Journal articles on the topic 'Volcanism Nevada'

To see the other types of publications on this topic, follow the link: Volcanism Nevada.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Volcanism Nevada.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Bradshaw, T. K., and Eugene I. Smith. "Polygenetic Quaternary volcanism at Crater Flat, Nevada." Journal of Volcanology and Geothermal Research 63, no. 3-4 (November 1994): 165–82. http://dx.doi.org/10.1016/0377-0273(94)90072-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

CADOUX, ANITA, YVES MISSENARD, RAYMUNDO G. MARTINEZ-SERRANO, and HERVÉ GUILLOU. "Trenchward Plio-Quaternary volcanism migration in the Trans-Mexican Volcanic Belt: the case of the Sierra Nevada range." Geological Magazine 148, no. 3 (January 28, 2011): 492–506. http://dx.doi.org/10.1017/s0016756810000993.

Full text
Abstract:
AbstractThe Miocene–Quaternary Trans-Mexican Volcanic arc is thought to have grown southwards (i.e. trenchward) since the Pliocene. This theory is mainly supported by roughly N–S-directed polygenetic volcanic ranges along which volcanic activity migrates southwards with time. We investigated the eruptive history of one of these ranges, the Sierra Nevada (east boundary of Mexico City basin), by compiling literature ages and providing new K–Ar dates. Our K–Ar ages are the first ones for the northernmost Tláloc and Telapón volcanoes and for the ancestral Popocatépetl (Nexpayantla). The obtained ages reveal that the four stratovolcanoes forming the range worked contemporaneously during most of the Middle to Late Pleistocene. However, taking into account the onset of the volcanic activity, a southward migration is evidenced along the Sierra Nevada: volcanism initiated at its northern tip at least 1.8 Ma ago at Tláloc volcano, extended southwards 1 Ma ago with Iztaccíhuatl and appeared at its southern end 329 ka ago with the Nexpayantla cone. Such a migration would be most probably primarily driven by Cocos slab roll-back and steepening rather than by regional crustal tectonics, which played a secondary role by controlling the apparent alignment of the volcanoes.
APA, Harvard, Vancouver, ISO, and other styles
3

Browne, Brandon L., Raul Becerra, Colin Campbell, Phillip Saleen, and Frank R. Wille. "Quaternary basaltic volcanism in the Golden Trout Volcanic Field, southern Sierra Nevada, California." Journal of Volcanology and Geothermal Research 343 (September 2017): 25–44. http://dx.doi.org/10.1016/j.jvolgeores.2017.05.028.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Rodríguez García, Gabriel, and Gloria Obando. "Volcanism of the La Quinta Formation in the Perijá mountain range." Boletín Geológico, no. 46 (June 30, 2020): 51–94. http://dx.doi.org/10.32685/0120-1425/boletingeo.46.2020.535.

Full text
Abstract:
This study reports new data on the petrography, total rock chemistry and U-Pb zircon geochronology of volcanic rocks of the La Quinta Formation that outcrop the western flank of the Perijá mountain range and the Cesar and La Guajira departments. The volcanic rocks consist of basaltic, andesitic, dacitic and rhyolitic lavas, and the volcaniclastic rocks consist of crystal-vitric and crystal-lithic tuffs and agglomerates of calc-alkaline affinity, formed in a continental margin arc setting. Geochronological data suggest that the La Quinta Formation was volcanically active for approximately 25 Ma, during which its composition varied from basaltic trachyandesites to rhyolites. U-Pb dating suggests that the volcanism began in approximately 191 Ma (Sinemurian age) and continued until approximately 164 Ma, with at least three periods of increased volcanic activity. The inherited zircons contain Triassic, Permian, Neoproterozoic and Mesoproterozoic populations, indicating that this arc was emplaced on rocks of the Chibcha Terrane along the South American paleomargin and that it is part of the same arc that formed the Jurassic volcanic rocks of the Sierra Nevada de Santa Marta, Cocinas and San Lucas mountain ranges and the Upper Magdalena Valley.
APA, Harvard, Vancouver, ISO, and other styles
5

Aalto, K. R. "Philip Tyson's 1849 study of California gold prospects." Earth Sciences History 36, no. 1 (January 1, 2017): 30–40. http://dx.doi.org/10.17704/1944-6178-36.1.30.

Full text
Abstract:
Philip Thomas Tyson (1799–1877) toured privately through central California, from San Francisco through the Sierra Nevada foothills gold prospects in 1849, to assess their potential and the general geology of the region. He produced the first regional map with geologic notations and several rough topographic/geologic cross-sections. He described Coast Range basement rocks, now described as Franciscan Complex mélange and broken formation, the stratigraphic configuration of the Great Valley, and general geology of the Sierra Nevada foothills. He recognized that the older Coast Range and Sierran basement were deformed prior to recent volcanism and extensive terrestrial fluvial sedimentation, likely Neogene in age.
APA, Harvard, Vancouver, ISO, and other styles
6

Saleeby, Jason, and Zorka Saleeby. "Late Cenozoic structure and tectonics of the southern Sierra Nevada–San Joaquin Basin transition, California." Geosphere 15, no. 4 (June 13, 2019): 1164–205. http://dx.doi.org/10.1130/ges02052.1.

