metals into the stream during intense melting periods in summer. This may have strong impacts on both water quantity and quality of fresh water resources but might also harm the aquatic fauna in mountain regions. In this context, the present study used stable isotopes of water and electrical conductivity (EC) combined with trace, major and minor elements to identify the influence of permafrost thawing on the water quality in the <span class="hlt">glacierized</span> Solda catchment (130 km2) in South Tyrol (Italy). We carried out a monthly sampling of two springs fed by an active rock <span class="hlt">glacier</span> at about 2600 m a.s.l. from July to October 2015. Furthermore, we took monthly water samples from different stream sections of the Solda River (1110 to m a.s.l.) from March to November 2015. Meteorological data were measured by an Automatic Weather Station at 2825 m a.s.l. of the Hydrographic Office (Autonomous Province of Bozen-Bolzano). First results show that water from the rock <span class="hlt">glacier</span> springs and stream water fell along the global meteoric water line. Spring water was slightly more variable in isotopic ratio (δ2H: -91 to - 105 ) and less variable in dissolved solutes (EC: 380 to 611 μS/cm) than stream water (δ2H: -96 to - 107 ‰ and EC: 212 to 927 μS/cm). Both spring water and stream water showed a pronounced drop in EC during July and August, very likely induced by increased melt water dilution. In both water types, element concentrations of Ca and Mg were highest (up to 160 and 20 mg/l, respectively). In September, spring water showed higher concentrations in Cu, As, and Pb than stream water, indicating that these elements partly exceeded the concentration limit for drinking water. These observations highlight the important control, which rock <span class="hlt">glacier</span> thawing may have on water quality</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/imap/2108/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/imap/2108/report.pdf"><span>Geologic map of Bryce Canyon National <span class="hlt">Park</span> and vicinity, southwestern Utah</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Bowers, William E.</p> <p>1991-01-01</p> <p>Bryce Canyon National <span class="hlt">Park</span> is located along the eastern escarpment of the Paunsaugunt Plateau, which along with the Markagunt Plateau to the west, form the southernmost of the High Plateaus of Utah. The park’s unique scenery has been created by forces of differential erosion acting on colorful rocks exposed along and below the rim of the plateau. <span class="hlt">Park</span> headquarters and major scenic viewpoints that lie on or near the rim of the plateau are accessible from Utah Highway 12 mi west of the <span class="hlt">park</span>. More remote parts of the <span class="hlt">park</span> are located in canyons beneath the rim and are accessible only by foot, along horse trails or from a few unimproved dirt roads that approach the <span class="hlt">park</span> boundary from the <span class="hlt">east</span> or south.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016CG.....94...68S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016CG.....94...68S"><span>A GRASS GIS module to obtain an estimation of <span class="hlt">glacier</span> behavior under climate change: A pilot study on Italian <span class="hlt">glacier</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Strigaro, Daniele; Moretti, Massimiliano; Mattavelli, Matteo; Frigerio, Ivan; Amicis, Mattia De; Maggi, Valter</p> <p>2016-09-01</p> <p>The aim of this work is to integrate the Minimal <span class="hlt">Glacier</span> Model in a Geographic Information System Python module in order to obtain spatial simulations of <span class="hlt">glacier</span> retreat and to assess the future scenarios with a spatial representation. The Minimal <span class="hlt">Glacier</span> Models are a simple yet effective way of estimating <span class="hlt">glacier</span> response to climate fluctuations. This module can be useful for the scientific and glaciological community in order to evaluate <span class="hlt">glacier</span> behavior, driven by climate forcing. The module, called r.glacio.model, is developed in a GRASS GIS (GRASS Development Team, 2016) environment using Python programming language combined with different libraries as GDAL, OGR, CSV, math, etc. The module is applied and validated on the Rutor <span class="hlt">glacier</span>, a <span class="hlt">glacier</span> in the south-western region of the Italian Alps. This <span class="hlt">glacier</span> is very large in size and features rather regular and lively dynamics. The simulation is calibrated by reconstructing the 3-dimensional dynamics flow line and analyzing the difference between the simulated flow line length variations and the observed <span class="hlt">glacier</span> fronts coming from ortophotos and DEMs. These simulations are driven by the past mass balance record. Afterwards, the future assessment is estimated by using climatic drivers provided by a set of General Circulation Models participating in the Climate Model Inter-comparison Project 5 effort. The approach devised in r.glacio.model can be applied to most alpine <span class="hlt">glaciers</span> to obtain a first-order spatial representation of <span class="hlt">glacier</span> behavior under climate change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..1710743B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..1710743B"><span>Geomorphology and Ice Content of <span class="hlt">Glacier</span> - Rock <span class="hlt">Glacier</span> &ndash; Moraine Complexes in Ak-Shiirak Range (Inner Tien Shan, Kyrgyzstan)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bolch, Tobias; Kutuzov, Stanislav; Rohrbach, Nico; Fischer, Andrea; Osmonov, Azamat</p> <p>2015-04-01</p> <p>Meltwater originating from the Tien Shan is of high importance for the runoff to the arid and semi-arid region of Central Asia. Previous studies estimate a <span class="hlt">glaciers</span>' contribution of about 40% for the Aksu-Tarim Catchment, a transboundary watershed between Kyrgyzstan and China. Large parts of the Ak-Shiirak Range drain into this watershed. <span class="hlt">Glaciers</span> in Central and Inner Tien Shan are typically polythermal or even cold and surrounded by permafrost. Several <span class="hlt">glaciers</span> terminate into large moraine complexes which show geomorphological indicators of ice content such as thermo-karst like depressions, and further downvalley signs of creep such as ridges and furrows and a fresh, steep rock front which are typical indicators for permafrost creep ("rock <span class="hlt">glacier</span>"). Hence, <span class="hlt">glaciers</span> and permafrost co-exist in this region and their interactions are important to consider, e.g. for the understanding of glacial and periglacial processes. It can also be assumed that the ice stored in these relatively large dead-ice/moraine-complexes is a significant amount of the total ice storage. However, no detailed investigations exist so far. In an initial study, we investigated the structure and ice content of two typical <span class="hlt">glacier</span>-moraine complexes in the Ak-Shiirak-Range using different ground penetrating radar (GPR) devices. In addition, the geomorphology was mapped using high resolution satellite imagery. The structure of the moraine-rock <span class="hlt">glacier</span> complex is in general heterogeneous. Several dead ice bodies with different thicknesses and moraine-derived rock <span class="hlt">glaciers</span> with different stages of activities could be identified. Few parts of these "rock <span class="hlt">glaciers</span>" contain also massive ice but the largest parts are likely characterised by rock-ice layers of different thickness and ice contents. In one <span class="hlt">glacier</span> forefield, the thickness of the rock-ice mixture is partly more than 300 m. This is only slightly lower than the maximum thickness of the <span class="hlt">glacier</span> ice. Our measurements revealed that up to 20% of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Geomo.311....1A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Geomo.311....1A"><span>Debris thickness patterns on debris-covered <span class="hlt">glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Anderson, Leif S.; Anderson, Robert S.</p> <p>2018-06-01</p> <p>Many debris-covered <span class="hlt">glaciers</span> have broadly similar debris thickness patterns: surface debris thickens and tends to transition from convex- to concave-up-down <span class="hlt">glacier</span>. We explain this pattern using theory (analytical and numerical models) paired with empirical observations. Down <span class="hlt">glacier</span> debris thickening results from the conveyor-belt-like nature of the <span class="hlt">glacier</span> surface in the ablation zone (debris can typically only be added but not removed) and from the inevitable decline in ice surface velocity toward the terminus. Down-<span class="hlt">glacier</span> thickening of debris leads to the reduction of sub-debris melt and debris emergence toward the terminus. Convex-up debris thickness patterns occur near the up-<span class="hlt">glacier</span> end of debris covers where debris emergence dominates (ablation controlled). Concave-up debris thickness patterns occur toward <span class="hlt">glacier</span> termini where declining surface velocities dominate (velocity controlled). A convex-concave debris thickness profile inevitably results from the transition between ablation-control and velocity-control down-<span class="hlt">glacier</span>. Debris thickness patterns deviating from this longitudinal shape are most likely caused by changes in hillslope debris supply through time. By establishing this expected debris thickness pattern, the effects of climate change on debris cover can be better identified.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19960000908','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19960000908"><span>SAR investigations of <span class="hlt">glaciers</span> in northwestern North America</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lingle, Craig S.; Harrison, William D.</p> <p>1995-01-01</p> <p>The objective of this project was to investigate the utility of satellite synthetic aperture radar (SAR) imagery for measurement of geophysical parameters on Alaskan <span class="hlt">glaciers</span> relevant to their mass balance and dynamics, including: (1) the positions of firn lines (late-summer snow lines); (2) surface velocities on fast-flowing (surging) <span class="hlt">glaciers</span>, and also on slower steady-flow <span class="hlt">glaciers</span>; and (3) the positions and changes in the positions of <span class="hlt">glacier</span> termini. Preliminary studies of topography and <span class="hlt">glacier</span> surface velocity with SAR interferometry have also been carried out. This project was motivated by the relationships of multi-year to decadal changes in <span class="hlt">glacier</span> geometry to changing climate, and the probable significant contribution of Alaskan <span class="hlt">glaciers</span> to rising sea level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C53D..02M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C53D..02M"><span>Remote Sensing Observations of Advancing and Surging Tidewater <span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McNabb, R. W.; Kääb, A.; Nuth, C.; Girod, L.; Truffer, M.; Fahnestock, M. A.</p> <p>2017-12-01</p> <p>Progress has been made in understanding the glaciological frontiers of tidewater <span class="hlt">glacier</span> dynamics and surge dynamics, though many aspects of these topics are not well-understood. Advances in the processing of digital elevation models (DEMs) from ASTER imagery, as well as the increased availability and temporal density of satellite images such as Landsat and the Sentinel missions, provide an unprecedented wealth of satellite data over <span class="hlt">glaciers</span>, providing new opportunities to learn about these topics. As one of the largest glaciated regions in the world outside of the Greenland and Antarctic ice sheets, <span class="hlt">glaciers</span> in Alaska and adjacent regions in Canada have been highlighted for their elevated contributions to global sea level rise, through both high levels of melt and frontal ablation/calving from a large number of tidewater <span class="hlt">glaciers</span>. The region is also home to a number of surging <span class="hlt">glaciers</span>. We focus on several tidewater <span class="hlt">glaciers</span> in the region, including Turner, Tsaa, Harvard, and Meares <span class="hlt">Glaciers</span>. Turner <span class="hlt">Glacier</span> is a surge-type tidewater <span class="hlt">glacier</span> with a surge period of approximately eight years, while Tsaa <span class="hlt">Glacier</span> is a tidwewater <span class="hlt">glacier</span> that has shown rapid swings in terminus position on the order of a year. Harvard and Meares <span class="hlt">Glaciers</span> have been steadily advancing since at least the mid-20th century, in contrast with neighboring <span class="hlt">glaciers</span> that are retreating. Using a combination of ASTER, Landsat, and Sentinel data, we present and examine high-resolution time series of elevation, velocity, and terminus position for these <span class="hlt">glaciers</span>, as well as updated estimates of volume change and frontal ablation rates, including on sub-annual time scales. Preliminary investigations of elevation change on Turner <span class="hlt">Glacier</span> show that changes are most pronounced in the lower reaches of the <span class="hlt">glacier</span>, below a prominent icefall approximately 15km from the head of the <span class="hlt">glacier</span>. On Harvard and Meares <span class="hlt">Glaciers</span>, elevation changes in the upper reaches of both <span class="hlt">glaciers</span> have been generally small or</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70031539','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70031539"><span>Oceanography of <span class="hlt">Glacier</span> Bay, Alaska: Implications for biological patterns in a glacial fjord estuary</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Etherington, L.L.; Hooge, P.N.; Hooge, Elizabeth Ross; Hill, D.F.</p> <p>2007-01-01</p> <p>Alaska, U.S.A, is one of the few remaining locations in the world that has fjords that contain temperate idewater <span class="hlt">glaciers</span>. Studying such estuarine systems provides vital information on how deglaciation affects oceanographic onditions of fjords and surrounding coastal waters. The oceanographic system of <span class="hlt">Glacier</span> Bay, Alaska, is of particular interest ue to the rapid deglaciation of the Bay and the resulting changes in the estuarine environment, the relatively high oncentrations of marine mammals, seabirds, fishes, and invertebrates, and the Bay’s status as a national <span class="hlt">park</span>, where ommercial fisheries are being phased out. We describe the first comprehensive broad-scale analysis of physical and iological oceanographic conditions within <span class="hlt">Glacier</span> Bay based on CTD measurements at 24 stations from 1993 to 2002. easonal patterns of near-surface salinity, temperature, stratification, turbidity, and euphotic depth suggest that freshwater nput was highest in summer, emphasizing the critical role of <span class="hlt">glacier</span> and snowmelt to this system. Strong and persistent tratification of surface waters driven by freshwater input occurred from spring through fall. After accounting for seasonal nd spatial variation, several of the external physical factors (i.e., air temperature, precipitation, day length) explained a large mount of variation in the physical properties of the surface waters. Spatial patterns of phytoplankton biomass varied hroughout the year and were related to stratification levels, euphotic depth, and day length. We observed hydrographic atterns indicative of strong competing forces influencing water column stability within <span class="hlt">Glacier</span> Bay: high levels of freshwater ischarge promoted stratification in the upper fjord, while strong tidal currents over the Bay’s shallow entrance sill enhanced ertical mixing. Where these two processes met in the central deep basins there were optimal conditions of intermediate tratification, higher light levels, and potential nutrient renewal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015GPC...128....1P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015GPC...128....1P"><span>Region-wide <span class="hlt">glacier</span> mass budgets and area changes for the Central Tien Shan between ~ 1975 and 1999 using Hexagon KH-9 imagery</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pieczonka, Tino; Bolch, Tobias</p> <p>2015-05-01</p> <p>The meltwater released by the <span class="hlt">glaciers</span> in the Central Tien Shan feeds in particular the Tarim River which is the main artery for the oases at the northern margin of the Taklamakan desert. The correct assessment of the contribution of the <span class="hlt">glaciers</span>' meltwater to the total runoff is hampered by the lack of long-term measurements of <span class="hlt">glacier</span> mass budgets. Digital terrain models (DTMs) for the different regions in the Central Tien Shan were generated based on ~ 1975 KH-9 Hexagon imagery and compared to the SRTM3 DTM acquired in February 2000. Moreover, <span class="hlt">glacier</span> area changes for the period ~ 1975-2008 have been measured by means of multi-temporal optical satellite imagery. The geodetic mass budget estimates for a <span class="hlt">glacierized</span> area of 5000 km2 revealed increasing mass loss <span class="hlt">east</span> to west and from the inner to the outer ranges. Highest mass loss accompanied by the most pronounced <span class="hlt">glacier</span> retreat was found for the Ak-Shirak massif with a region-wide mass balance of - 0.51 ± 0.36 m w.e. a- 1 and a rate of area change of - 0.27 ± 0.15% a- 1, whilst moderate mass loss was observed for the Inylchek (0.20 ± 0.44 m w.e. a- 1) and Tomur area (0.33 ± 0.30 m w.e. a- 1) despite partly debris cover. These latter regions also revealed the lowest <span class="hlt">glacier</span> shrinkage within the entire Central Tien Shan. The total <span class="hlt">glacier</span> mass loss of 0.35 ± 0.34 m w.e. a- 1 is, however, within the global average whilst the <span class="hlt">glacier</span> area shrinkage is comparatively low. On average, the investigated <span class="hlt">glacierized</span> area of ~ 6600 km2 shrank by 0.11 ± 0.15% a- 1 only. We could also identify several surge-type <span class="hlt">glaciers</span>. The results are consistent with in-situ mass balance measurements for Karabatkak <span class="hlt">Glacier</span> and previously published results of the Ak-Shirak range proving the suitability of declassified imagery for <span class="hlt">glacier</span> change investigations. The contribution to the runoff of Aksu River, the largest tributary of the Tarim River, due to <span class="hlt">glacier</span> imbalance has been determined at ~ 20% for the 1975-2000 period.