Full text
Abstract:
AbstractThis paper presents a new synthesis for the late Cenozoic tectonic, paleogeographic, and geomorphologic evolution of the southern Sierra Nevada and adjacent eastern San Joaquin Basin. The southern Sierra Nevada and San Joaquin Basin contrast sharply, with the former constituting high-relief basement exposures and the latter constituting a Neogene marine basin with superposed low-relief uplifts actively forming along its margins. Nevertheless, we show that Neogene basinal conditions extended continuously eastward across much of the southern Sierra Nevada, and that during late Neogene–Quaternary time, the intra-Sierran basinal deposits were uplifted and fluvially reworked into the San Joaquin Basin. Early Neogene normal-sense growth faulting was widespread and instrumental in forming sediment accommodation spaces across the entire basinal system. Upon erosion of the intra-Sierran basinal deposits, structural relief that formed on the basement surface by the growth faults emerged as topographic relief. Such “weathered out” fossil fault scarps control much of the modern southern Sierra landscape. This Neogene high-angle fault system followed major Late Cretaceous basement structures that penetrated the crust and that formed in conjunction with partial loss of the region’s underlying mantle lithosphere. This left the region highly prone to surface faulting, volcanism, and surface uplift and/or subsidence transients during subsequent tectonic regimes. The effects of the early Neogene passage of the Mendocino Triple Junction were amplified as a result of the disrupted state of the region’s basement. This entailed widespread high-angle normal faulting, convecting mantle-sourced volcanism, and epeirogenic transients that were instrumental in sediment dispersal, deposition, and reworking patterns. Subsequent phases of epeirogenic deformation forced additional sediment reworking episodes across the southern Sierra Nevada–eastern San Joaquin Basin region during the late Miocene break-off and west tilt of the Sierra Nevada microplate and the Pliocene–Quaternary loss of the region’s residual mantle lithosphere that was left intact from the Late Cretaceous tectonic regime. These late Cenozoic events have left the high local-relief southern Sierra basement denuded of its Neogene basinal cover and emergent immediately adjacent to the eastern San Joaquin Basin and its eastern marginal uplift zone.
APA, Harvard, Vancouver, ISO, and other styles
7

Best, Myron G., and Eric H. Christiansen. "Limited extension during peak Tertiary volcanism, Great Basin of Nevada and Utah." Journal of Geophysical Research: Solid Earth 96, B8 (July 30, 1991): 13509–28. http://dx.doi.org/10.1029/91jb00244.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Putirka, Keith, and Cathy J. Busby. "The tectonic significance of high-K2O volcanism in the Sierra Nevada, California." Geology 35, no. 10 (2007): 923. http://dx.doi.org/10.1130/g23914a.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Colgan, Joseph P., Trevor A. Dumitru, and Elizabeth L. Miller. "Diachroneity of Basin and Range extension and Yellowstone hotspot volcanism in northwestern Nevada." Geology 32, no. 2 (2004): 121. http://dx.doi.org/10.1130/g20037.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Grunder, Anita L., Erik W. Klemetti, Todd C. Feeley, and Claire M. McKee. "Eleven million years of arc volcanism at the Aucanquilcha Volcanic Cluster, northern Chilean Andes: implications for the life span and emplacement of plutons." Transactions of the Royal Society of Edinburgh: Earth Sciences 97, no. 4 (December 2006): 415–36. http://dx.doi.org/10.1017/s0263593300001541.

Full text
Abstract:
ABSTRACTThe arid climate of the Altiplano has preserved a volcanic history of ∼11 million years at the Aucanquilcha Volcanic Cluster (AVC), northern Chile, which is built on thick continental crust. The AVC has a systematic temporal, spatial, compositional and mineralogical development shared by other long-lived volcanic complexes, indicating a common pattern in continental magmatism with implications for the development of underlying plutonic complexes, that in turn create batholiths.The AVC is a ∼700-km2, Tertiary to Recent cluster of at least 19 volcanoes that have erupted andesite and dacite lavas (∼55 to 68 wt.% SiO2) and a small ash-flow tuff, totalling 327 ± 20 km3. Forty 40Ar/39Ar ages for the AVC range from 10·97 ± 0·35 to 0·24 ± 0·05 Ma and define three major 1·5 to 3 million-year pulses of volcanism followed by the present pulse expressed as Volcán Aucanquilcha. The first stage of activity (∼11–8 Ma, Alconcha Group) produced seven volcanoes and the 2-km3 Ujina ignimbrite and is a crudely bimodal suite of pyroxene andesite and dacite. After a possible two million year hiatus, the second stage of volcanism (∼6–4 Ma, Gordo Group) produced at least five volcanoes ranging from pyroxene andesite to dacite. The third stage (∼4–2 Ma, Polan Group) represents a voluminous pulse of activity, with eruption of at least another five volcanoes, broadly distributed in the centre of the AVC, and composed dominantly of biotite amphibole dacite; andesites at this stage occur as magmatic inclusions. The most recent activity (1 Ma to recent) is in the centre of the AVC at Volcán Aucanquilcha, a potentially active composite volcano made of biotite-amphibole dacite with andesite and dacite magmatic inclusions.These successive eruptive groups describe (1) a spatial pattern of volcanism from peripheral to central, (2) a corresponding change from compositionally diverse andesite-dacite volcanism to compositionally increasingly restricted and increasingly silicic dacite, (3) a change from early anhydrous mafic silicate assemblages (pyroxene dominant) to later biotite amphibole dacite, (4) an abrupt increase in eruption rate; and (5) the onset of pervasive hydrothermal alteration.The evolutionary succession of the 327-km3 AVC is similar to other long-lived intermediate volcanic complexes of very different volumes, e.g., eastern Nevada (thousands of km3, Gans et al. 1989; Grunder 1995), Yanacocha, Perú (tens of km3, Longo 2005), and the San Juan Volcanic System (tens of thousands of km3, Lipman 2007) and finds an analogue in the 10-m. y. history and incremental growth of the Cretaceous Tuolumne Intrusive Suite (Coleman et al. 2004; Glazner et al. 2004). The present authors interpret the AVC to reflect episodic sampling of the protracted and fitful development of an integrated and silicic middle to upper crustal magma reservoir over a period of at least 11 million years.
APA, Harvard, Vancouver, ISO, and other styles
11

Connor, Charles B., John A. Stamatakos, David A. Ferrill, Brittain E. Hill, Goodluck I. Ofoegbu, F. Michael Conway, Budhi Sagar, and John Trapp. "Geologic factors controlling patterns of small-volume basaltic volcanism: Application to a volcanic hazards assessment at Yucca Mountain, Nevada." Journal of Geophysical Research: Solid Earth 105, B1 (January 10, 2000): 417–32. http://dx.doi.org/10.1029/1999jb900353.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Lang Farmer, G., Allen F. Glazner, and Curtis R. Manley. "Did lithospheric delamination trigger late Cenozoic potassic volcanism in the southern Sierra Nevada, California?" Geological Society of America Bulletin 114, no. 6 (June 2002): 754–68. http://dx.doi.org/10.1130/0016-7606(2002)114<0754:dldtlc>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Cousens, Brian, Christopher D. Henry, and Vishal Gupta. "Distinct mantle sources for Pliocene–Quaternary volcanism beneath the modern Sierra Nevada and adjacent Great Basin, northern California and western Nevada, USA." Geosphere 8, no. 3 (June 2012): 562–80. http://dx.doi.org/10.1130/ges00741.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Canada, Andrew S., Elizabeth J. Cassel, Allen J. McGrew, M. Elliot Smith, Daniel F. Stockli, Kenneth A. Foland, Brian R. Jicha, and Brad S. Singer. "Eocene exhumation and extensional basin formation in the Copper Mountains, Nevada, USA." Geosphere 15, no. 5 (July 16, 2019): 1577–97. http://dx.doi.org/10.1130/ges02101.1.

Full text
Abstract:
Abstract Within extended orogens, records that reflect the driving processes and dynamics of early extension are often overprinted by subsequent orogenic collapse. The Copper Mountains of northeastern Nevada preserve an exceptional record of hinterland extensional deformation and high-elevation basin formation, but current geochronology and thermochronology are insufficient to relate this to broader structural trends in the region. This extension occurred concurrent with volcanism commonly attributed to Farallon slab removal. We combine thermochronology of both synextensional hanging-wall strata and footwall rocks to comprehensively evaluate the precise timing and style of this deformation. Specifically, we apply (U-Th)/(He-Pb) double dating of minerals extracted from Eocene–Oligocene Copper Basin strata with multi-mineral (U-Th)/He and 40Ar/39Ar thermochronology of rocks sampled across an ∼20 km transect of the Copper Mountains. We integrate basement and detrital thermochronology records to comprehensively evaluate the timing and rates of hinterland extension and basin sedimentation. Cooling and U-Pb crystallization ages show the Coffeepot Stock, which spans the width of the Copper Mountains, was emplaced at ca. 109–108 Ma, and then cooled through the 40Ar/39Ar muscovite and biotite closure temperatures by ca. 90 Ma, the zircon (U-Th)/He closure temperature between ca. 90 and 70 Ma, and the apatite (U-Th)/He closure temperature between 43 and 40 Ma. Detrital apatite and zircon (U-Th)/(He-Pb) double dating of late Eocene fluvial and lacustrine strata of the Dead Horse Formation and early Oligocene fluvial strata of the Meadow Fork Formation, both deposited in Copper Basin, shows that Early Cretaceous age detrital grains have a cooling history that is analogous to proximal intrusive rocks of the Coffeepot Stock. At ca. 38 Ma, cooling and depositional ages for Copper Basin strata reveal rapid exhumation of proximal source terranes (cooling rate of ∼37 °C/m.y.); in these terranes, 8–12 km of slip along the low-angle Copper Creek normal fault exhumed the Coffeepot Stock in the footwall. Late Eocene–early Oligocene slip along this fault and an upper fault splay, the Meadow Fork fault, created a half graben that accommodated ∼1.4 km of volcaniclastic strata, including ∼20 m of lacustrine strata that preserve the renowned Copper Basin flora. Single-crystal sanidine 40Ar/39Ar geochronology of interbedded tuffs in Copper Basin constrains the onset of rapid exhumation to 38.0 ± 0.9 Ma, indicating that surface-breaching extensional deformation was coincident with intense proximal volcanism. Coarse-grained syndeformational sediments of the Oligocene Meadow Fork Formation were deposited just prior to formation of an extensive regional Oligocene–Miocene unconformity and represent one of the most complete hinterland stratigraphic records of this time. We interpret this history of rapid late Eocene exhumation across the Copper Mountains, coeval volcanism, and subsequent unconformity formation to reflect dynamic and thermal effects associated with Farallon slab removal. The final phase of extension is recorded by late, high-angle normal faults that cut and rotate the early middle Miocene Jarbidge Rhyolite sequence, deposited unconformably in the hanging wall. These results provide an independent record of episodic Paleogene to Miocene exhumation documented in Cordilleran metamorphic core complexes and establish that substantial extension occurred locally in the hinterland prior to province-wide Miocene extensional break-up.
APA, Harvard, Vancouver, ISO, and other styles
15