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.C43C0689E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.C43C0689E"><span>A High-Resolution Sensor Network for Monitoring <span class="hlt">Glacier</span> Dynamics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Edwards, S.; Murray, T.; O'Farrell, T.; Rutt, I. C.; Loskot, P.; Martin, I.; Selmes, N.; Aspey, R.; James, T.; Bevan, S. L.; Baugé, T.</p> <p>2013-12-01</p> <p>Changes in Greenland and Antarctic ice sheets due to ice flow/ice-berg calving are a major uncertainty affecting sea-level rise forecasts. Latterly GNSS (Global Navigation Satellite Systems) have been employed extensively to monitor such <span class="hlt">glacier</span> dynamics. Until recently however, the favoured methodology has been to deploy sensors onto the <span class="hlt">glacier</span> surface, collect data for a period of time, then retrieve and download the sensors. This approach works well in less dynamic environments where the risk of sensor loss is low. In more extreme environments e.g. approaching the glacial calving front, the risk of sensor loss and hence data loss increases dramatically. In order to provide glaciologists with new insights into flow dynamics and calving processes we have developed a novel sensor network to increase the robustness of data capture. We present details of the technological requirements for an in-situ Zigbee wireless streaming network infrastructure supporting instantaneous data acquisition from high resolution GNSS sensors thereby increasing data capture robustness. The data obtained offers new opportunities to investigate the interdependence of mass flow, uplift, velocity and geometry and the network architecture has been specifically designed for deployment by helicopter close to the calving front to yield unprecedented detailed information. Following successful field trials of a pilot three node network during 2012, a larger 20 node network was deployed on the fast-flowing Helheim <span class="hlt">glacier</span>, south-<span class="hlt">east</span> Greenland over the summer months of 2013. The utilisation of dual wireless transceivers in each <span class="hlt">glacier</span> node, multiple frequencies and four ';collector' stations located on the valley sides creates overlapping networks providing enhanced capacity, diversity and redundancy of data 'back-haul', even close to ';floor' RSSI (Received Signal Strength Indication) levels around -100 dBm. Data loss through radio packet collisions within sub-networks are avoided through the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70036254','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70036254"><span>A complex relationship between calving <span class="hlt">glaciers</span> and climate</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Post, A.; O'Neel, S.; Motyka, R.J.; Streveler, G.</p> <p>2011-01-01</p> <p>Many terrestrial <span class="hlt">glaciers</span> are sensitive indicators of past and present climate change as atmospheric temperature and snowfall modulate <span class="hlt">glacier</span> volume. However, climate interpretations based on <span class="hlt">glacier</span> behavior require careful selection of representative <span class="hlt">glaciers</span>, as was recently pointed out for surging and debris-covered <span class="hlt">glaciers</span>, whose behavior often defies regional <span class="hlt">glacier</span> response to climate [Yde and Paasche, 2010]. Tidewater calving <span class="hlt">glaciers</span> (TWGs)mountain <span class="hlt">glaciers</span> whose termini reach the sea and are generally grounded on the seaflooralso fall into the category of non-representative <span class="hlt">glaciers</span> because the regional-scale asynchronous behavior of these <span class="hlt">glaciers</span> clouds their complex relationship with climate. TWGs span the globe; they can be found both fringing ice sheets and in high-latitude regions of each hemisphere. TWGs are known to exhibit cyclic behavior, characterized by slow advance and rapid, unstable retreat, largely independent of short-term climate forcing. This so-called TWG cycle, first described by Post [1975], provides a solid foundation upon which modern investigations of TWG stability are built. Scientific understanding has developed rapidly as a result of the initial recognition of their asynchronous cyclicity, rendering greater insight into the hierarchy of processes controlling regional behavior. This has improved the descriptions of the strong dynamic feedbacks present during retreat, the role of the ocean in TWG dynamics, and the similarities and differences between TWG and ice sheet outlet <span class="hlt">glaciers</span> that can often support floating tongues.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.C53E0718A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.C53E0718A"><span>Remote Sensing Estimates of <span class="hlt">Glacier</span> Mass Balance Changes in the Himalayas of Nepal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ambinakudige, S.; Joshi, K.</p> <p>2011-12-01</p> <p>Mass balance changes of <span class="hlt">glaciers</span> are important indicators of climate change. There are only 30 'reference' <span class="hlt">glaciers</span> in the world that have continuous mass balance data with world <span class="hlt">glacier</span> monitoring service since 1976. Especially, Himalayan <span class="hlt">glaciers</span> are conspicuously absent from global mass balance records. This shows the urgent need for mass balance data for <span class="hlt">glaciers</span> throughout the world. In this study, we estimated mass balance of some major <span class="hlt">glaciers</span> in the Sagarmatha National <span class="hlt">Park</span> (SNP) in Nepal using remote sensing applications. The SNP is one of the densest glaciated regions in the Himalayan range consisting approximately 296 glacial lakes. The region has experienced several glacial lake outburst floods (GLOFs) in recent years, causing extensive damage to local infrastructure and loss of human life. In general, mass balance is determined at seasonal or yearly intervals. Because of the rugged and difficult terrain of the Himalayan region, there are only a few field based measurements of mass balance available. Moreover, there are only few cases where the applications of remote sensing methods were used to calculate mass balance of the Himalayan <span class="hlt">glaciers</span> due to the lack of accurate elevation data. Studies have shown that estimations of mass balance using remote sensing applications were within the range of field-based mass balance measurements from the same period. This study used ASTER VNIR, 3N (nadir view) and 3B (backward view) bands to generate Digital Elevation Models (DEMs) for the SNP area. 3N and 3B bands generate an along track stereo pair with a base-to-height (B/H) ratio of about 0.6. Accurate measurement of ground control points (GCPs), their numbers and distribution are important inputs in creating accurate DEMs. Because of the availability of topographic maps for this area, we were able to provide very accurate GCPs, in sufficient numbers and distribution. We created DEMs for the years 2002, 2003, 2004 and 2005 using ENVI DEM extraction tool. Bands</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhDT........60D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhDT........60D"><span>Evaluating <span class="hlt">glacier</span> movement fluctuations using remote sensing: A case study of the Baird, Patterson, LeConte, and Shakes <span class="hlt">glaciers</span> in central Southeastern Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Davidson, Robert Howard</p> <p></p> <p>Global Land Survey (GLS) data encompassing Landsat Multispectral Scanner (MSS), Landsat 5's Thematic Mapper (TM), and Landsat 7's Enhanced Thematic Mapper Plus (ETM+) were used to determine the terminus locations of Baird, Patterson, LeConte, and Shakes <span class="hlt">Glaciers</span> in Alaska in the time period 1975-2010. The sequences of the terminuses locations were investigated to determine the movement rates of these <span class="hlt">glaciers</span> with respect to specific physical and environmental conditions. GLS data from 1975, 1990, 2000, 2005, and 2010 in false-color composite images enhancing ice-snow differentiation and Iterative Self-Organizing (ISO) Data Cluster Unsupervised Classifications were used to 1) quantify the movement rates of Baird, Patterson, LeConte, and Shakes <span class="hlt">Glaciers</span>; 2) analyze the movement rates for <span class="hlt">glaciers</span> with similar terminal terrain conditions and; 3) analyze the movement rates for <span class="hlt">glaciers</span> with dissimilar terminal terrain conditions. From the established sequence of terminus locations, movement distances were quantified between the <span class="hlt">glacier</span> locations. Movement distances were then compared to see if any correlation existed between <span class="hlt">glaciers</span> with similar or dissimilar terminal terrain conditions. The Global Land Ice Measurement from Space (GLIMS) data was used as a starting point from which <span class="hlt">glacier</span> movement was measured for Baird, Patterson, and LeConte <span class="hlt">Glaciers</span> only as the Shakes <span class="hlt">Glacier</span> is currently not included in the GLIMS database. The National Oceanographic and Atmospheric Administration (NOAA) temperature data collected at the Petersburg, Alaska, meteorological station (from January 1, 1973 to December 31, 2009) were used to help in the understanding of the climatic condition in this area and potential impact on <span class="hlt">glaciers</span> terminus. Results show that <span class="hlt">glaciers</span> with similar terminal terrain conditions (Patterson and Shakes <span class="hlt">Glaciers</span>) and <span class="hlt">glaciers</span> with dissimilar terminal terrain conditions (Baird, Patterson, and LeConte <span class="hlt">Glaciers</span>) did not exhibit similar movement rates</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4981079','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4981079"><span>Dynamics of <span class="hlt">glacier</span> calving at the ungrounded margin of Helheim <span class="hlt">Glacier</span>, southeast Greenland</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Selmes, Nick; James, Timothy D.; Edwards, Stuart; Martin, Ian; O'Farrell, Timothy; Aspey, Robin; Rutt, Ian; Nettles, Meredith; Baugé, Tim</p> <p>2015-01-01</p> <p>Abstract During summer 2013 we installed a network of 19 GPS nodes at the ungrounded margin of Helheim <span class="hlt">Glacier</span> in southeast Greenland together with three cameras to study iceberg calving mechanisms. The network collected data at rates up to every 7 s and was designed to be robust to loss of nodes as the <span class="hlt">glacier</span> calved. Data collection covered 55 days, and many nodes survived in locations right at the <span class="hlt">glacier</span> front to the time of iceberg calving. The observations included a number of significant calving events, and as a consequence the <span class="hlt">glacier</span> retreated ~1.5 km. The data provide real‐time, high‐frequency observations in unprecedented proximity to the calving front. The <span class="hlt">glacier</span> calved by a process of buoyancy‐force‐induced crevassing in which the ice downglacier of flexion zones rotates upward because it is out of buoyant equilibrium. Calving then occurs back to the flexion zone. This calving process provides a compelling and complete explanation for the data. Tracking of oblique camera images allows identification and characterisation of the flexion zones and their propagation downglacier. Interpretation of the GPS data and camera data in combination allows us to place constraints on the height of the basal cavity that forms beneath the rotating ice downglacier of the flexion zone before calving. The flexion zones are probably formed by the exploitation of basal crevasses, and theoretical considerations suggest that their propagation is strongly enhanced when the <span class="hlt">glacier</span> base is deeper than buoyant equilibrium. Thus, this calving mechanism is likely to dominate whenever such geometry occurs and is of increasing importance in Greenland. PMID:27570721</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Geomo.290...58B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Geomo.290...58B"><span>Contrasting medial moraine development at adjacent temperate, maritime <span class="hlt">glaciers</span>: Fox and Franz Josef <span class="hlt">Glaciers</span>, South Westland, New Zealand</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brook, Martin; Hagg, Wilfried; Winkler, Stefan</p> <p>2017-08-01</p> <p>Medial moraines form important pathways for sediment transportation in valley <span class="hlt">glaciers</span>. Despite the existence of well-defined medial moraines on several <span class="hlt">glaciers</span> in the New Zealand Southern Alps, medial moraines there have hitherto escaped attention. The evolving morphology and debris content of medial moraines on Franz Josef <span class="hlt">Glacier</span> and Fox <span class="hlt">Glacier</span> on the western flank of the Southern Alps is the focus of this study. These temperate maritime <span class="hlt">glaciers</span> exhibit accumulation zones of multiple basins that feed narrow tongues flowing down steep valleys and terminate 400 m above sea level. The medial moraines at both <span class="hlt">glaciers</span> become very prominent in the lower ablation zones, where the medial moraines widen, and develop steeper flanks coeval with an increase in relative relief. Medial moraine growth appears somewhat self-limiting in that relief and slope angle increase eventually lead to transport of debris away from the medial moraine by mass-movement-related processes. Despite similarities in overall morphologies, a key contrast in medial moraine formation exists between the two <span class="hlt">glaciers</span>. At Fox <span class="hlt">Glacier</span>, the medial moraine consists of angular rockfall-derived debris, folded to varying degrees along flow-parallel axes throughout the tongue. The debris originates above the ELA, coalesces at flow-unit boundaries, and takes a medium/high level transport pathway before subsequently emerging at point-sources aligned with gently dipping fold hinges near the snout. In contrast at Franz Josef <span class="hlt">Glacier</span>, the medial moraine emerges farther down-<span class="hlt">glacier</span> immediately below a prominent rock knob. Clasts show a mix of angular to rounded shapes representing high level transport and subglacially transported materials, the latter facies possibly also elevated by supraglacial routing of subglacial meltwater. Our observations confirm that a variety of different debris sources, transport pathways, and structural glaciological processes can interact to form medial moraines within New Zealand</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/964677','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/964677"><span>Environmental Management Waste Management Facility Proxy Waste Lot Profile 6.999 for Building K-25 West Wing, <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee</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>Rigsby V.P.</p> <p>2009-02-12</p> <p>In 1989, the Oak Ridge Reservation (ORR), which includes the <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> (ETTP), was placed on the Comprehensive Environmental Response, Compensation, and Liability Act of 1980 (CERCLA) National Priorities List. The Federal Facility Agreement (FFA) (DOE 1992), effective January 1, 1992, now governs environmental restoration activities conducted under CERCLA at the ORR. Following signing of the FFA, U.S. Department of Energy (DOE), U.S. Environmental Protection Agency (EPA), and the state of Tennessee signed the Oak Ridge Accelerated Cleanup Plan Agreement on June 18, 2002. The purpose of this agreement is to define a streamlined decision-making process to facilitatemore » the accelerated implementation of cleanup, resolve ORR milestone issues, and establish future actions necessary to complete the accelerated cleanup plan by the end of fiscal year 2008. While the FFA continues to serve as the overall regulatory framework for remediation, the Accelerated Cleanup Plan Agreement supplements existing requirements to streamline the decision-making process. Decontamination and decommissioning (D&D) activities of Bldg. K-25, the original gaseous diffusion facility, is being conducted by Bechtel Jacobs Company LLC (BJC) on behalf of the DOE. The planned CERCLA action covering disposal of building structure and remaining components from the K-25 building is scheduled as a non-time-critical CERCLA action as part of DOE's continuous risk reduction strategy for ETTP. The K-25 building is proposed for D&D because of its poor physical condition and the expense of surveillance and maintenance activities. The K-25/K-27 D&D Project proposes to dispose of the commingled waste listed below from the K-25 west side building structure and remaining components and process gas equipment and piping at the Environmental Management Waste Management Facility (EMWMF) under waste disposal proxy lot (WPXL) 6.999: (1) Building structure (e.g. concrete floors [excluding</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/984476','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/984476"><span>Environmental Baseline Survey Report for the Title Transfer of Parcel ED-9 at the <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span>, Oak Ridge, Tennessee</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>SAIC</p> <p>2010-05-01</p> <p>This environmental baseline survey (EBS) report documents the baseline environmental conditions of the U. S. Department of Energy's (DOE's) Parcel ED-9 at the <span class="hlt">East</span> Tennessee Technology <span class="hlt">Park</span> (ETTP). Parcel ED-9 consists of about 13 acres that DOE proposes to transfer to Heritage Center, LLC (hereafter referred to as 'Heritage Center'), a subsidiary of the Community Reuse Organization of <span class="hlt">East</span> Tennessee (CROET). The 13 acres include two tracts of land, referred to as ED-9A (7.06 acres) and ED-9B (5.02 acres), and a third tract consisting of about 900 linear feet of paved road and adjacent right-of-way, referred to as ED-9C (0.98more » acres). Transfer of the title to ED-9 will be by deed under a Covenant Deferral Request (CDR) pursuant to Section 120(h)(3)(C) of the Comprehensive Environmental Response, Compensation, and Liability Act of 1980 (CERCLA). This report provides a summary of information to support the transfer of this government-owned property at ETTP to a non-federal entity.