Bond, David P. G., and Stephen E. Grasby. "Late Ordovician mass extinction caused by volcanism, warming, and anoxia, not cooling and glaciation." Geology 48, no. 8 (May 18, 2020): 777–81. http://dx.doi.org/10.1130/g47377.1.

Full text
Abstract:
Abstract The Ordovician saw major diversification in marine life abruptly terminated by the Late Ordovician mass extinction (LOME). Around 85% of species were eliminated in two pulses 1 m.y. apart. The first pulse, in the basal Hirnantian, has been linked to cooling and Gondwanan glaciation. The second pulse, later in the Hirnantian, is attributed to warming and anoxia. Previously reported mercury (Hg) spikes in Nevada (USA), South China, and Poland implicate an unknown large igneous province (LIP) in the crisis, but the timing of Hg loading has led to different interpretations of the LIP-extinction scenario in which volcanism causes cooling, warming, or both. We report close correspondence between Hg, Mo, and U anomalies, declines in enrichment factors of productivity proxies, and the two LOME pulses at the Ordovician-Silurian boundary stratotype (Dob’s Linn, Scotland). These support an extinction scenario in which volcanogenic greenhouse gases caused warming around the Katian-Hirnantian boundary that led to expansion of a preexisting deepwater oxygen minimum zone, productivity collapse, and the first LOME pulse. Renewed volcanism in the Hirnantian stimulated further warming and anoxia and the second LOME pulse. Rather than being the odd-one-out of the “Big Five” extinctions with origins in cooling, the LOME is similar to the others in being caused by volcanism, warming, and anoxia.
APA, Harvard, Vancouver, ISO, and other styles
16

Smith, Eugene I., Deborah L. Keenan, and Terry Plank. "Episodic Volcanism and Hot Mantle: Implications for Volcanic Hazard Studies at the Proposed Nuclear Waste Repository at Yucca Mountain, Nevada." GSA Today 12, no. 4 (2002): 4. http://dx.doi.org/10.1130/1052-5173(2002)012<0004:evahmi>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

DeGraaff Surpless, Kathleen, Diane Clemens-Knott, Andrew P. Barth, and Michelle Gevedon. "A survey of Sierra Nevada magmatism using Great Valley detrital zircon trace-element geochemistry: View from the forearc." Lithosphere 11, no. 5 (June 27, 2019): 603–19. http://dx.doi.org/10.1130/l1059.1.

Full text
Abstract:
AbstractThe well-characterized Sierra Nevada magmatic arc offers an unparalleled opportunity to improve our understanding of continental arc magmatism, but present bedrock exposure provides an incomplete record that is dominated by Cretaceous plutons, making it challenging to decipher details of older magmatism and the dynamic interplay between plutonism and volcanism. Moreover, the forearc detrital record includes abundant zircon formed during apparent magmatic lulls, suggesting that understanding the long-term history of arc magmatism requires integrating plutonic, volcanic, and detrital records. We present trace-element geochemistry of detrital zircon grains from the Great Valley forearc basin to survey Sierra Nevadan arc magmatism through Mesozoic time. We analyzed 257 previously dated detrital zircon grains from seven sandstone samples of volcanogenic, arkosic, and mixed compositions deposited ca. 145–80 Ma along the length of the forearc basin. Detrital zircon trace-element geochemistry is largely consistent with continental arc derivation and shows similar geochemical ranges between samples, regardless of location along strike of the forearc basin, depositional age, or sandstone composition. Comparison of zircon trace-element data from the forearc, arc, and retroarc regions revealed geochemical asymmetry across the arc that was persistent through time and demonstrated that forearc and retroarc basins sampled different parts of the arc and therefore recorded different magmatic histories. In addition, we identified a minor group of Jurassic detrital zircon grains with oceanic geochemical signatures that may have provenance in the Coast Range ophiolite. Taken together, these results suggest that the forearc detrital zircon data set reveals information different from that gleaned from the arc itself and that zircon compositions can help to identify and differentiate geochemically distinct parts of continental arc systems. Our results highlight the importance of integrating multiple proxies to fully document arc magmatism, demonstrating that detrital zircon geochemical data can enhance understanding of a well-characterized arc, and these data may prove an effective means by which to survey an arc that is inaccessible and therefore poorly characterized.
APA, Harvard, Vancouver, ISO, and other styles
18

Aalto, K. "Clarence King's Geology." Earth Sciences History 23, no. 1 (January 1, 2004): 9–31. http://dx.doi.org/10.17704/eshi.23.1.rx018782662jv071.