« less</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('http://adsabs.harvard.edu/abs/2017EGUGA..1912518P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1912518P"><span>Influencing factors on the cooling effect of coarse blocky top-layers on relict rock <span class="hlt">glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pauritsch, Marcus; Wagner, Thomas; Mayaud, Cyril; Thalheim, Felix; Kellerer-Pirklbauer, Andreas; Winkler, Gerfried</p> <p>2017-04-01</p> <p> impact on the thermal regime of the Schöneben Rock <span class="hlt">Glacier</span> and, as the major wind direction, especially for higher wind speeds, is from west towards <span class="hlt">east</span>, it is suspected that wind-forced convection is even more important at the Dürrtal Rock <span class="hlt">Glacier</span>. The limited incident solar radiation at the Schöneben Rock <span class="hlt">Glacier</span> results in a longer seasonal snow cover that appears earlier in autumn and can persist longer during the melting season. Moreover, with the predominant westerly wind, snow is supposedly transported from neighboring catchments (i.a. the Dürrtal Rock <span class="hlt">Glacier</span> catchment) towards the Schöneben Rock <span class="hlt">Glacier</span> catchment. Thus, in times with relatively cold air temperatures the coarse blocky surface at the Dürrtal Rock <span class="hlt">Glacier</span> is better connected to the atmosphere than the more northern exposed Schöneben rock <span class="hlt">glacier</span> because of the missing or only partial snow cover, which results in an enhanced cooling effect. It can be concluded that the cooling effect of coarse blocky debris is highly variable in alpine environments and can show considerable variations, depending on the heterogeneous structure of the layer itself and a complex interplay of slope aspect-related microclimatic effects such as incident solar radiation, predominant wind direction and snow cover dynamics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009JGlac..55..292M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009JGlac..55..292M"><span>Solar radiation, cloudiness and longwave radiation over low-latitude <span class="hlt">glaciers</span>: implications for mass-balance modelling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mölg, Thomas; Cullen, Nicolas J.; Kaser, Georg</p> <p></p> <p>Broadband radiation schemes (parameterizations) are commonly used tools in <span class="hlt">glacier</span> mass-balance modelling, but their performance at high altitude in the tropics has not been evaluated in detail. Here we take advantage of a high-quality 2 year record of global radiation (G) and incoming longwave radiation (L↓) measured on Kersten <span class="hlt">Glacier</span>, Kilimanjaro, <span class="hlt">East</span> Africa, at 5873 m a.s.l., to optimize parameterizations of G and L↓. We show that the two radiation terms can be related by an effective cloud-cover fraction neff, so G or L↓ can be modelled based on neff derived from measured L↓ or G, respectively. At neff = 1, G is reduced to 35% of clear-sky G, and L↓ increases by 45-65% (depending on altitude) relative to clear-sky L↓. Validation for a 1 year dataset of G and L↓ obtained at 4850 m on Glaciar Artesonraju, Peruvian Andes, yields a satisfactory performance of the radiation scheme. Whether this performance is acceptable for mass-balance studies of tropical <span class="hlt">glaciers</span> is explored by applying the data from Glaciar Artesonraju to a physically based mass-balance model, which requires, among others, G and L↓ as forcing variables. Uncertainties in modelled mass balance introduced by the radiation parameterizations do not exceed those that can be caused by errors in the radiation measurements. Hence, this paper provides a tool for inclusion in spatially distributed mass-balance modelling of tropical <span class="hlt">glaciers</span> and/or extension of radiation data when only G or L↓ is measured.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..1212086L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..1212086L"><span>A new satellite-derived <span class="hlt">glacier</span> inventory for Western Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Le Bris, Raymond; Frey, Holger; Paul, Frank; Bolch, Tobias</p> <p>2010-05-01</p> <p><span class="hlt">Glaciers</span> and ice caps are essential components of studies related to climate change impact assessment. <span class="hlt">Glacier</span> inventories provide the required baseline data to perform the related analysis in a consistent and spatially representative manner. In particular, the calculation of the current and future contribution to global sea-level rise from heavily <span class="hlt">glacierized</span> regions is a major demand. One of the regions, where strong mass losses and geometric changes of <span class="hlt">glaciers</span> have been observed recently is Alaska. Unfortunately, the digitally available data base of <span class="hlt">glacier</span> extent is quite rough and based on rather old maps from the 1960s. Accordingly, the related calculations and extrapolations are imprecise and an updated <span class="hlt">glacier</span> inventory is urgently required. Here we present first results of a new <span class="hlt">glacier</span> inventory for Western Alaska that is prepared in the framework of the ESA project Glob<span class="hlt">Glacier</span> and is based on freely available orthorectified Landsat TM and ETM+ scenes from USGS. The analysed region covers the Tordrillo, Chigmit and Chugach Mts. as well as the Kenai Peninsula. In total, 8 scenes acquired between 2002 and 2009 were used covering c. 20.420 km2 of <span class="hlt">glaciers</span>. All <span class="hlt">glacier</span> types are present in this region, incl. outlet <span class="hlt">glaciers</span> from icefields, <span class="hlt">glacier</span> clad volcanoes, and calving <span class="hlt">glaciers</span>. While well established automated <span class="hlt">glacier</span> mapping techniques (band rationing) are applied to map clean and slightly dirty <span class="hlt">glacier</span> ice, many <span class="hlt">glaciers</span> are covered by debris or volcanic ash and outlines need manual corrections during post-processing. Prior to the calculation of drainage divides from DEM-based watershed analysis, we performed a cross-comparative analysis of DEMs from USGS, ASTER (GDEM) and SRTM 1 for Kenai Peninsula. This resulted in the decision to use the USGS DEM for calculating the drainage divides and most of the topographic inventory parameters, and the more recent GDEM to derive minimum elevation for each <span class="hlt">glacier</span>. A first statistical analysis of the results</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1169678','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1169678"><span>Los Alamos Canyon Ice Rink <span class="hlt">Parking</span> Flood Plain 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>Hathcock, Charles Dean; Keller, David Charles</p> <p>2015-02-10</p> <p>The project location is in Los Alamos Canyon <span class="hlt">east</span> of the ice rink facility at the intersection of West and Omega roads (Figure 1). Forty eight <span class="hlt">parking</span> spaces will be constructed on the north and south side of Omega Road, and a lighted walking path will be constructed to the ice rink. Some trees will be removed during this action. A guardrail of approximately 400 feet will be constructed along the north side of West Road to prevent unsafe <span class="hlt">parking</span> in that area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004ggav.rept......','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004ggav.rept......"><span>Geenland <span class="hlt">Glacier</span> Albedo Variability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p></p> <p>2004-01-01</p> <p>The program for Arctic Regional Climate Assessment (PARCA) is a NASA-funded project with the prime goal of addressing the mass balance of the Greenland ice sheet. Since the formal initiation of the program in 1995, there has been a significant improvement in the estimates of the mass balance of the ice sheet. Results from this program reveal that the high-elevation regions of the ice sheet are approximately in balance, but the margins are thinning. Laser surveys reveal significant thinning along 70 percent of the ice sheet periphery below 2000 m elevations, and in at least one outlet <span class="hlt">glacier</span>, Kangerdlugssuaq in southeast Greenland, thinning has been as much as 10 m/yr. This study examines the albedo variability in four outlet <span class="hlt">glaciers</span> to help separate out the relative contributions of surface melting versus ice dynamics to the recent mass balance changes. Analysis of AVHRR Polar Pathfinder albedo shows that at the Petermann and Jakobshavn <span class="hlt">glaciers</span>, there has been a negative trend in albedo at the <span class="hlt">glacier</span> terminus from 1981 to 2000, whereas the Stor+strommen and Kangerdlugssuaq <span class="hlt">glaciers</span> show slightly positive trends in albedo. These findings are consistent with recent observations of melt extent from passive microwave data which show more melt on the western side of Greenland and slightly less on the eastern side. Significance of albedo trends will depend on where and when the albedo changes occur. Since the majority of surface melt occurs in the shallow sloping western margin of the ice sheet where the shortwave radiation dominates the energy balance in summer (e.g. Jakobshavn region) this region will be more sensitive to changes in albedo than in regions where this is not the case. Near the Jakobshavn <span class="hlt">glacier</span>, even larger changes in albedo have been observed, with decreases as much as 20 percent per decade.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040050637','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040050637"><span>Greenland <span class="hlt">Glacier</span> Albedo Variability</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2004-01-01</p> <p>The program for Arctic Regional Climate Assessment (PARCA) is a NASA-funded project with the prime goal of addressing the mass balance of the Greenland ice sheet. Since the formal initiation of the program in 1995, there has been a significant improvement in the estimates of the mass balance of the ice sheet. Results from this program reveal that the high-elevation regions of the ice sheet are approximately in balance, but the margins are thinning. Laser surveys reveal significant thinning along 70 percent of the ice sheet periphery below 2000 m elevations, and in at least one outlet <span class="hlt">glacier</span>, Kangerdlugssuaq in southeast Greenland, thinning has been as much as 10 m/yr. This study examines the albedo variability in four outlet <span class="hlt">glaciers</span> to help separate out the relative contributions of surface melting versus ice dynamics to the recent mass balance changes. Analysis of AVHRR Polar Pathfinder albedo shows that at the Petermann and Jakobshavn <span class="hlt">glaciers</span>, there has been a negative trend in albedo at the <span class="hlt">glacier</span> terminus from 1981 to 2000, whereas the Stor+strommen and Kangerdlugssuaq <span class="hlt">glaciers</span> show slightly positive trends in albedo. These findings are consistent with recent observations of melt extent from passive microwave data which show more melt on the western side of Greenland and slightly less on the eastern side. Significance of albedo trends will depend on where and when the albedo changes occur. Since the majority of surface melt occurs in the shallow sloping western margin of the ice sheet where the shortwave radiation dominates the energy balance in summer (e.g. Jakobshavn region) this region will be more sensitive to changes in albedo than in regions where this is not the case. Near the Jakobshavn <span class="hlt">glacier</span>, even larger changes in albedo have been observed, with decreases as much as 20 percent per decade.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AGUFMIP52A0750K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AGUFMIP52A0750K"><span>Five 'Supercool' Icelandic <span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Knudsen, O.; Roberts, M. J.; Roberts, M. J.; Tweed, F. S.; Russell, A. J.; Lawson, D. E.; Larson, G. J.; Evenson, E. B.; Bjornsson, H.</p> <p>2001-12-01</p> <p>Sediment entrainment by glaciohydraulic supercooling has recently been demonstrated as an effective process at Matanuska <span class="hlt">glacier</span>, Alaska. Although subfreezing meltwater temperatures have been recorded at several Alaskan <span class="hlt">glaciers</span>, the link between supercooling and sediment accretion remains confined to Matanuska. This study presents evidence of glaciohydraulic supercooling and associated basal ice formation from five Icelandic <span class="hlt">glaciers</span>: Skeidarárjökull, Skaftafellsjökull, Kvíárjökull, Flaájökull, and Hoffellsjökull. These observations provide the best example to-date of glaciohydraulic supercooling and related sediment accretion outside Alaska. Fieldwork undertaken in March, July and August 2001 confirmed that giant terraces of frazil ice, diagnostic of the presence of supercooled water, are forming around subglacial artesian vents. Frazil flocs retrieved from these vents contained localised sandy nodules at ice crystal boundaries. During periods of high discharge, sediment-laden frazil flocs adhere to the inner walls of vents, and continue to trap suspended sediment. Bands of debris-rich frazil ice, representing former vents, are texturally similar to basal ice exposures at the <span class="hlt">glacier</span> margins, implying a process-form relationship between glaciohydraulic freeze-on and basal ice formation. It is hypothesised that glaciohydraulic supercooling is generating thick sequences of basal ice. Observations also confirm that in situ melting of basal ice creates thick sedimentary sequences, as sediment structures present in the basal ice can be clearly traced into ice-marginal ridges. Glaciohydraulic supercooling is an effective sediment entrainment mechanism at Icelandic <span class="hlt">glaciers</span>. Supercooling has the capacity to generate thick sequences of basal ice and the sediments present in basal ice can be preserved. These findings are incompatible with established theories of intraglacial sediment entrainment and basal ice formation; instead, they concur with, and extend, the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/dc0983/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/dc0983/"><span>Rainbow Pool, Eastern portion of West Potomac <span class="hlt">Park</span>; bounded 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>Rainbow Pool, Eastern portion of West Potomac <span class="hlt">Park</span>; bounded by Elm Walks to the north and south, Seventeenth Street to the <span class="hlt">east</span> and the Reflecting Pool to the west, Washington, District of Columbia, DC</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca1118.photos.010724p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca1118.photos.010724p/"><span>30. Photocopy of photograph (from National <span class="hlt">Park</span> Service, San Francisco, ...</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>30. Photocopy of photograph (from National <span class="hlt">Park</span> Service, San Francisco, California, 1930 (?) EXTERIOR, <span class="hlt">EAST</span> SIDE OF MISSIONA AFTER RESTORATION, C. 1930 (?) - Mission San Francisco Solano de Sonoma, First & Spain Streets, Sonoma, Sonoma County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C12B..02A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C12B..02A"><span>Bed Topography of Jakobshavn Isbræ and Helheim <span class="hlt">Glacier</span>, Greenland from High-Resolution Gravity Data Combined with Other Observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>An, L.; Rignot, E. J.; Morlighem, M.; Paden, J. D.; Holland, D.</p> <p>2016-12-01</p> <p>Jakobshavn Isbræ (JKS) is the most active and largest outlet <span class="hlt">glacier</span> in West Greenland, draining approximately 6.5% of the ice sheet. JKS sped up more than twofold since 2002 and contributed nearly 1 mm of global sea level rise during the period from 2000 to 2011. Helheim <span class="hlt">glacier</span> is the fastest flowing outlet <span class="hlt">glacier</span> in <span class="hlt">East</span> Greenland and accelerated by a factor two during a strong thinning period in early 2000s. To interpret the recent and future evolution of these <span class="hlt">glaciers</span>, it is essential to know their ice thickness and bed topography as well as the bathymetry in the fjords. Here, we present a novel approach to infer the <span class="hlt">glacier</span> bed topography, ice thickness and sea floor bathymetry near the grounding line using high-resolution airborne gravity data from AIRGrav. AIRGrav data were collected in August 2012 with a helicopter platform, at 500 m spacing grid, 50 knots ground speed, 80 m ground clearance, with sub-milligal accuracy, i.e. higher than NASA Operation IceBridge (OIB)'s 5.2 km resolution, 290 knots, and 450 m clearance. We use a 3D inversion of the gravity data combining our observations and a forward modeling of the surrounding gravity field with point measurements of the bathymetry at the ice-ocean boundary and a reconstruction of the <span class="hlt">glacier</span> bed topography upstream using a mass conservation method combining re-analyzed airborne radar-derived ice thickness data from CReSIS with ice flow motion vectors from satellite radar interferometry. The results provide a more accurate view of the bed topography of these <span class="hlt">glaciers</span> and resolve major uncertainties from past attempts to probe the deepest part of the bed near the ice front from radio echo sounding data alone. The results reveal that the JKS is now retreating into an even deeper bed, from 600 m in 1996 to 900 m at present and 1,400 m in the next 25 km. The <span class="hlt">glacier</span> will continue to retreat probably at an increasing rate (0.6 km/yr at present) along a retrograde bed, i.e. into thicker ice. On Helheim</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015QSRv..114...78W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015QSRv..114...78W"><span>Reconstructing Holocene <span class="hlt">glacier</span> activity at Langfjordjøkelen, Arctic Norway, using multi-proxy fingerprinting of distal <span class="hlt">glacier</span>-fed lake sediments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wittmeier, Hella E.; Bakke, Jostein; Vasskog, Kristian; Trachsel, Mathias</p> <p>2015-04-01</p> <p>Late Glacial and Holocene <span class="hlt">glacier</span> fluctuations are important indicators of climate variability in the northern polar region and contain knowledge vital to understanding and predicting present and future climate changes. However, there still is a lack of robustly dated terrestrial climate records from Arctic Norway. Here, we present a high-resolution relative <span class="hlt">glacier</span> activity record covering the past ∼10,000 cal. a BP from the northern outlet of the Langfjordjøkelen ice cap in Arctic Norway. This record is reconstructed from detailed geomorphic mapping, multi-proxy sedimentary fingerprinting and analyses of distal <span class="hlt">glacier</span>-fed lake sediments. We used Principal Component Analysis to characterize sediments of glacial origin and trace them in a chain of downstream lakes. Of the variability in the sediment record of the uppermost Lake Jøkelvatnet, 73% can be explained by the first Principal Component axis and tied directly to upstream <span class="hlt">glacier</span> erosion, whereas the glacial signal becomes weaker in the more distal Lakes Store Rundvatnet and Storvatnet. Magnetic susceptibility and titanium count rates were found to be the most suitable indicators of Holocene <span class="hlt">glacier</span> activity in the distal <span class="hlt">glacier</span>-fed lakes. The complete deglaciation of the valley of Sør-Tverrfjorddalen occurred ∼10,000 cal. a BP, followed by a reduced or absent <span class="hlt">glacier</span> during the Holocene Thermal Optimum. The Langfjordjøkelen ice cap reformed with the onset of the Neoglacial ∼4100 cal. a BP, and the gradually increasing <span class="hlt">glacier</span> activity culminated at the end of the Little Ice Age in the early 20th century. Over the past 2000 cal. a BP, the record reflects frequent high-amplitude <span class="hlt">glacier</span> fluctuations. Periods of reduced <span class="hlt">glacier</span> activity were centered around 1880, 1600, 1250 and 950 cal. a BP, while intervals of increased <span class="hlt">glacier</span> activity occurred around 1680, 1090, 440 and 25 cal. a BP. The large-scale Holocene <span class="hlt">glacier</span> activity of the Langfjordjøkelen ice cap is consistent with regional temperature</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFM.C33C1303J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFM.C33C1303J"><span>The energy balance on the surface of a tropical <span class="hlt">glacier</span> tongue. Investigations on <span class="hlt">glacier</span> Artesonraju, Cordillera Blanca, Perú.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Juen, I.; Mölg, T.; Wagnon, P.; Cullen, N. J.; Kaser, G.</p> <p>2006-12-01</p> <p>The Cordillera Blanca in Perú is situated in the Outer Tropics spanning from 8 to 10 ° South. Solar incidence and air temperature show only minor seasonal variations whereas precipitation occurs mainly from October to April. An energy balance station was installed on the tongue of <span class="hlt">glacier</span> Artesonraju (4850 m a.s.l.) in March 2004. In this study each component of the energy balance on the <span class="hlt">glacier</span> surface is analysed separately over a full year, covering one dry and one wet season. During the dry season <span class="hlt">glacier</span> melt at the <span class="hlt">glacier</span> tongue is app. 0.5 m we per month. In the wet season <span class="hlt">glacier</span> melt is twice as much with 1 m we per month. This is due to higher energy fluxes and decreased sublimation during the wet season. With an energy balance model that has already been proved under tropical climate conditions (Mölg and Hardy, 2004) each energy flux is changed individually to evaluate the change in the amount of <span class="hlt">glacier</span> melt. First results indicate that a change in humidity related variables affects <span class="hlt">glacier</span> melt very differently in the dry and wet season, whereas a change in air temperature changes <span class="hlt">glacier</span> melt more constantly throughout the year.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.C11B0675M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.C11B0675M"><span>Comparison of the 2008-2011 and 1993-1995 Surges of Bering <span class="hlt">Glacier</span>, Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molnia, B. F.; Angeli, K.</p> <p>2011-12-01</p> <p> of August 1993 and left it heavily fractured. With the current surge, until July 2009, surface displacements were restricted to the area from west of, to northeast of the Grindle Hills. By November 18, 2010, the surge front reached Bering's terminus and left it more heavily fractured than in 1993. The current surge shows the same style and types of surface disruptions and deformations at the same locations as did the earlier surge. For example, in both surges, sinusoidal crevasses were first noted north of the Grindle Hills, while rifts were noted in the upper central piedmont lobe. The current surge has produced much more fracturing of the Medial Moraine Band than did the 1993-95 surge. Similarly, the extent of surface fracturing up-<span class="hlt">glacier</span> from the piedmont lobe is significantly greater in the current surge. During the 1993-95 surge, surface expression of the surge extended about 45 km <span class="hlt">east</span> of the western end of Juniper Island. In late July 2011, surge-related surface fractures extended nearly 90 km to the <span class="hlt">east</span>. The Steller lobe of the Bering <span class="hlt">Glacier</span> System has not been involved in either surge. Continued observations of the current surge, in the context of the 1993-95 surge, are providing significant insights into repeatable patterns of surging <span class="hlt">glacier</span> behavior. Bering <span class="hlt">Glacier</span> is an amazing natural laboratory at which to conduct these observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.C21D0667A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.C21D0667A"><span>An Analysis of Mass Balance of Chilean <span class="hlt">Glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ambinakudige, S.; Tetteh, L.</p> <p>2013-12-01</p> <p><span class="hlt">Glaciers</span> in Chile range from very small glacierets found on the isolated volcanoes of northern Chile to the 13,000 sq.km Southern Patagonian Ice Field. Regular monitoring of these <span class="hlt">glaciers</span> is very important as they are considered as sensitive indicators of climate change. Millions of people's lives are dependent on these <span class="hlt">glaciers</span> for fresh water and irrigation purpose. In this study, mass balances of several Chilean <span class="hlt">glaciers</span> were estimated using Aster satellite images between 2007 and 2012. Highly accurate DEMs were created with supplementary information from IceSat data. The result indicated a negative mass balance for many <span class="hlt">glaciers</span> indicating the need for further monitoring of <span class="hlt">glaciers</span> in the Andes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..1512341E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..1512341E"><span>Arctic polynya and <span class="hlt">glacier</span> interactions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Edwards, Laura</p> <p>2013-04-01</p> <p>Major uncertainties surround future estimates of sea level rise attributable to mass loss from the polar ice sheets and ice caps. Understanding changes across the Arctic is vital as major potential contributors to sea level, the Greenland Ice Sheet and the ice caps and <span class="hlt">glaciers</span> of the Canadian Arctic archipelago, have experienced dramatic changes in recent times. Most ice mass loss is currently focused at a relatively small number of <span class="hlt">glacier</span> catchments where ice acceleration, thinning and calving occurs at ocean margins. Research suggests that these tidewater <span class="hlt">glaciers</span> accelerate and iceberg calving rates increase when warming ocean currents increase melt on the underside of floating <span class="hlt">glacier</span> ice and when adjacent sea ice is removed causing a reduction in 'buttressing' back stress. Thus localised changes in ocean temperatures and in sea ice (extent and thickness) adjacent to major glacial catchments can impact hugely on the dynamics of, and hence mass lost from, terrestrial ice sheets and ice caps. Polynyas are areas of open water within sea ice which remain unfrozen for much of the year. They vary significantly in size (~3 km2 to > ~50,000 km2 in the Arctic), recurrence rates and duration. Despite their relatively small size, polynyas play a vital role in the heat balance of the polar oceans and strongly impact regional oceanography. Where polynyas develop adjacent to tidewater <span class="hlt">glaciers</span> their influence on ocean circulation and water temperatures may play a major part in controlling subsurface ice melt rates by impacting on the water masses reaching the calving front. Areas of open water also play a significant role in controlling the potential of the atmosphere to carry moisture, as well as allowing heat exchange between the atmosphere and ocean, and so can influence accumulation on (and hence thickness of) <span class="hlt">glaciers</span> and ice caps. Polynya presence and size also has implications for sea ice extent and therefore potentially the buttressing effect on neighbouring</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://wa.water.usgs.gov/projects/glacier/data/bidlake_AGU_2010.pdf','USGSPUBS'); return false;" href="http://wa.water.usgs.gov/projects/glacier/data/bidlake_AGU_2010.pdf"><span><span class="hlt">Glacier</span> modeling in support of field observations of mass balance at South Cascade <span class="hlt">Glacier</span>, Washington, USA</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Josberger, Edward G.; Bidlake, William R.</p> <p>2010-01-01</p> <p>The long-term USGS measurement and reporting of mass balance at South Cascade <span class="hlt">Glacier</span> was assisted in balance years 2006 and 2007 by a new mass balance model. The model incorporates a temperature-index melt computation and accumulation is modeled from <span class="hlt">glacier</span> air temperature and gaged precipitation at a remote site. Mass balance modeling was used with glaciological measurements to estimate dates and magnitudes of critical mass balance phenomena. In support of the modeling, a detailed analysis was made of the "<span class="hlt">glacier</span> cooling effect" that reduces summer air temperature near the ice surface as compared to that predicted on the basis of a spatially uniform temperature lapse rate. The analysis was based on several years of data from measurements of near-surface air temperature on the <span class="hlt">glacier</span>. The 2006 and 2007 winter balances of South Cascade <span class="hlt">Glacier</span>, computed with this new, model-augmented methodology, were 2.61 and 3.41 mWE, respectively. The 2006 and 2007 summer balances were -4.20 and -3.63 mWE, respectively, and the 2006 and 2007 net balances were -1.59 and -0.22 mWE. PDF version of a presentation on the mass balance of South Cascade <span class="hlt">Glacier</span> in Washington state. Presented at the American Geophysical Union Fall Meeting 2010.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.C21B0736R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C21B0736R"><span>A new <span class="hlt">glacier</span> inventory for the Karakoram-Pamir region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rastner, P.; Paul, F.; Bolch, T.; Moelg, N.</p> <p>2015-12-01</p> <p>High-quality <span class="hlt">glacier</span> inventories are required as a reference dataset to determine <span class="hlt">glacier</span> changes and model their reaction to climate change, among others. In particular in High Mountain Asia such an inventory was missing for several heavily <span class="hlt">glacierized</span> regions with reportedly strongly changing <span class="hlt">glaciers</span>. As a contribution to GLIMS and the Randolph <span class="hlt">Glacier</span> Inventory (RGI) we have mapped all <span class="hlt">glaciers</span> in the Karakoram and Pamir region within the framework of ESAs <span class="hlt">Glaciers</span>_cci project. <span class="hlt">Glacier</span> mapping was performed using the band ratio method (TM3/TM5) and manual editing of Landsat TM/ETM+ imagery acquired around the year 2000. The mapping was challenged by frequent seasonal snow at high elevations, debris-covered <span class="hlt">glacier</span> tongues, and several surging <span class="hlt">glaciers</span>. We addressed the snow issue by utilizing multi-temporal imagery and improved manual mapping of debris-covered <span class="hlt">glacier</span> tongues with ALOS PALSAR coherence images. Slow disintegration of <span class="hlt">glacier</span> tongues after a surge (leaving still-connected dead ice) results in a difficult identification of the terminus and assignment of entities. Drainage divides were derived from the ASTER GDEM II and manually corrected to calculate topographic parameters. All <span class="hlt">glaciers</span> larger 0.02 km2 cover an area of about 21,700 km2 in the Karakoram and about 11,800 km² in the Pamir region. Most <span class="hlt">glaciers</span> are in the 0.1-0.5 km2 size class for Pamir, whereas for the Karakoram they are in the class <0.1 km2. <span class="hlt">Glaciers</span> between 1 and 5 km2 contribute more than 30% to the total area in Pamir, whereas for the Karakoram region it is only 17%. The mean <span class="hlt">glacier</span> elevation in the Karakoram (Pamir) region is 5426 (4874) m. A comparison with other recently published inventories reveals differences in the interpretation of <span class="hlt">glacier</span> extents (mainly in the accumulation region) that would lead to large area changes if unconsidered for change assessment across different inventories.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.7488S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.7488S"><span>Integrated firn elevation change model for <span class="hlt">glaciers</span> and ice caps</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saß, Björn; Sauter, Tobias; Braun, Matthias</p> <p>2016-04-01</p> <p>We present the development of a firn compaction model in order to improve the volume to mass conversion of geodetic <span class="hlt">glacier</span> mass balance measurements. The model is applied on the Arctic ice cap Vestfonna. Vestfonna is located on the island Nordaustlandet in the north <span class="hlt">east</span> of Svalbard. Vestfonna covers about 2400 km² and has a dome like shape with well-defined outlet <span class="hlt">glaciers</span>. Elevation and volume changes measured by e.g. satellite techniques are becoming more and more popular. They are carried out over observation periods of variable length and often covering different meteorological and snow hydrological regimes. The elevation change measurements compose of various components including dynamic adjustments, firn compaction and mass loss by downwasting. Currently, geodetic <span class="hlt">glacier</span> mass balances are frequently converted from elevation change measurements using a constant conversion factor of 850 kg m-³ or the density of ice (917 kg m-³) for entire <span class="hlt">glacier</span> basins. However, the natural conditions are rarely that static. Other studies used constant densities for the ablation (900 kg m-³) and accumulation (600 kg m-³) areas, whereby density variations with varying meteorological and climate conditions are not considered. Hence, each approach bears additional uncertainties from the volume to mass conversion that are strongly affected by the type and timing of the repeat measurements. We link and adapt existing models of surface energy balance, accumulation and snow and firn processes in order to improve the volume to mass conversion by considering the firn compaction component. Energy exchange at the surface is computed by a surface energy balance approach and driven by meteorological variables like incoming short-wave radiation, air temperature, relative humidity, air pressure, wind speed, all-phase precipitation, and cloud cover fraction. Snow and firn processes are addressed by a coupled subsurface model, implemented with a non-equidistant layer discretisation. On</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/0387b/report.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/0387b/report.pdf"><span>Recent Activity of <span class="hlt">Glaciers</span> of Mount Rainier, Washington</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sigafoos, Robert S.; Hendricks, E.L.</p> <p>1972-01-01</p> <p>Knowing the ages of trees growing on recent moraines at Mount Rainier, Wash., permits the moraines to be dated. Moraines which are ridges of boulders, gravel, sand, and dust deposited at the margins of a <span class="hlt">glacier</span>, mark former limits of a receding <span class="hlt">glacier</span>. Knowing past glacial activity aids our understanding of past climatic variations. The report documents the ages of moraines deposited by eight <span class="hlt">glaciers</span>. Aerial photographs and planimetric maps show areas where detailed field studies were made below seven <span class="hlt">glaciers</span>. Moraines, past ice positions, and sample areas are plotted on the photographs and maps, along with trails, roads, streams, and landforms, to permit critical areas to be identified in the future. Ground photographs are included so that sample sites and easily accessible moraines can be found along trails. Tables present data about trees sampled in areas near the <span class="hlt">glaciers</span> of Mount Rainier, Wash. The data in the tables show there are modern moraines of different age around the mountain; some valleys contain only one modern moraiine; others contain as many as nine. The evidence indicates a sequence of modern glacial advances terminating at about the following A.D. dates: 1525, 1550, 1625-60, 1715, 1730-65, 1820-60, 1875, and 1910. Nisqually River valley near Nisqually <span class="hlt">Glacier</span> contains one moraine formed before A.D. 1842; Tahoma Creek valley near South Tahoma <span class="hlt">Glacier</span> contains three moraines formed before A.D. 1528; 1843, and 1864; South Puyallup River valley near Tahoma <span class="hlt">Glacier</span>, six moraines A.D. 1544, 1761, 1841, 1851, 1863, 1898; Puyallup <span class="hlt">Glacier</span>, one moraine, A.D. 1846; Carbon <span class="hlt">Glacier</span>, four moraines, 1519, 1763, 1847, 1876; Winthrop <span class="hlt">Glacier</span>, four moraines, 1655, 1716, 1760, amid 1822; Emmons <span class="hlt">Glacier</span>, nine moraines, 1596, 1613, 1661, 1738, 1825, 1850, 1865, 1870, 1901; and Ohanapecosh <span class="hlt">Glacier</span>, three moraines, 1741, 1846, and 1878. Abandoned melt-water and flood channels were identified within moraine complexes below three <span class="hlt">glaciers</span>, and their time of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.C23C0669E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.C23C0669E"><span>Satellite Observations of <span class="hlt">Glacier</span> Surface Velocities in Southeast Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Elliott, J.; Melkonian, A. K.; Pritchard, M. E.