Full text
Abstract:
Clarence King (1842-1901) studied geology at Yale, served as a volunteer on Josiah Dwight Whitney's (1819-1896) Geological Survey of California, and directed the Fortieth Parallel Survey (1867-1872) from the Sierra Nevada across the Rocky Mountains, topo-graphically and geologically mapping some 100,000 square miles. He established a framework for orogenic history of the American Cordillera that has remained unchanged. Within this framework he recognized what we know today as the Sonoma, Sevier, and Laramide orogenies. He noted that folding of Paleozoic strata in the Great Basin recorded east-west crustal shortening, he delineated trends of Laramide folds, he determined that extensional Tertiary faulting that accompanied rhyolitic volcanism resulted in dislocation of old folds, and that ranges were broken into irregular blocks with considerable vertical displacement. King rejected strict Lyellian uniformitarianism and related Darwinian evolution to episodes of enhanced selection pressure engendered by natural catastrophes. His refinement to 24 Ma (million years) of Kelvin's earth age estimate from terrestrial refrigeration reinforced his conception that inadequate time existed to explain the Fortieth-Parallel geologic record by uniformitarianism, and that accelerated geologic processes best accounted for episodes of uplift/subsidence, faulting, volcanism, and landscape degradation. King thus stands out as an early actualist, quite modern in his approach to event stratigraphy.
APA, Harvard, Vancouver, ISO, and other styles
19

SAWYER, DAVID A., R. J. FLECK, M. A. LANPHERE, R. G. WARREN, D. E. BROXTON, and MARK R. HUDSON. "Episodic caldera volcanism in the Miocene southwestern Nevada volcanic field: Revised stratigraphic framework, 40Ar/39Ar geochronology, and implications for magmatism and extension." Geological Society of America Bulletin 106, no. 10 (October 1994): 1304–18. http://dx.doi.org/10.1130/0016-7606(1994)106<1304:ecvitm>2.3.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Manley, Curtis R., Allen F. Glazner, and G. Lang Farmer. "Timing of volcanism in the Sierra Nevada of California: Evidence for Pliocene delamination of the batholithic root?" Geology 28, no. 9 (September 2000): 811–14. http://dx.doi.org/10.1130/0091-7613(2000)028<0811:tovits>2.3.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Manley, Curtis R., Allen F. Glazner, and G. Lang Farmer. "Timing of volcanism in the Sierra Nevada of California: Evidence for Pliocene delamination of the batholithic root?" Geology 28, no. 9 (2000): 811. http://dx.doi.org/10.1130/0091-7613(2000)28<811:tovits>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Connor, Charles B., and Brittain E. Hill. "Three nonhomogeneous Poisson models for the probability of basaltic volcanism: Application to the Yucca Mountain region, Nevada." Journal of Geophysical Research: Solid Earth 100, B6 (June 10, 1995): 10107–25. http://dx.doi.org/10.1029/95jb01055.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Druschke, P., A. D. Hanson, and M. L. Wells. "Structural, stratigraphic, and geochronologic evidence for extension predating Palaeogene volcanism in the Sevier hinterland, east-central Nevada." International Geology Review 51, no. 7-8 (July 2009): 743–75. http://dx.doi.org/10.1080/00206810902917941.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Stahl, T. A., N. A. Niemi, M. P. Bunds, J. Andreini, and J. D. Wells. "Paleoseismic patterns of Quaternary tectonic and magmatic surface deformation in the eastern Basin and Range, USA." Geosphere 16, no. 1 (December 19, 2019): 435–55. http://dx.doi.org/10.1130/ges02156.1.

Full text
Abstract:
Abstract The competing contributions of tectonic and magmatic processes in accommodating continental extension are commonly obscured by a lack of on-fault paleoseismic information. This is especially true of the Sevier Desert, located at the eastern margin of the Basin and Range in central Utah (USA), where surface-rupturing faults are spatially associated with both regional detachment faults and Quaternary volcanism. Here, we use high-resolution topographic surveys (terrestrial lidar scans and real-time kinematic GPS), terrestrial cosmogenic nuclide (10Be and 3He) exposure dating, 40Ar/39Ar geochronology, and new neotectonic mapping to distinguish between modes of faulting and extension in a transect across the Sevier Desert. In the western Sevier Desert, the House Range and Cricket Mountains faults each have evidence of a single surface-rupturing earthquake in the last 20–30 k.y. and have time-integrated slip and extension rates of &lt;0.1 and ∼0.05 mm yr−1, respectively, since ca. 15–30 ka. These rates are similar to near-negligible modern geodetic extension estimates. Despite relatively low geologic, paleoseismic, and modern extension rates, both faults show evidence of contributing to the long-term growth of topographic relief and the structural development of the region. In the eastern Sevier Desert, the intrabasin Tabernacle, Pavant, and Deseret fault systems show markedly different surface expressions and behavior from the range-bounding normal faults farther west. Pleistocene to Holocene extension rates on faults in the eastern Sevier Desert are &gt;10× higher than those on their western counterparts. Faults here are co-located with Late Pleistocene to Holocene volcanic centers, have events temporally clustered around the timing of Pleistocene volcanism in at least one instance, and have accommodated extension ∼2×–10× above geodetic and long-term geologic rates. We propose a model whereby Pliocene to recent extension in the Sevier Desert is spatially partitioned into an eastern magma-assisted rifting domain, characterized by transient episodes of higher extension rates during volcanism, and a western tectonic-dominated domain, characterized by slower-paced faulting in the Cricket Mountains and House Range and more typical of the “Basin and Range style” that continues westward into Nevada. The Sevier Desert, with near-complete exposure and the opportunity to utilize a range of geophysical instrumentation, provides a globally significant laboratory for understanding the different modes of faulting in regions of continental extension.
APA, Harvard, Vancouver, ISO, and other styles
25

Timmermans, Ann C., Brian L. Cousens, and Christopher D. Henry. "Geochemical study of Cenozoic mafic volcanism in the west-central Great Basin, western Nevada, and the Ancestral Cascades Arc, California." Geosphere 16, no. 5 (July 10, 2020): 1179–207. http://dx.doi.org/10.1130/ges01535.1.