</p> <p>2012-12-01</p> <p><span class="hlt">Glaciers</span> in southeast Alaska are undergoing rapid changes and are significant contributors to sea level rise. A key to understanding the ice dynamics is knowledge of the surface velocities, which can be used with ice thickness measurements to derive mass flux rates. For many <span class="hlt">glaciers</span> in Alaska, surface velocity estimates either do not exist or are based on data that are at least a decade old. Here we present updated maps of <span class="hlt">glacier</span> surface velocities in southeast Alaska produced through a pixel tracking technique using synthetic aperture radar data and high-resolution optical imagery. For <span class="hlt">glaciers</span> with previous velocity estimates, we will compare the results and discuss possible implications for ice dynamics. We focus on <span class="hlt">Glacier</span> Bay and the Stikine Icefield, which contain a number of fast-flowing tidewater <span class="hlt">glaciers</span> including LeConte, Johns Hopkins, and La Perouse. For the Johns Hopkins, we will also examine the influence a massive landslide in June 2012 had on flow dynamics. Our velocity maps show that within <span class="hlt">Glacier</span> Bay, the highest surface velocities occur on the tidewater <span class="hlt">glaciers</span>. La Perouse, the only <span class="hlt">Glacier</span> Bay <span class="hlt">glacier</span> to calve directly into the Pacific Ocean, has maximum velocities of 3.5 - 4 m/day. Johns Hopkins <span class="hlt">Glacier</span> shows 4 m/day velocities at both its terminus and in its upper reaches, with lower velocities of ~1-3 m/day in between those two regions. Further north, the Margerie <span class="hlt">Glacier</span> has a maximum velocity of ~ 4.5 m/day in its upper reaches and a velocity of ~ 2 m/day at its terminus. Along the Grand Pacific terminus, the western terminus fed by the Ferris <span class="hlt">Glacier</span> displays velocities of about 1 m/day while the eastern terminus has lower velocities of < 0.5 m/day. The lake terminating <span class="hlt">glaciers</span> along the Pacific coast have overall lower surface velocities, but they display complex flow patterns. The Alsek <span class="hlt">Glacier</span> displays maximum velocities of 2.5 m/day above where it divides into two branches. Velocities at the terminus of the northern branch reach 1</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('https://www.loc.gov/pictures/collection/hh/item/sd0078.photos.203513p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/sd0078.photos.203513p/"><span>4. Inside perimeter fence, view towards <span class="hlt">east</span> and launch closure, ...</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. Inside perimeter fence, view towards <span class="hlt">east</span> and launch closure, sensor EMP antenna left center - Ellsworth Air Force Base, Delta Flight, Launch Facility D-6, 4 miles north of Badlands National <span class="hlt">Park</span> Headquarters, 4.5 miles <span class="hlt">east</span> of Jackson County line on county road, Interior, Jackson County, SD</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C11A0888M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C11A0888M"><span>Characterization of meltwater 'ingredients' at the Haig <span class="hlt">Glacier</span>, Canadian Rockies: the importance of <span class="hlt">glaciers</span> to regional water resources</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Miller, K.; Marshall, S.</p> <p>2017-12-01</p> <p>With rising temperatures, Alberta's <span class="hlt">glaciers</span> are under stresses which change and alter the timing, amount, and composition of meltwater contributions to rivers that flow from the Rocky Mountains. Meltwater can be stored within a <span class="hlt">glacier</span> or it can drain through the groundwater system, reducing and delaying meltwater delivery to <span class="hlt">glacier</span>-fed streams. This study tests whether the <span class="hlt">glacier</span> meltwater is chemically distinct from rain or snow melt, and thus whether meltwater contributions to higher-order streams that flow from the mountains can be determined through stream chemistry. Rivers like the Bow, North Saskatchewan, and Athabasca are vital waterways for much of Alberta's population. Assessing the extent of <span class="hlt">glacier</span> meltwater is vital to future water resource planning. <span class="hlt">Glacier</span> snow/ice and meltwater stream samples were collected during the 2017 summer melt season (May- September) and analyzed for isotope and ion chemistry. The results are being used to model water chemistry evolution in the melt stream through the summer season. A chemical mixing model will be constructed to determine the fractional contributions to the Haig meltwater stream from precipitation, surface melt, and subglacial meltwaters. Distinct chemical water signatures have not been used to partition water sources and understand <span class="hlt">glacier</span> contributions to rivers in the Rockies. The goal of this work is to use chemical signatures of glacial meltwater to help assess the extent of <span class="hlt">glacier</span> meltwater in Alberta rivers and how this varies through the summer season.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001488.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001488.html"><span><span class="hlt">Glaciers</span> and Sea Level Rise</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-12-08</p> <p>Small valley <span class="hlt">glacier</span> exiting the Devon Island Ice Cap in Canada. To learn about the contributions of <span class="hlt">glaciers</span> to sea level rise, visit: www.nasa.gov/topics/earth/features/<span class="hlt">glacier</span>-sea-rise.html Credit: Alex Gardner, Clark University 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/pp/1746/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/pp/1746/"><span>Geographic Names of Iceland's <span class="hlt">Glaciers</span>: Historic and Modern</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Sigurðsson, Oddur; Williams, Richard S.</p> <p>2008-01-01</p> <p>Climatic changes and resulting <span class="hlt">glacier</span> fluctuations alter landscapes. In the past, such changes were noted by local residents who often documented them in historic annals; eventually, <span class="hlt">glacier</span> variations were recorded on maps and scientific reports. In Iceland, 10 <span class="hlt">glacier</span> place-names are to be found in Icelandic sagas, and one of Iceland's ice caps, Snaefellsjokull, appeared on maps of Iceland published in the 16th century. In the late 17th century, the first description of eight of Iceland's <span class="hlt">glaciers</span> was written. Therefore, Iceland distinguishes itself in having a more than 300-year history of observations by Icelanders on its <span class="hlt">glaciers</span>. A long-term collaboration between Oddur Sigurdsson and Richard S. Williams, Jr., led to the authorship of three books on the <span class="hlt">glaciers</span> of Iceland. Much effort has been devoted to documenting historical <span class="hlt">glacier</span> research and related nomenclature and to physical descriptions of Icelandic <span class="hlt">glaciers</span> by Icelanders and other scientists from as far back as the Saga Age to recent (2008) times. The first book, Icelandic Ice Mountains, was published by the Icelandic Literary Society in 2004 in cooperation with the Icelandic Glaciological Society and the International Glaciological Society. Icelandic Ice Mountains was a <span class="hlt">glacier</span> treatise written by Sveinn Palsson in 1795 and is the first English translation of this important scientific document. Icelandic Ice Mountains includes a Preface, including a summary of the history and facsimiles of page(s) from the original manuscript, a handwritten copy, and an 1815 manuscript (without maps and drawings) by Sveinn Palsson on the same subject which he wrote for Rev. Ebenezer Henderson; an Editor's Introduction; 82 figures, including facsimiles of Sveinn Palsson's original maps and perspective drawings, maps, and photographs to illustrate the text; a comprehensive Index of Geographic Place-Names and Other Names in the treatise; References, and 415 Endnotes. Professional Paper 1746 (this book) is the second</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/al1194.photos.226340p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/al1194.photos.226340p/"><span>View of <span class="hlt">parking</span> (resting) frame that supported the Shuttle assembly ...</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>View of <span class="hlt">parking</span> (resting) frame that supported the Shuttle assembly when the hydrodynamic supports were not engaged (removed from structure). - Marshall Space Flight Center, Saturn V Dynamic Test Facility, <span class="hlt">East</span> Test Area, Huntsville, Madison County, AL</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Remote+AND+sensing&pg=6&id=EJ343018','ERIC'); return false;" href="https://eric.ed.gov/?q=Remote+AND+sensing&pg=6&id=EJ343018"><span>Get Close to <span class="hlt">Glaciers</span> with Satellite Imagery.</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>Hall, Dorothy K.</p> <p>1986-01-01</p> <p>Discusses the use of remote sensing from satellites to monitor <span class="hlt">glaciers</span>. Discusses efforts to use remote sensing satellites of the Landsat series for examining the global distribution, mass, balance, movements, and dynamics of the world's <span class="hlt">glaciers</span>. Includes several Landsat images of various <span class="hlt">glaciers</span>. (TW)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C42A..04C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C42A..04C"><span>An Active Englacial Hydrological System in a Cold <span class="hlt">Glacier</span>: Blood Falls, Taylor <span class="hlt">Glacier</span>, Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Carr, C. G.; Pettit, E. C.; Carmichael, J.; Badgeley, J.; Tulaczyk, S. M.; Lyons, W. B.; Mikucki, J.</p> <p>2016-12-01</p> <p>Blood Falls is a supraglacial hydrological feature formed by episodic release of iron-rich subglacial brine derived from an extensive aquifer beneath the cold, polar, Taylor <span class="hlt">Glacier</span>. While fluid transport in non-temperate ice typically occurs through meltwater delivery from the <span class="hlt">glacier</span> surface to the bed (hydrofracturing, supraglacial lake drainage), Blood Falls represents the opposite situation: brine moves from a subglacial source to the <span class="hlt">glacier</span> surface. Here, we present the first complete conceptual model for brine transport and release, as well as the first direct evidence of a wintertime brine release at Blood Falls obtained through year-round time-lapse photography. Related analyses show that brine pools subglacially underneath the northern terminus of Taylor <span class="hlt">Glacier</span>, rather than flowing directly into proglacial Lake Bonney because ice-cored moraines and channelized surface topography provide hydraulic barriers. This pooled brine is pressurized by hydraulic head from the upglacier brine source region. Based on seismic data, we propose that episodic supraglacial release is initiated by high strain rates coupled with pressurized subglacial brine that drive intermittent subglacial and englacial fracturing. Ultimately, brine-filled basal crevasses propagate upward to link with surface crevasses, allowing brine to flow from the bed to the surface. The observation of wintertime brine release indicates that surface-generated meltwater is not necessary to trigger crack propagation or to maintain the conduit as previously suggested. The liquid brine persists beneath and within the cold ice (-17°C) despite ambient ice/brine temperature differences of as high as 10°C through both locally depressed brine freezing temperatures through cryoconcentration of salts and increased ice temperatures through release of latent heat during partial freezing of brine. The existence of an englacial hydrological system initiated by basal crevassing extends to polar <span class="hlt">glaciers</span> a process</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://files.eric.ed.gov/fulltext/ED132796.pdf','ERIC'); return false;" href="http://files.eric.ed.gov/fulltext/ED132796.pdf"><span>Nassau County Department of Recreation and <span class="hlt">Parks</span>.</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>Iowa Univ., Iowa City. Recreation Education Program.</p> <p></p> <p>Presented are duplications of the responses given by the Nassau County Department of Recreation and <span class="hlt">Parks</span> (<span class="hlt">East</span> Meadow, New York) as part of a project to collect, share, and compile information about, and techniques in the operation of 18 community action models for recreation services to the disabled. Model programs are categorized as consumer,…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.C33E0863M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.C33E0863M"><span>Hypsometric control on <span class="hlt">glacier</span> mass balance sensitivity in Alaska</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McGrath, D.; Sass, L.; Arendt, A. A.; O'Neel, S.; Kienholz, C.; Larsen, C.; Burgess, E. W.</p> <p>2015-12-01</p> <p>Mass loss from <span class="hlt">glaciers</span> in Alaska is dominated by strongly negative surface balances, particularly on small, continental <span class="hlt">glaciers</span> but can be highly variable from <span class="hlt">glacier</span> to <span class="hlt">glacier</span>. <span class="hlt">Glacier</span> hypsometry can exert significant control on mass balance sensitivity, particularly if the equilibrium line altitude (ELA) is in a broad area of low surface slope. In this study, we explore the spatial variability in <span class="hlt">glacier</span> response to future climate forcings on the basis of hypsometry. We first derive mass balance sensitivities (30-70 m ELA / 1° C and 40-90 m ELA / 50% decrease in snow accumulation) from the ~50-year USGS Benchmark <span class="hlt">glaciers</span> mass balance record. We subsequently assess mean climate fields in 2090-2100 derived from the IPCC AR5/CMIP5 RCP 6.0 5-model mean. Over <span class="hlt">glaciers</span> in Alaska, we find 2-4° C warming and 10-20% increase in precipitation relative to 2006-2015, but a corresponding 0-50% decrease in snow accumulation due to rising temperatures. We assess changes in accumulation area ratios (AAR) to a rising ELA using binned individual <span class="hlt">glacier</span> hypsometries. For an ELA increase of 150 m, the mean statewide AAR drops by 0.45, representing a 70% reduction in accumulation area on an individual <span class="hlt">glacier</span> basis. Small, interior <span class="hlt">glaciers</span> are the primary drivers of this reduction and for nearly 25% of all <span class="hlt">glaciers</span>, the new ELA exceeds the <span class="hlt">glacier</span>'s maximum elevation, portending eventual loss. The loss of small <span class="hlt">glaciers</span>, particularly in the drier interior of Alaska will significantly modify streamflow properties (flashy hydrographs, earlier and reduced peak flows, increased interannual variability, warmer temperatures) with poorly understood downstream ecosystem and oceanographic impacts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018TCry...12...81T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018TCry...12...81T"><span>The Greater Caucasus <span class="hlt">Glacier</span> Inventory (Russia, Georgia and Azerbaijan)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tielidze, Levan G.; Wheate, Roger D.</p> <p>2018-01-01</p> <p>There have been numerous studies of <span class="hlt">glaciers</span> in the Greater Caucasus, but none that have generated a modern <span class="hlt">glacier</span> database across the whole mountain range. Here, we present an updated and expanded <span class="hlt">glacier</span> inventory at three time periods (1960, 1986, 2014) covering the entire Greater Caucasus. Large-scale topographic maps and satellite imagery (Corona, Landsat 5, Landsat 8 and ASTER) were used to conduct a remote-sensing survey of <span class="hlt">glacier</span> change, and the 30 m resolution Advanced Spaceborne Thermal Emission and Reflection Radiometer Global Digital Elevation Model (ASTER GDEM; 17 November 2011) was used to determine the aspect, slope and height distribution of <span class="hlt">glaciers</span>. <span class="hlt">Glacier</span> margins were mapped manually and reveal that in 1960 the mountains contained 2349 <span class="hlt">glaciers</span> with a total <span class="hlt">glacier</span> surface area of 1674.9 ± 70.4 km2. By 1986, <span class="hlt">glacier</span> surface area had decreased to 1482.1 ± 64.4 km2 (2209 <span class="hlt">glaciers</span>), and by 2014 to 1193.2 ± 54.0 km2 (2020 <span class="hlt">glaciers</span>). This represents a 28.8 ± 4.4 % (481 ± 21.2 km2) or 0.53 % yr-1 reduction in total <span class="hlt">glacier</span> surface area between 1960 and 2014 and an increase in the rate of area loss since 1986 (0.69 % yr-1) compared to 1960-1986 (0.44 % yr-1). <span class="hlt">Glacier</span> mean size decreased from 0.70 km2 in 1960 to 0.66 km2 in 1986 and to 0.57 km2 in 2014. This new <span class="hlt">glacier</span> inventory has been submitted to the Global Land Ice Measurements from Space (GLIMS) database and can be used as a basis data set for future studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMPP31C1293L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMPP31C1293L"><span>Southwest Greenland's Alpine <span class="hlt">Glacier</span> History: Recent <span class="hlt">Glacier</span> Change in the Context of the Holocene Geologic Record</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Larocca, L. J.; Axford, Y.; Lasher, G. E.; Lee, C. W.