Full text
Abstract:
Abstract Processes linked to shallow subduction, slab rollback, and extension are recorded in the whole-rock major-, trace-element, and Sr, Nd, and Pb isotopic compositions of mafic magmatic rocks in both time and space over southwestern United States. Eocene to Mio-Pliocene volcanic rocks were sampled along a transect across the west-central Great Basin (GB) in Nevada to the Ancestral Cascade Arc (ACA) in the northern Sierra Nevada, California (∼39°–40° latitude), which are interpreted to represent a critical segment of a magmatic sweep that occurred as a result of subduction from east-northeast convergence between the Farallon and North American plates and extension related to the change from a convergent to a transform margin along the western edge of North America. Mafic volcanic rocks from the study area can be spatially divided into three broad regions: GB (5–35 Ma), eastern ACA, and western ACA (2.5–16 Ma). The volcanic products are dominantly calc-alkalic but transition to alkalic toward the east. Great Basin lavas erupted far inland from the continental margin and have higher K, P, Ti, and La/Sm as well as lower (Sr/P)pmn, Th/Rb, and Ba/Nb compared to ACA lavas. Higher Pb isotopic values, combined with lower Ce/Ce* and high Th/Nb ratios in some ACA lavas, are interpreted to come from slab sediment. Mafic lavas from the GB and ACA have overlapping 87Sr/86Sr and 143Nd/144Nd values that are consistent with mantle wedge melts mixing with a subduction-modified lithospheric mantle source. Eastern and western ACA lavas largely overlap in age and elemental and isotopic composition, with the exception of a small subset of lavas from the westernmost ACA region; these lavas show lower 87Sr/86Sr at a given 143Nd/144Nd. Results show that although extension contributes to melting in some regions (e.g., selected lavas in the GB and Pyramid Lake), chemical signatures for most mafic melts are dominated by subduction-related mantle wedge and a lithospheric mantle component.
APA, Harvard, Vancouver, ISO, and other styles
26

Stamatakos, John A., Charles B. Connor, and Ronald H. Martin. "Quaternary Basin Evolution and Basaltic Volcanism of Crater Flat, Nevada, From Detailed Ground Magnetic Surveys of the Little Cones." Journal of Geology 105, no. 3 (May 1997): 319–30. http://dx.doi.org/10.1086/515926.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Hagan, Jeanette C., Cathy J. Busby, Keith Putirka, and Paul R. Renne. "Cenozoic palaeocanyon evolution, Ancestral Cascades arc volcanism, and structure of the Hope Valley–Carson Pass region, Sierra Nevada, California." International Geology Review 51, no. 9-11 (August 12, 2009): 777–823. http://dx.doi.org/10.1080/00206810903028102.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Wallace, Alan R. "Geology of the Ivanhoe Hg-Au District, Northern Nevada: Influence of Miocene Volcanism, Lakes, and Active Faulting on Epithermal Mineralization." Economic Geology 98, no. 2 (April 2003): 409–24. http://dx.doi.org/10.2113/gsecongeo.98.2.409.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Wallace, A. R. "Geology of the Ivanhoe Hg-Au District, Northern Nevada: Influence of Miocene Volcanism, Lakes, and Active Faulting on Epithermal Mineralization." Economic Geology 98, no. 2 (April 1, 2003): 409–24. http://dx.doi.org/10.2113/98.2.409.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Colgan, J. P., T. A. Dumitru, M. McWilliams, and E. L. Miller. "Timing of Cenozoic volcanism and Basin and Range extension in northwestern Nevada: New constraints from the northern Pine Forest Range." Geological Society of America Bulletin 118, no. 1-2 (January 1, 2006): 126–39. http://dx.doi.org/10.1130/b25681.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Canada, Andrew S., Elizabeth J. Cassel, Daniel F. Stockli, M. Elliot Smith, Brian R. Jicha, and Brad S. Singer. "Accelerating exhumation in the Eocene North American Cordilleran hinterland: Implications from detrital zircon (U-Th)/(He-Pb) double dating." GSA Bulletin 132, no. 1-2 (May 16, 2019): 198–214. http://dx.doi.org/10.1130/b35160.1.

Full text
Abstract:
AbstractBasins in orogenic hinterlands are directly coupled to crustal thickening and extension through landscape processes and preserve records of deformation that are unavailable in footwall rocks. Following prolonged late Mesozoic–early Cenozoic crustal thickening and plateau construction, the hinterland of the Sevier orogen of western North America underwent late Cenozoic extension and formation of metamorphic core complexes. While the North American Cordillera is one of Earth’s best-studied orogens, estimates for the spatial and temporal patterns of initial extensional faulting differ greatly and thus limit understanding of potential drivers for deformation. We employed (U-Th)/(He-Pb) double dating of detrital zircon and (U-Th)/He thermochronology of detrital apatite from precisely dated Paleogene terrestrial strata to quantify the timing and magnitude of exhumation and explore the linkages between tectonic unroofing and basin evolution in northeastern Nevada. We determined sediment provenance and lag time evolution (i.e., the time between cooling and deposition, which is a measure of upper-crustal exhumation) during an 8 m.y. time span of deposition within the Eocene Elko Basin. Fluvial strata deposited between 49 and 45 Ma yielded Precambrian (U-Th)/He zircon cooling ages (ZHe) with 105–740 m.y. lag times dominated by unreset detrital ages, suggesting limited exhumation and Proterozoic through early Eocene sediment burial (&lt;4–6 km) across the region. Minimum nonvolcanic detrital ZHe lag times decreased to &lt;100 m.y. in 45–43 Ma strata and to &lt;10 m.y. in 43–41 Ma strata, illustrating progressive and rapid hinterland unroofing in Eocene time. Detrital apatite (U-Th)/He ages present in ca. 44 and 39 Ma strata record Eocene cooling ages with 1–20 m.y. lag times. These data reflect acceleration of basement exhumation rates by &gt;1 km/m.y., indicative of rapid, large-magnitude extensional faulting and metamorphic core complex formation. Contemporaneous with this acceleration of hinterland exhumation, syntectonic freshwater lakes developed in the hanging wall of the Ruby Mountains–East Humboldt Range metamorphic core complex at ca. 43 Ma. Volcanism driven by Farallon slab removal migrated southward across northeastern Nevada, resulting in voluminous rhyolitic eruptions at 41.5 and 40.1 Ma, and marking the abrupt end of fluvial and lacustrine deposition across much of the Elko Basin. Thermal and rheologic weakening of the lithosphere and/or partial slab removal likely initiated extensional deformation, rapidly unroofing deeper crustal levels. We attribute the observed acceleration in exhumation, expansion of sedimentary basins, and migrating volcanism across the middle Eocene to record the thermal and isostatic effects of Farallon slab rollback and subsequent removal of the lowermost mantle lithosphere.
APA, Harvard, Vancouver, ISO, and other styles
32

Titus, Sarah J., Ryan Clark, and Basil Tikoff. "Geologic and geophysical investigation of two fine-grained granites, Sierra Nevada Batholith, California: Evidence for structural controls on emplacement and volcanism." Geological Society of America Bulletin 117, no. 9 (2005): 1256. http://dx.doi.org/10.1130/b25689.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Millar, Constance I., John C. King, Robert D. Westfall, Harry A. Alden, and Diane L. Delany. "Late Holocene forest dynamics, volcanism, and climate change at Whitewing Mountain and San Joaquin Ridge, Mono County, Sierra Nevada, CA, USA." Quaternary Research 66, no. 2 (September 2006): 273–87. http://dx.doi.org/10.1016/j.yqres.2006.05.001.

Full text
Abstract:
AbstractDeadwood tree stems scattered above treeline on tephra-covered slopes of Whitewing Mtn (3051 m) and San Joaquin Ridge (3122 m) show evidence of being killed in an eruption from adjacent Glass Creek Vent, Inyo Craters. Using tree-ring methods, we dated deadwood to AD 815–1350 and infer from death dates that the eruption occurred in late summer AD 1350. Based on wood anatomy, we identified deadwood species as Pinus albicaulis, P. monticola, P. lambertiana, P. contorta, P. jeffreyi, and Tsuga mertensiana. Only P. albicaulis grows at these elevations currently; P. lambertiana is not locally native. Using contemporary distributions of the species, we modeled paleoclimate during the time of sympatry to be significantly warmer (+3.2°C annual minimum temperature) and slightly drier (−24 mm annual precipitation) than present, resembling values projected for California in the next 70–100 yr.
APA, Harvard, Vancouver, ISO, and other styles
34

Henry, Christopher D., and James E. Faulds. "Ash-flow tuffs in the Nine Hill, Nevada, paleovalley and implications for tectonism and volcanism of the western Great Basin, USA." Geosphere 6, no. 4 (August 2010): 339–69. http://dx.doi.org/10.1130/ges00548.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Resmini, Ronald G., and Bruce D. Marsh. "Steady-state volcanism, paleoeffusion rates, and magma system volume inferred from plagioclase crystal size distributions in mafic lavas: Dome Mountain, Nevada." Journal of Volcanology and Geothermal Research 68, no. 4 (November 1995): 273–96. http://dx.doi.org/10.1016/0377-0273(95)00003-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Wells, S. G., L. D. McFadden, C. E. Renault, and B. M. Crowe. "Geomorphic assessment of late Quaternary volcanism in the Yucca Mountain area, southern Nevada: Implications for the proposed high-level radioactive waste repository." Geology 18, no. 6 (1990): 549. http://dx.doi.org/10.1130/0091-7613(1990)018<0549:gaolqv>2.3.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Benson, Thomas R., Gail A. Mahood, and Marty Grove. "Geology and 40Ar/39Ar geochronology of the middle Miocene McDermitt volcanic field, Oregon and Nevada: Silicic volcanism associated with propagating flood basalt dikes at initiation of the Yellowstone hotspot." GSA Bulletin 129, no. 9-10 (June 23, 2017): 1027–51. http://dx.doi.org/10.1130/b31642.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Brueseke, M. E., J. S. Callicoat, W. Hames, and P. B. Larson. "Mid-Miocene rhyolite volcanism in northeastern Nevada: The Jarbidge Rhyolite and its relationship to the Cenozoic evolution of the northern Great Basin (USA)." Geological Society of America Bulletin 126, no. 7-8 (April 1, 2014): 1047–67. http://dx.doi.org/10.1130/b30736.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Wannamaker, Philip E., Jeffery M. Johnston, John A. Stodt, and John R. Booker. "Anatomy of the southern Cordilleran hingeline, Utah and Nevada, from deep electrical resistivity profiling." GEOPHYSICS 62, no. 4 (July 1997): 1069–86. http://dx.doi.org/10.1190/1.1444208.