</p> <p>2017-12-01</p> <p>Due to anthropogenic climate change, the Arctic region is currently undergoing major transformation, and is expected to continue warming much faster than the global average. To put recent and future changes into context, a longer-term understanding of this region's past response to natural climate variability is needed. Given their sensitivity to modest climate change, small alpine <span class="hlt">glaciers</span> and ice caps on Greenland's coastal margin (beyond the Greenland Ice Sheet) represent ideal features to record climate variability through the Holocene. Here we investigate the Holocene history of a small ( 160 square km) ice cap and adjacent alpine <span class="hlt">glaciers</span>, located in southwest Greenland approximately 50 km south of Nuuk. We employ measurements on sediment cores from a <span class="hlt">glacier</span>-fed lake in combination with geospatial analysis of satellite images spanning the past several decades. Sedimentary indicators of sediment source and thus glacial activity, including organic matter abundance, inferred chlorophyll-a content, sediment major element abundances, grain size, and magnetic susceptibility are presented from cores collected from a distal <span class="hlt">glacier</span>-fed lake (informally referred to here as Per's Lake) in the summer of 2015. These parameters reflect changes in the amount and character of inorganic detrital input into the lake, which may be linked to the size of the upstream <span class="hlt">glaciers</span> and ice cap and allow us to reconstruct their status through the Holocene. Additionally, we present a complementary record of recent changes in Equilibrium Line Altitude (ELA) for the upstream alpine <span class="hlt">glaciers</span>. Modern ELAs are inferred using the accumulation area ratio (AAR) method in ArcGIS via Landsat and Worldview-2 satellite imagery, along with elevation data obtained from digital elevation models (DEMs). Paleo-ELAs are inferred from the positions of moraines and trim lines marking the <span class="hlt">glaciers</span>' most recent expanded state, which we attribute to the Little Ice Age (LIA). This approach will allow us to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5371072','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5371072"><span>Bacterial Microbiota Associated with the <span class="hlt">Glacier</span> Ice Worm Is Dominated by Both Worm-Specific and <span class="hlt">Glacier</span>-Derived Facultative Lineages</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Murakami, Takumi; Segawa, Takahiro; Dial, Roman; Takeuchi, Nozomu; Kohshima, Shiro; Hongoh, Yuichi</p> <p>2017-01-01</p> <p>The community structure of bacteria associated with the <span class="hlt">glacier</span> ice worm Mesenchytraeus solifugus was analyzed by amplicon sequencing of 16S rRNA genes and their transcripts. Ice worms were collected from two distinct <span class="hlt">glaciers</span> in Alaska, Harding Icefield and Byron <span class="hlt">Glacier</span>, and <span class="hlt">glacier</span> surfaces were also sampled for comparison. Marked differences were observed in bacterial community structures between the ice worm and <span class="hlt">glacier</span> surface samples. Several bacterial phylotypes were detected almost exclusively in the ice worms, and these bacteria were phylogenetically affiliated with either animal-associated lineages or, interestingly, clades mostly consisting of <span class="hlt">glacier</span>-indigenous species. The former included bacteria that belong to Mollicutes, Chlamydiae, Rickettsiales, and Lachnospiraceae, while the latter included Arcicella and Herminiimonas phylotypes. Among these bacteria enriched in ice worm samples, Mollicutes, Arcicella, and Herminiimonas phylotypes were abundantly and consistently detected in the ice worm samples; these phylotypes constituted the core microbiota associated with the ice worm. A fluorescence in situ hybridization analysis showed that Arcicella cells specifically colonized the epidermis of the ice worms. Other bacterial phylotypes detected in the ice worm samples were also abundantly recovered from the respective habitat <span class="hlt">glaciers</span>; these bacteria may be food for ice worms to digest or temporary residents. Nevertheless, some were overrepresented in the ice worm RNA samples; they may also function as facultative gut bacteria. Our results indicate that the community structure of bacteria associated with ice worms is distinct from that in the associated <span class="hlt">glacier</span> and includes worm-specific and facultative, <span class="hlt">glacier</span>-indigenous lineages. PMID:28302989</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28302989','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28302989"><span>Bacterial Microbiota Associated with the <span class="hlt">Glacier</span> Ice Worm Is Dominated by Both Worm-Specific and <span class="hlt">Glacier</span>-Derived Facultative Lineages.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Murakami, Takumi; Segawa, Takahiro; Dial, Roman; Takeuchi, Nozomu; Kohshima, Shiro; Hongoh, Yuichi</p> <p>2017-03-31</p> <p>The community structure of bacteria associated with the <span class="hlt">glacier</span> ice worm Mesenchytraeus solifugus was analyzed by amplicon sequencing of 16S rRNA genes and their transcripts. Ice worms were collected from two distinct <span class="hlt">glaciers</span> in Alaska, Harding Icefield and Byron <span class="hlt">Glacier</span>, and <span class="hlt">glacier</span> surfaces were also sampled for comparison. Marked differences were observed in bacterial community structures between the ice worm and <span class="hlt">glacier</span> surface samples. Several bacterial phylotypes were detected almost exclusively in the ice worms, and these bacteria were phylogenetically affiliated with either animal-associated lineages or, interestingly, clades mostly consisting of <span class="hlt">glacier</span>-indigenous species. The former included bacteria that belong to Mollicutes, Chlamydiae, Rickettsiales, and Lachnospiraceae, while the latter included Arcicella and Herminiimonas phylotypes. Among these bacteria enriched in ice worm samples, Mollicutes, Arcicella, and Herminiimonas phylotypes were abundantly and consistently detected in the ice worm samples; these phylotypes constituted the core microbiota associated with the ice worm. A fluorescence in situ hybridization analysis showed that Arcicella cells specifically colonized the epidermis of the ice worms. Other bacterial phylotypes detected in the ice worm samples were also abundantly recovered from the respective habitat <span class="hlt">glaciers</span>; these bacteria may be food for ice worms to digest or temporary residents. Nevertheless, some were overrepresented in the ice worm RNA samples; they may also function as facultative gut bacteria. Our results indicate that the community structure of bacteria associated with ice worms is distinct from that in the associated <span class="hlt">glacier</span> and includes worm-specific and facultative, <span class="hlt">glacier</span>-indigenous lineages.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26045157','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26045157"><span>Pyrosequencing-Based Assessment of the Microbial Community Structure of Pastoruri <span class="hlt">Glacier</span> Area (Huascarán National <span class="hlt">Park</span>, Perú), a Natural Extreme Acidic Environment.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>González-Toril, Elena; Santofimia, Esther; Blanco, Yolanda; López-Pamo, Enrique; Gómez, Manuel J; Bobadilla, Miguel; Cruz, Rolando; Palomino, Edwin Julio; Aguilera, Ángeles</p> <p>2015-11-01</p> <p>The exposure of fresh sulfide-rich lithologies by the retracement of the Nevado Pastoruri <span class="hlt">glacier</span> (Central Andes, Perú) is increasing the presence of heavy metals in the water as well as decreasing the pH, producing an acid rock drainage (ARD) process in the area. We describe the microbial communities of an extreme ARD site in Huascarán National <span class="hlt">Park</span> as well as their correlation with the water physicochemistry. Microbial biodiversity was analyzed by FLX 454 sequencing of the 16S rRNA gene. The suggested geomicrobiological model of the area distinguishes three different zones. The proglacial zone is located in the upper part of the valley, where the ARD process is not evident yet. Most of the OTUs detected in this area were related to sequences associated with cold environments (i.e., psychrotolerant species of Cyanobacteria or Bacteroidetes). After the proglacial area, an ARD-influenced zone appeared, characterized by the presence of phylotypes related to acidophiles (Acidiphilium) as well as other species related to acidic and cold environments (i.e., acidophilic species of Chloroflexi, Clostridium and Verrumicrobia). Sulfur- and iron-oxidizing acidophilic bacteria (Acidithiobacillus) were also identified. The post-ARD area was characterized by the presence of OTUs related to microorganisms detected in soils, permafrost, high mountain environments, and deglaciation areas (Sphingomonadales, Caulobacter or Comamonadaceae).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA03386&hterms=Glacier+retreat+global&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DGlacier%2Bretreat%2Bglobal','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA03386&hterms=Glacier+retreat+global&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DGlacier%2Bretreat%2Bglobal"><span>Malaspina <span class="hlt">Glacier</span>, Alaska, Perspective with Landsat Overlay</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2003-01-01</p> <p><p/>Malaspina <span class="hlt">Glacier</span> in southeastern Alaska is considered the classic example of a piedmont <span class="hlt">glacier</span>. Piedmont <span class="hlt">glaciers</span> occur where valley <span class="hlt">glaciers</span> exit a mountain range onto broad lowlands, are no longer laterally confined, and spread to become wide lobes. Malaspina <span class="hlt">Glacier</span> is actually a compound <span class="hlt">glacier</span>, formed by the merger of several valley <span class="hlt">glaciers</span>, the most prominent of which seen here are Agassiz <span class="hlt">Glacier</span> (left) and Seward <span class="hlt">Glacier</span> (right). In total, Malaspina <span class="hlt">Glacier</span> is up to 65 kilometers (40 miles) wide and extends up to 45 kilometers (28 miles) from the mountain front nearly to the sea. <p/>This perspective view was created from a Landsat satellite image and an elevation model generated by the Shuttle Radar Topography Mission (SRTM). Landsat views both visible and infrared light, which have been combined here into a color composite that generally shows glacial ice in light blue, snow in white, vegetation in green, bare rock in grays and tans, and the ocean (foreground) in dark blue. The back (northern) edge of the data set forms a false horizon that meets a false sky. <p/><span class="hlt">Glaciers</span> erode rocks, carry them down slope, and deposit them at the edge of the melting ice, typically in elongated piles called moraines. The moraine patterns at Malaspina <span class="hlt">Glacier</span> are quite spectacular in that they have huge contortions that result from the <span class="hlt">glacier</span> crinkling as it gets pushed from behind by the faster-moving valley <span class="hlt">glaciers</span>. <p/><span class="hlt">Glaciers</span> are sensitive indicators of climatic change. They can grow and thicken with increasing snowfall and/or decreased melting. Conversely, they can retreat and thin if snowfall decreases and/or atmospheric temperatures rise and cause increased melting. Landsat imaging has been an excellent tool for mapping the changing geographic extent of <span class="hlt">glaciers</span> since 1972. The elevation measurements taken by SRTM in February 2000 now provide a near-global baseline against which future non-polar region glacial thinning or thickening can be assessed. <p</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.A11G0090G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.A11G0090G"><span>The Association Between Fog and Temperature Inversions from Ground and Radiosonde Observations in <span class="hlt">East</span> Greenland</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gilson, G.; Jiskoot, H.</p> <p>2016-12-01</p> <p>Many Arctic <span class="hlt">glaciers</span> terminate along coasts where temperature inversions and sea fog are frequent during summer. Both can influence <span class="hlt">glacier</span> ablation, but the effects of fog may be complex. To understand fog's physical and radiative properties and its association to temperature inversions it is important to determine accurate Arctic coastal fog climatologies In previous research we determined that fog in <span class="hlt">East</span> Greenland peaks in the melt season and can be spatially extensive over <span class="hlt">glacierized</span> terrain. In this study we aim to understand which environmental factors influence fog occurrence in <span class="hlt">East</span> Greenland; understand the association between fog and temperature inversions; and quantify fog height. We analyzed fog observations and other weather data from coastal synoptic weather stations, and extracted temperature inversions from the Integrated Global Radiosonde Archive radiosonde profiles. Fog height was calculated from radiosonde profiles, based on a method developed for radiation fog which we expanded to include advection and steam fog. Our results show that Arctic coastal fog requires sea ice breakup and a sea breeze with wind speed between 1-4 m/s. Fog is mostly advective, occurring under stable synoptic conditions characterized by deep and strong low-level temperature inversions. Steam fog may occur 5-30% of the time. Fog can occur under near-surface subsidence, with a subsaturated inversion base, or a saturated inversion base. We classified five types of fog based on their vertical sounding characteristics: only at the surface, below an inversion, capped by an inversion, inside a surface-based inversion, or inside a low-level inversion. Fog is commonly 100-400 m thick, often reaching the top of the boundary layer. Fog height is greater at northern stations, where daily fog duration is longer and relative humidity lower. Our results will be included in <span class="hlt">glacier</span> energy-balance models to account for the influence of fog and temperature inversions on <span class="hlt">glacier</span> melt.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-ED04-0056-098.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-ED04-0056-098.html"><span>A fox at Torres del Paine National <span class="hlt">Park</span> in Chile during NASA's AirSAR 2004 campaign</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2004-03-11</p> <p>A fox at Torres del Paine National <span class="hlt">Park</span> in Chile during NASA's AirSAR 2004 campaign. AirSAR 2004 is a three-week expedition in Central and South America by an international team of scientists that is using an all-weather imaging tool, called the Airborne Synthetic Aperture Radar (AirSAR), located onboard NASA's DC-8 airborne laboratory. Scientists from many parts of the world are combining ground research with NASA's AirSAR technology to improve and expand on the quality of research they are able to conduct. Founded in 1959, Torres del Paine National <span class="hlt">Park</span> encompasses 450,000 acres in the Patagonia region of Chile. This region is being studied by NASA using a DC-8 equipped with an Airborne Synthetic Aperture Radar (AirSAR) developed by scientists from NASA’s Jet Propulsion Laboratory. This is a very sensitive region that is important to scientists because the temperature has been consistently rising causing a subsequent melting of the region’s <span class="hlt">glaciers</span>. AirSAR will provide a baseline model and unprecedented mapping of the region. This data will make it possible to determine whether the warming trend is slowing, continuing or accelerating. AirSAR will also provide reliable information on ice shelf thickness to measure the contribution of the <span class="hlt">glaciers</span> to sea level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ESuD....5..493M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ESuD....5..493M"><span>Pluri-decadal (1955-2014) evolution of <span class="hlt">glacier</span>-rock <span class="hlt">glacier</span> transitional landforms in the central Andes of Chile (30-33° S)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Monnier, Sébastien; Kinnard, Christophe</p> <p>2017-08-01</p> <p>Three <span class="hlt">glacier</span>-rock <span class="hlt">glacier</span> transitional landforms in the central Andes of Chile are investigated over the last decades in order to highlight and question the significance of their landscape and flow dynamics. Historical (1955-2000) aerial photos and contemporary (> 2000) Geoeye satellite images were used together with common processing operations, including imagery orthorectification, digital elevation model generation, and image feature tracking. At each site, the rock <span class="hlt">glacier</span> morphology area, thermokarst area, elevation changes, and horizontal surface displacements were mapped. The evolution of the landforms over the study period is remarkable, with rapid landscape changes, particularly an expansion of rock <span class="hlt">glacier</span> morphology areas. Elevation changes were heterogeneous, especially in debris-covered <span class="hlt">glacier</span> areas with large heaving or lowering up to more than ±1 m yr-1. The use of image feature tracking highlighted spatially coherent flow vector patterns over rock <span class="hlt">glacier</span> areas and, at two of the three sites, their expansion over the studied period; debris-covered <span class="hlt">glacier</span> areas are characterized by a lack of movement detection and/or chaotic displacement patterns reflecting thermokarst degradation; mean landform displacement speeds ranged between 0.50 and 1.10 m yr-1 and exhibited a decreasing trend over the studied period. One important highlight of this study is that, especially in persisting cold conditions, rock <span class="hlt">glaciers</span> can develop upward at the expense of debris-covered <span class="hlt">glaciers</span>. Two of the studied landforms initially (prior to the study period) developed from an alternation between glacial advances and rock <span class="hlt">glacier</span> development phases. The other landform is a small debris-covered <span class="hlt">glacier</span> having evolved into a rock <span class="hlt">glacier</span> over the last half-century. Based on these results it is proposed that morphological and dynamical interactions between <span class="hlt">glaciers</span> and permafrost and their resulting hybrid landscapes may enhance the resilience of the mountain cryosphere</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=10532&hterms=glacier+melt&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dglacier%2Bmelt','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=10532&hterms=glacier+melt&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dglacier%2Bmelt"><span>Alaska <span class="hlt">Glaciers</span> and Rivers</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2007-01-01</p> <p>The Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA's Terra satellite captured this image on October 7, 2007, showing the Alaska Mountains of south-central Alaska already coated with snow. Purple shadows hang in the lee of the peaks, giving the snow-clad land a crumpled appearance. White gives way to brown on the right side of the image where the mountains yield to the lower-elevation Susitna River Valley. The river itself cuts a silver, winding path through deep green forests and brown wetlands and tundra. Extending from the river valley, are smaller rivers that originated in the Alaska Mountains. The source of these rivers is evident in the image. Smooth white tongues of ice extend into the river valleys, the remnants of the <span class="hlt">glaciers</span> that carved the valleys into the land. Most of the water flowing into the Gulf of Alaska from the Susitna River comes from these mountain <span class="hlt">glaciers</span>. <span class="hlt">Glacier</span> melt also feeds <span class="hlt">glacier</span> lakes, only one of which is large enough to be visible in this image. Immediately left of the Kahiltna River, the aquamarine waters of Chelatna Lake stand out starkly against the brown and white landscape.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70033573','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70033573"><span>Passive microwave (SSM/I) satellite predictions of valley <span class="hlt">glacier</span> hydrology, Matanuska <span class="hlt">Glacier</span>, Alaska</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Kopczynski, S.E.; Ramage, J.; Lawson, D.; Goetz, S.; Evenson, E.; Denner, J.; Larson, G.</p> <p>2008-01-01</p> <p>We advance an approach to use satellite passive microwave observations to track valley <span class="hlt">glacier</span> snowmelt and predict timing of spring snowmelt-induced floods at the terminus. Using 37 V GHz brightness temperatures (Tb) from the Special Sensor Microwave hnager (SSM/I), we monitor snowmelt onset when both Tb and the difference between the ascending and descending overpasses exceed fixed thresholds established for Matanuska <span class="hlt">Glacier</span>. Melt is confirmed by ground-measured air temperature and snow-wetness, while <span class="hlt">glacier</span> hydrologic responses are monitored by a stream gauge, suspended-sediment sensors and terminus ice velocity measurements. Accumulation area snowmelt timing is correlated (R2 = 0.61) to timing of the annual snowmelt flood peak and can be predicted within ??5 days. Copyright 2008 by the American Geophysical Union.</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('http://adsabs.harvard.edu/abs/2003AGUFM.C11C0845H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003AGUFM.C11C0845H"><span>Comparative Analysis of <span class="hlt">Glaciers</span> in the Chugach-St.-Elias Mountains</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Herzfeld, U. C.; Mayer, H.</p> <p>2003-12-01</p> <p>The phenomenon of <span class="hlt">glacier</span> surges has to date been studied for only relatively few examples. 136 of the 204 surge-type <span class="hlt">glaciers</span> in North America listed by Post (1969) are located in the St. Elias Mountains. In August 2003 we increased our data inventory of observations on surge <span class="hlt">glaciers</span> by collecting material for 19 <span class="hlt">glaciers</span> in the <span class="hlt">Glacier</span> Bay area and neighboring regions in the eastern St. Elias Mountains, including 6 surge-type <span class="hlt">glaciers</span> (Carroll, Rendu, Ferris, Grand Pacific, Margerie, and Johns Hopkins <span class="hlt">Glaciers</span>). Analyses utilize digital video and photographic data, satellite data and GPS data. Geostatistical classification parameters and algebraic parameters characteristic of surge motions are derived for selected <span class="hlt">glaciers</span>. During the 1993-1995 surge of Bering <span class="hlt">Glacier</span> the entire surface of Alaska's longest <span class="hlt">glacier</span> was crevassed and could be segmented into several dynamic provinces, where patterns changed as the surge progressed and the affected areas expanded downglacier and upglacier, finally affecting the Bagley Ice Field. The middle moraine of Grand Pacific and Ferris <span class="hlt">Glaciers</span> is pushed over to the Grand Pacific side, caused by a recent surge of the heavily crevassed Ferris <span class="hlt">Glacier</span>. The front of Johns Hopkins <span class="hlt">Glacier</span> advances, as its lower reaches are affected by a surge. The surge history of Bering <span class="hlt">Glacier</span> goes back to the Holocene, whereas Carroll and Rendu <span class="hlt">Glaciers</span> have surged only 3-4 times. These observations pose questions on the possible relationship between surge dynamics and climatic changes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1160.pdf','CFR2014'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2014-title36-vol1/pdf/CFR-2014-title36-vol1-sec13-1160.pdf"><span>36 CFR 13.1160 - Restrictions on vessel entry.</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>2014-07-01</p> <p>... INTERIOR NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-<span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve... vessels in this subpart as required to protect the values and purposes of <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and... used only to directly exit <span class="hlt">Glacier</span> Bay from Bartlett Cove and return directly to Bartlett Cove. The...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2011-title36-vol1/pdf/CFR-2011-title36-vol1-sec13-1160.pdf','CFR2011'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2011-title36-vol1/pdf/CFR-2011-title36-vol1-sec13-1160.pdf"><span>36 CFR 13.1160 - Restrictions on vessel entry.</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>... INTERIOR NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-<span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve... vessels in this subpart as required to protect the values and purposes of <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and... used only to directly exit <span class="hlt">Glacier</span> Bay from Bartlett Cove and return directly to Bartlett Cove. The...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.gpo.gov/fdsys/pkg/CFR-2012-title36-vol1/pdf/CFR-2012-title36-vol1-sec13-1160.pdf','CFR2012'); return false;" href="https://www.gpo.gov/fdsys/pkg/CFR-2012-title36-vol1/pdf/CFR-2012-title36-vol1-sec13-1160.pdf"><span>36 CFR 13.1160 - Restrictions on vessel entry.</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>... INTERIOR NATIONAL <span class="hlt">PARK</span> SYSTEM UNITS IN ALASKA Special Regulations-<span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve... vessels in this subpart as required to protect the values and purposes of <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and... used only to directly exit <span class="hlt">Glacier</span> Bay from Bartlett Cove and return directly to Bartlett Cove. The...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA217638','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA217638"><span>Airfields on Antarctic <span class="hlt">Glacier</span> Ice</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1989-12-01</p> <p><span class="hlt">glacier</span> ice Vu., vA2 2~ FEB 0C DLSPM ONSAEM- T r it Cover: Blue ice areas near the Scott <span class="hlt">Glacier</span>. There is a possible landing field at 86035"S, 148025"W...pi. Ii7 t E 9 v 1.. - Site$ At Moliunt HoWe t87*20S. 14W 0W) -nd P-411 lardain t leois lower than that of clear <span class="hlt">Glacier</span> (85ൎ’S, 16795T~) wur-a...emphasis much more vigorous than isthecasein thehighin- on the area of Mount Howe and D’Angelo Bluff teior of Antarctica. For example, near Mawson</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19544737','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19544737"><span>Assessment of lake sensitivity to acidic deposition in national <span class="hlt">parks</span> of the Rocky Mountains.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nanus, L; Williams, M W; Campbell, D H; Tonnessen, K A; Blett, T; Clow, D W</p> <p>2009-06-01</p> <p>The sensitivity of high-elevation lakes to acidic deposition was evaluated in five national <span class="hlt">parks</span> of the Rocky Mountains based on statistical relations between lake acid-neutralizing capacity concentrations and basin characteristics. Acid-neutralizing capacity (ANC) of 151 lakes sampled during synoptic surveys and basin-characteristic information derived from geographic information system (GIS) data sets were used to calibrate the statistical models. The explanatory basin variables that were considered included topographic parameters, bedrock type, and vegetation type. A logistic regression model was developed, and modeling results were cross-validated through lake sampling during fall 2004 at 58 lakes. The model was applied to lake basins greater than 1 ha in area in <span class="hlt">Glacier</span> National <span class="hlt">Park</span> (n = 244 lakes), Grand Teton National <span class="hlt">Park</span> (n = 106 lakes), Great Sand Dunes National <span class="hlt">Park</span> and Preserve (n = 11 lakes), Rocky Mountain National <span class="hlt">Park</span> (n = 114 lakes), and Yellowstone National <span class="hlt">Park</span> (n = 294 lakes). Lakes that had a high probability of having an ANC concentration <100 microeq/L, and therefore sensitive to acidic deposition, are located in basins with elevations >3000 m, with <30% of the catchment having northeast aspect and with >80% of the catchment bedrock having low buffering capacity. The modeling results indicate that the most sensitive lakes are located in Rocky Mountain National <span class="hlt">Park</span> and Grand Teton National <span class="hlt">Park</span>. This technique for evaluating the lake sensitivity to acidic deposition is useful for designing long-term monitoring plans and is potentially transferable to other remote mountain areas of the United States and the world.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018E%26ES..107a2039O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018E%26ES..107a2039O"><span>Relationship between <span class="hlt">glacier</span> melting and atmospheric circulation in the southeast Siberia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Osipova, O. P.; Osipov, E. Y.</p> <p>2018-01-01</p> <p>The interaction between climate and cryosphere is a key issue in recent years. Changes in surface mass balance of mountain <span class="hlt">glaciers</span> closely correspond to differential changes in atmospheric circulation. Mountain <span class="hlt">glaciers</span> in southeast Siberia located on <span class="hlt">East</span> Sayan, Baikalsky and Kodar ridges have been continuously shrinking since the end of the Little Ice Age. In this study we used daily synoptic weather maps (Irkutsk Center of Hydrometeorology and Environmental Monitoring), 500 hPa, 700 hPa and 850 hPa geopotential height and air temperature data of NCEP/NCAR reanalysis to assess relationships between atmospheric circulation patterns and the sum of positive temperature (SPT), a predictor of summer ice/snow ablation. Results show that increased SPT (ablation) is generally associated with anticyclones and anticyclonic pressure fields (with cloudless weather conditions) and warm atmospheric fronts. Decreased SPT (ablation) is strongly correlated with cyclones and cyclonic type pressure fields, cold atmospheric fronts and air advections. Significant correlations have been found between ablation and cyclonic/anticyclonic activity. Revealed decreasing trends in the SPT in three glaciarized ridges at the beginning of the 21st century led to changes of air temperature and snow/ice melt climates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26632967','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26632967"><span>Polychlorinated Biphenyls in a Temperate Alpine <span class="hlt">Glacier</span>: 1. Effect of Percolating Meltwater on their Distribution in <span class="hlt">Glacier</span> Ice.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pavlova, Pavlina Aneva; Jenk, Theo Manuel; Schmid, Peter; Bogdal, Christian; Steinlin, Christine; Schwikowski, Margit</p> <p>2015-12-15</p> <p>In Alpine regions, <span class="hlt">glaciers</span> act as environmental archives and can accumulate significant amounts of atmospherically derived pollutants. Due to the current climate-warming-induced accelerated melting, these pollutants are being released at correspondingly higher rates. To examine the effect of melting on the redistribution of legacy pollutants in Alpine <span class="hlt">glaciers</span>, we analyzed polychlorinated biphenyls in an ice core from the temperate Silvretta <span class="hlt">glacier</span>, located in eastern Switzerland. This <span class="hlt">glacier</span> is affected by surface melting in summer. As a result, liquid water percolates down and particles are enriched in the current annual surface layer. Dating the ice core was a challenge because meltwater percolation also affects the traditionally used parameters. Instead, we counted annual layers of particulate black carbon in the ice core, adding the years with negative <span class="hlt">glacier</span> mass balance, that is, years with melting and subsequent loss of the entire annual snow accumulation. The analyzed samples cover the time period 1930-2011. The concentration of indicator PCBs (iPCBs) in the Silvretta ice core follows the emission history, peaking in the 1970s (2.5 ng/L). High PCB values in the 1990s and 1930s are attributed to meltwater-induced relocation within the <span class="hlt">glacier</span>. The total iPCB load at the Silvretta ice core site is 5 ng/cm(2). A significant amount of the total PCB burden in the Silvretta <span class="hlt">glacier</span> has been released to the environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27266318','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27266318"><span>Distribution and transportation of mercury from <span class="hlt">glacier</span> to lake in the Qiangyong <span class="hlt">Glacier</span> Basin, southern Tibetan Plateau, China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sun, Shiwei; Kang, Shichang; Huang, Jie; Li, Chengding; Guo, Junming; Zhang, Qianggong; Sun, Xuejun; Tripathee, Lekhendra</p> <p>2016-06-01</p> <p>The Tibetan Plateau is home to the largest aggregate of <span class="hlt">glaciers</span> outside the Polar Regions and is a source of fresh water to 1.4 billion people. Yet little is known about the transportation and cycling of Hg in high-elevation <span class="hlt">glacier</span> basins on Tibetan Plateau. In this study, surface snow, <span class="hlt">glacier</span> melting stream water and lake water samples were collected from the Qiangyong <span class="hlt">Glacier</span> Basin. The spatiotemporal distribution and transportation of Hg from <span class="hlt">glacier</span> to lake were investigated. Significant diurnal variations of dissolved Hg (DHg) concentrations were observed in the river water, with low concentrations in the morning (8:00am-14:00pm) and high concentrations in the afternoon (16:00pm-20:00pm). The DHg concentrations were exponentially correlated with runoff, which indicated that runoff was the dominant factor affecting DHg concentrations in the river water. Moreover, significant decreases of Hg were observed during transportation from <span class="hlt">glacier</span> to lake. DHg adsorption onto particulates followed by the sedimentation of particulate-bound Hg (PHg) could be possible as an important Hg removal mechanism during the transportation process. Significant decreases in Hg concentrations were observed downstream of Xiao Qiangyong Lake, which indicated that the high-elevation lake system could significantly affect the distribution and transportation of Hg in the Qiangyong <span class="hlt">Glacier</span> Basin. Copyright © 2016. Published by Elsevier B.V.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.2632A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.2632A"><span>Effect of fjord geometry on tidewater <span class="hlt">glacier</span> stability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Åkesson, Henning; Nisancioglu, Kerim H.; Nick, Faezeh M.</p> <p>2016-04-01</p> <p>Many marine-terminating <span class="hlt">glaciers</span> have thinned, accelerated and retreated during the last two decades, broadly consistent with warmer atmospheric and oceanic conditions. However, these patterns involve considerable spatial and temporal variability, with diverse <span class="hlt">glacier</span> behavior within the same regions. Similarly, reconstructions of marine-terminating <span class="hlt">glaciers</span> indicate highly asynchronous retreat histories. While it is well known that retrograde slopes can cause marine ice sheet instabilities, the effect of lateral drag and fjord width has received less attention. Here, we test the hypothesis that marine outlet <span class="hlt">glacier</span> stability is largely controlled by fjord width, and to a less extent by regional climate forcing. We employ a dynamic flowline model on idealized <span class="hlt">glacier</span> geometries (representative of different outlet <span class="hlt">glaciers</span>) to investigate geometric controls on decadal and longer times scales. The model accounts for driving and resistive stresses of <span class="hlt">glacier</span> flow as well as along-flow stress transfer. It has a physical treatment of iceberg calving and a time-adaptive grid allowing for continuous tracking of grounding-line migration. We apply changes in atmospheric and oceanic forcing and show how wide and narrow fjord sections foster <span class="hlt">glacier</span> (in)stabilities. We also evaluate the effect of including a surface mass balance - elevation feedback in such a setting. Finally, the relevance of these results to past and future marine-terminating <span class="hlt">glacier</span> stability is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005HyPr...19..231M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005HyPr...19..231M"><span>A revised Canadian perspective: progress in <span class="hlt">glacier</span> hydrology</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Munro, D. Scott</p> <p>2005-01-01</p> <p>Current research into <span class="hlt">glacier</span> hydrology is occurring at a time when <span class="hlt">glaciers</span> around the world, particularly those whose hydrological regimes affect populated areas, are shrinking as they go through a state of perpetual negative annual mass balance. Small <span class="hlt">glaciers</span> alone are likely to contribute 0·5 to 1 mm year-1 to global sea-level rise, with associated reductions in local freshwater resources, impacts upon freshwater ecosystems and increased risk of hazard due to outburst floods. Changes to the accumulation regimes of <span class="hlt">glaciers</span> and ice sheets may be partly responsible, so the measurement and distribution of snowfall in <span class="hlt">glacierized</span> basins, a topic long represented in non-<span class="hlt">glacierized</span> basin research, is now beginning to receive more attention than it did before, aided by the advent of reliable automatic weather stations that provide data throughout the year. Satellite data continue to be an important information source for summer meltwater estimation, as distributed models, and their need for albedo maps, continue to develop. This further entails the need for simplifications to energy balance components, sacrificing point detail so that spatial calculation may proceed more quickly. The understanding of surface meltwater routing through the <span class="hlt">glacier</span> to produce stream outflow continues to be a stimulating area of research, as demonstrated by activity at the Trapridge <span class="hlt">Glacier</span>, Canada, and Canadian involvement in the Haut <span class="hlt">Glacier</span> d'Arolla, Switzerland. As Canadian <span class="hlt">glacier</span> monitoring continues to evolve, effort must be directed toward developing situations where mass balance, meltwater generation and flow routing studies can be done together at selected sites. Copyright</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70029799','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70029799"><span>Development of a spatial analysis method using ground-based repeat photography to detect changes in the alpine treeline ecotone, <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, U.S.A.</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Roush, W.; Munroe, Jeffrey S.; Fagre, D.B.</p> <p>2007-01-01</p> <p>Repeat photography is a powerful tool for detection of landscape change over decadal timescales. Here a novel method is presented that applies spatial analysis software to digital photo-pairs, allowing vegetation change to be categorized and quantified. This method is applied to 12 sites within the alpine treeline ecotone of <span class="hlt">Glacier</span> National <span class="hlt">Park</span>, Montana, and is used to examine vegetation changes over timescales ranging from 71 to 93 years. Tree cover at the treeline ecotone increased in 10 out of the 12 photo-pairs (mean increase of 60%). Establishment occurred at all sites, infilling occurred at 11 sites. To demonstrate the utility of this method, patterns of tree establishment at treeline are described and the possible causes of changes within the treeline ecotone are discussed. Local factors undoubtedly affect the magnitude and type of the observed changes, however the ubiquity of the increase in tree cover implies a common forcing mechanism. Mean minimum summer temperatures have increased by 1.5??C over the past century and, coupled with variations in the amount of early spring snow water equivalent, likely account for much of the increase in tree cover at the treeline ecotone. Lastly, shortcomings of this method are presented along with possible solutions and areas for future research. ?? 2007 Regents of the University of Colorado.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/sim/3122/','USGSPUBS'); return false;" href="https://pubs.usgs.gov/sim/3122/"><span>Marine benthic habitat mapping of Muir Inlet, <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, Alaska, with an evaluation of the Coastal and Marine Ecological Classification Standard III</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Trusel, Luke D.; Cochrane, Guy R.; Etherington, Lisa L.; Powell, Ross D.; Mayer, Larry A.</p> <p>2010-01-01</p> <p>Seafloor geology and potential benthic habitats were mapped in Muir Inlet, <span class="hlt">Glacier</span> Bay National <span class="hlt">Park</span> and Preserve, Alaska, using multibeam sonar, ground-truth information, and geological interpretations. Muir Inlet is a recently deglaciated fjord that is under the influence of glacial and paraglacial marine processes. High glacially derived sediment and meltwater fluxes, slope instabilities, and variable bathymetry result in a highly dynamic estuarine environment and benthic ecosystem. We characterize the fjord seafloor and potential benthic habitats using the Coastal and Marine Ecological Classification Standard (CMECS) recently developed by the National Oceanic and Atmospheric Administration (NOAA) and NatureServe. Substrates within Muir Inlet are dominated by mud, derived from the high glacial debris flux. Water-column characteristics are derived from a combination of conductivity temperature depth (CTD) measurements and circulation-model results. We also present modern glaciomarine sediment accumulation data from quantitative differential bathymetry. These data show Muir Inlet is divided into two contrasting environments: a dynamic upper fjord and a relatively static lower fjord. The accompanying maps represent the first publicly available high-resolution bathymetric surveys of Muir Inlet. The results of these analyses serve as a test of the CMECS and as a baseline for continued mapping and correlations among seafloor substrate, benthic habitats, and glaciomarine processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.7162I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.7162I"><span>Geomorphic consequences of two large <span class="hlt">glacier</span> and rock <span class="hlt">glacier</span> destabilizations in the Central and northern Chilean Andes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Iribarren Anacona, Pablo; Bodin, Xavier</p> <p>2010-05-01</p> <p>Mountain areas are occasionaly affected by complex mass movements of high magnitude and large extent, which generally involve water, snow, rock and ice in variable proportions. Those events can take the form of rock avalanche, landslide, debris flow, <span class="hlt">glacier</span> collapse or a combination of these phenomenons. In the Central Andes of Chile, they affect hardly accessible regions with low population, explaining the scarcity of previous studies. Nevertheless, during the last 30 years, some documented examples of such events in this region have shown that the volume of material involved is in the order of several millions of m³, the areas affected can reach several tenth of km² and the velocity of the movement can exceed several tenths of m/s. In this context, this study intends i) to inventory and to describe the main characteristics of events previously documented in the Central Andes of Chile, and ii) analyse in detail two recent events (2005-2007) never described before which have affected in one case a <span class="hlt">glacier</span> and in another case a rock <span class="hlt">glacier</span>. With the objectives of determining the possible chain of triggering factors and interpreting the event's significance in terms of geomorphic, cryogenic and climatic dynamics, we used air photographs, satellite imagery (Landsat TM & ETM+; Quick Bird when available in Google Earth 5.0), data from the closest meteorological stations, <span class="hlt">glacier</span> mass balance data and seismic records to investigate the collapse of a rock <span class="hlt">glacier</span> occurred in 2006 on the west-facing flank of the Cerro Las Tórtolas (6160 m asl; 29°58' S. - 69°55' W.), in the arid North of Chile, and the collapse of a <span class="hlt">glacier</span> that occurred during austral summer 2006-2007 on the South side of the Tinguiririca Volcano (4075 m asl; 34°48' S. - 70°21' W.). The rock <span class="hlt">glacier</span> collapse of the Cerro Las Tórtolas West flank occurred during the spring of 2006, but signs of destabilization were already observable since the end of 2005. The deposit of the collapsed mass of the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/nv0242.photos.380512p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/nv0242.photos.380512p/"><span>1. View of Building 802 from the <span class="hlt">parking</span> lot, 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>1. View of Building 802 from the <span class="hlt">parking</span> lot, Building 800 in the background, facing <span class="hlt">east</span>. - Naval Air Station Fallon, 100-man Fallout Shelter, 800 Complex, off Carson Road near intersection of Pasture & Berney Roads, Fallon, Churchill County, NV</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..14..939L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..14..939L"><span>Uncovering <span class="hlt">glacier</span> dynamics beneath a debris mantle</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lefeuvre, P.-M.; Ng, F. S. L.</p> <p>2012-04-01</p> <p>Debris-covered <span class="hlt">glaciers</span> (DCGs) have an extensive sediment mantle whose low albedo influences their surface energy balance to cause a buffering effect that could enhance or reduce ablation rates depending on the sediment thickness. The last effect suggests that some DCGs may be less sensitive to climate change and survive for longer than debris-free (or 'clean') <span class="hlt">glaciers</span> under sustained climatic warming. However, the origin of DCGs is debated and the precise impact of the debris mantle on their flow dynamics and surface geometry has not been quantified. Here we investigate these issues with a numerical model that encapsulates ice-flow physics and surface debris evolution and transport along a <span class="hlt">glacier</span> flow-line, as well as couples these with <span class="hlt">glacier</span> mass balance. We model the impact of surface debris on ablation rates by a mathematical function based on published empirical data (including Ostrem's curve). A key interest is potential positive feedback of ablation on debris thickening and lowering of surface albedo. Model simulations show that when DCGs evolve to attain steady-state profiles, they reach lower elevations than clean <span class="hlt">glaciers</span> do for the same initial and climatic conditions. Their mass-balance profile at steady state displays an inversion near the snout (where the debris cover is thickest) that is not observed in the clean-<span class="hlt">glacier</span> simulations. In these cases, where the mantle causes complete buffering to inhibit ablation, the DCG does not reach a steady-state profile, and the sediment thickness evolves to a steady value that depends sensitively on the <span class="hlt">glacier</span> surface velocities. Variation in the assumed englacial debris concentration in our simulations also determines <span class="hlt">glacier</span> behaviour. With low englacial debris concentration, the DCG retreats initially while its mass-balance gradient steepens, but the <span class="hlt">glacier</span> re-advances if it subsequently builds up a thick enough debris cover to cause complete buffering. We identify possible ways and challenges of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016TCry...10.2129T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016TCry...10.2129T"><span>ICESat laser altimetry over small mountain <span class="hlt">glaciers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Treichler, Désirée; Kääb, Andreas</p> <p>2016-09-01</p> <p>Using sparsely glaciated southern Norway as a case study, we assess the potential and limitations of ICESat laser altimetry for analysing regional <span class="hlt">glacier</span> elevation change in rough mountain terrain. Differences between ICESat GLAS elevations and reference elevation data are plotted over time to derive a <span class="hlt">glacier</span> surface elevation trend for the ICESat acquisition period 2003-2008. We find spatially varying biases between ICESat and three tested digital elevation models (DEMs): the Norwegian national DEM, SRTM DEM, and a high-resolution lidar DEM. For regional <span class="hlt">glacier</span> elevation change, the spatial inconsistency of reference DEMs - a result of spatio-temporal merging - has the potential to significantly affect or dilute trends. Elevation uncertainties of all three tested DEMs exceed ICESat elevation uncertainty by an order of magnitude, and are thus limiting the accuracy of the method, rather than ICESat uncertainty. ICESat matches <span class="hlt">glacier</span> size distribution of the study area well and measures small ice patches not commonly monitored in situ. The sample is large enough for spatial and thematic subsetting. Vertical offsets to ICESat elevations vary for different <span class="hlt">glaciers</span> in southern Norway due to spatially inconsistent reference DEM age. We introduce a per-<span class="hlt">glacier</span> correction that removes these spatially varying offsets, and considerably increases trend significance. Only after application of this correction do individual campaigns fit observed in situ <span class="hlt">glacier</span> mass balance. Our correction also has the potential to improve <span class="hlt">glacier</span> trend significance for other causes of spatially varying vertical offsets, for instance due to radar penetration into ice and snow for the SRTM DEM or as a consequence of mosaicking and merging that is common for national or global DEMs. After correction of reference elevation bias, we find that ICESat provides a robust and realistic estimate of a moderately negative <span class="hlt">glacier</span> mass balance of around -0.36 ± 0.07 m ice per year. This regional</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.C41E0457P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.C41E0457P"><span>Airborne geophysical investigations of basal conditions at flow transitions of outlet <span class="hlt">glaciers</span> on the Greenland Ice Sheet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Palmer, S. J.; Dowdeswell, J. A.; Christoffersen, P.; Siegert, M. J.; Blankenship, D. D.; Young, D. A.; Greenbaum, J.</p> <p>2011-12-01</p> <p>Recent observations have shown that the fast flowing marine-terminating outlet <span class="hlt">glaciers</span> which drain the Greenland Ice Sheet (GrIS) have thinned in places at rates in excess of 10 m yr-1. The 21 largest outlet <span class="hlt">glaciers</span> in Greenland accelerated by 57 % between 1996 and 2005, leading to a 100 Gt yr-1 increase in mass loss due to ice discharge over the same period and a 150 % increase of the GrIS's contribution to sea level. Observations that thinning rates are greater than those expected from changes in surface mass balance alone suggest thinning of some GrIS marine-terminating outlet <span class="hlt">glaciers</span> can be attributed to changes in ice dynamics. An important question for both scientists and policy makers is how the GrIS will react to projected temperature increases, particularly in the context that the Arctic is likely to warm at a greater rate than the global average due to the ice-albedo feedback. As the combined width of all major marine-terminating <span class="hlt">glaciers</span> draining the GrIS (as measured at the narrowest point in each case) is less 200 km, an understanding of their dynamics is crucial in predicting the effect of future warming on the ice sheet as a whole. During April 2011, we used a Basler BT-67 aircraft equipped with a suite of geophysical instruments to investigate three major <span class="hlt">glacier</span> systems in Greenland. Data were acquired at the Sermeq Kujatdl and Rink <span class="hlt">Glacier</span> systems in West Greenland; and Daugaard Jensen <span class="hlt">Glacier</span> in <span class="hlt">East</span> Greenland. The study areas were selected because they are major drainage basins (c. 103-105 km2) which provide a high ice flux to the sea (c. 10-20 km3 yr-1); and are located in different regions of the GrIS with correspondingly different atmospheric and oceanic settings. Here we present results from the High Capability Radar Sounder instrument, a phase coherent VHF ice-penetrating radar which operates in frequency-chirped mode from 52.5 to 67.5 MHz. We use these data to determine ice thickness along flightlines both parallel and perpendicular to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.7677A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.7677A"><span>Melting beneath Greenland outlet <span class="hlt">glaciers</span> and ice streams</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alexander, David; Perrette, Mahé; Beckmann, Johanna</p> <p>2015-04-01</p> <p>Basal melting of fast-flowing Greenland outlet <span class="hlt">glaciers</span> and ice streams due to frictional heating at the ice-bed interface contributes significantly to total <span class="hlt">glacier</span> mass balance and subglacial meltwater flux, yet modelling this basal melt process in Greenland has received minimal research attention. A one-dimensional dynamic ice-flow model is calibrated to the present day longitudinal profiles of 10 major Greenland outlet <span class="hlt">glaciers</span> and ice streams (including the Jakobshavn Isbrae, Petermann <span class="hlt">Glacier</span> and Helheim <span class="hlt">Glacier</span>) and is validated against published ice flow and surface elevation measurements. Along each longitudinal profile, basal melt is calculated as a function of ice flow velocity and basal shear stress. The basal shear stress is dependent on the effective pressure (difference between ice overburden pressure and water pressure), basal roughness and a sliding parametrization. Model output indicates that where outlet <span class="hlt">glaciers</span> and ice streams terminate into the ocean with either a small floating ice tongue or no floating tongue whatsoever, the proportion of basal melt to total melt (surface, basal and submarine melt) is 5-10% (e.g. Jakobshavn Isbrae; Daugaard-Jensen <span class="hlt">Glacier</span>). This proportion is, however, negligible where larger ice tongues lose mass mostly by submarine melt (~1%; e.g. Nioghalvfjerdsfjorden <span class="hlt">Glacier</span>). Modelled basal melt is highest immediately upvalley of the grounding line, with contributions typically up to 20-40% of the total melt for slippery beds and up to 30-70% for resistant beds. Additionally, modelled grounding line and calving front migration inland for all outlet <span class="hlt">glaciers</span> and ice streams of hundreds of metres to several kilometres occurs. Including basal melt due to frictional heating in outlet <span class="hlt">glacier</span> and ice stream models is important for more accurately modelling mass balance and subglacial meltwater flux, and therefore, more accurately modelling outlet <span class="hlt">glacier</span> and ice stream dynamics and responses to future climate change.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.6261D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.6261D"><span><span class="hlt">Glacier</span> discharge and climate variations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dominguez, M. Carmen; Rodriguez-Puebla, Concepcion; Encinas, Ascension H.; Visus, Isabel; Eraso, Adolfo</p> <p>2010-05-01</p> <p>Different studies account for the warming in the polar regions that consequently would affect <span class="hlt">Glacier</span> Discharge (GD). Since changes in GD may cause large changes in sensible and latent heat fluxes, we ask about the relationships between GD and climate anomalies, which have not been quantified yet. In this study we apply different statistical methods such as correlation, Singular Spectral Analysis and Wavelet to compare the behaviour of GD data in two Experimental Pilot Catchments (CPE), one (CPE-KG-62°S) in the Antarctica and the other (CPE-KVIA-64°N) in the Arctic regions. Both CPE's are measuring sub- and endo-<span class="hlt">glacier</span> drainage for recording of <span class="hlt">glacier</span> melt water run-off. The CPE-KG-62°S is providing hourly GD time series since January 2002 in Collins <span class="hlt">glacier</span> of the Maxwell Bay in King George Island (62°S, 58°W). The second one, CPE-KVIA-64°N, is providing hourly GD time series since September 2003 in the Kviarjökull <span class="hlt">glacier</span> of the Vatnajökull ice cap in Iceland (64°N, 16°W). The soundings for these measurements are pressure sensors installed in the river of the selected catchments for the ice cap (CPE-KG-62°S) and in the river of the <span class="hlt">glacier</span> for (CPE-KVIA-64°N). In each CPE, the calibration function between level and discharge has been adjusted, getting a very high correlation coefficient (0.99 for the first one and 0.95 for the second one), which let us devise a precise discharge law for the <span class="hlt">glacier</span>. We obtained relationships between GD with atmospheric variables such as radiation, temperature, relative humidity, atmospheric pressure and precipitation. We also found a negative response of GD to El Niño teleconnection index. The results are of great interest due to the GD impact on the climate system and in particular for sea level rise.</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>