Full text
Abstract:
To address outstanding questions in Mesozoic‐Cenozoic structure and present‐day deep physicochemical state in the region of the southern Cordilleran hingeline, a detailed, east‐west profile of magnetotelluric (MT) soundings 155 km in length was acquired. From these soundings, a resistivity interpretation was produced using an inversion algorithm based on a structural parameterization. In the upper ten kilometers of the transect, the interpretation shows two segments of low resistivity lying beneath allochthonous rocks of the Late Mesozoic, Sevier thrust sheet. Subsequent industry drilling motivated in part by our surveying confirms the existence and position of the eastern subthrust conductor and, more spectacularly, identifies the presence of yet deeper, autochthonous Mesozoic rocks. The conductors cannot be specified uniquely with present public data, because their electrical characteristics appear consistent with Paleozoic, pyrolized graphitic strata of either Late Devonian‐Mississippian or Middle Ordovician age. However, the drilling results show that Late Paleozoic and younger rocks lie underthrust much farther west than recognized previously, and perhaps as far west as the Utah‐Nevada border. A simple structural interpretation is offered where one underthrust segment of low‐resistivity sediments was created originally, but this segment was broken later into two major ones during higher‐angle Tertiary extension. For the middle and lower crust, the MT data imply a nearly 1-D resistivity structure of remarkable uniformity across the entire transect. In particular, there occurs a deep low‐resistivity layer most pronounced (about 8 ohm-m) in the nominal depth interval of 17.5 to 40 km. The MT data indicate that the layer cannot be confined to a single thin layer in the lower crust but instead represents vertically distributed low resistivity. With temperatures estimated from surface heat flow to range from 550°C to 1050°C with depth in the layer, and with a metaigneous mineralogy of high metamorphic grade assumed, mechanisms to produce the low resistivity can be constrained. The deep layer is thus consistent with [Formula: see text] brines at its upper levels, fluids of lower [Formula: see text] activity toward middle levels, and [Formula: see text] melting below about 30 km. The marked uniformity of the deep conductive layer across the transect suggests a similar uniformity of deep physicochemical state. However, this is not at odds with recent analyses of heat flow, Curie depth, Quaternary extension, and basaltic volcanism. Pre‐existing structural fabrics have had no measureable influence on localizing regions of high temperature, fluids and melting in the lower crust, at least averaged over the scale of tens of kilometers. Given its uniformity over a distance of 155 km or more, the depth to the regional deep conductor does not appear related to the distribution of high‐temperature geothermal resources.
APA, Harvard, Vancouver, ISO, and other styles
40

Cousens, Brian L., Christopher D. Henry, Bradley J. Harvey, Tim Brownrigg, Julie Prytulak, and James F. Allan. "Secular variations in magmatism during a continental arc to post-arc transition: Plio-Pleistocene volcanism in the Lake Tahoe/Truckee area, Northern Sierra Nevada, California." Lithos 123, no. 1-4 (April 2011): 225–42. http://dx.doi.org/10.1016/j.lithos.2010.09.009.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Bernardino, Melissa V., Craig H. Jones, William Levandowski, Ian Bastow, Thomas J. Owens, and Hersh Gilbert. "A multicomponent Isabella anomaly: Resolving the physical state of the Sierra Nevada upper mantle from Vp/Vs anisotropy tomography." Geosphere 15, no. 6 (November 8, 2019): 2018–42. http://dx.doi.org/10.1130/ges02093.1.

Full text
Abstract:
Abstract The Isabella anomaly, a prominent upper-mantle high-speed P-wave anomaly located within the southern Great Valley and southwestern foothills of the Sierra Nevada, has been interpreted either as foundering sub-Sierran lithosphere or as remnant oceanic lithosphere. We used Vp/Vs anisotropy tomography to distinguish among the probable origins of the Isabella anomaly. S waveforms were rotated into the Sierran SKSFast and SKSSlow directions determined from SKS-splitting studies. Teleseismic P-, SFast-, SSlow-, SKSFast-, and SKSSlow-wave arrival times were then inverted to obtain three-dimensional (3-D) perturbations in Vp, Vp/VsMean, and percent azimuthal anisotropy using three surface wave 3-D starting models and one one-dimensional (1-D) model. We observed the highest Vp/Vs anomalies associated with slower velocities in regions marked by young volcanism, with the largest of these anomalies being the Mono anomaly under the Long Valley region, which extends to depths of at least 75 km. Peak Vp/Vs perturbations of +4% were found at 40 km depth. The low velocities and high Vp/Vs values of this anomaly could be related to partial melt. The high wave speeds of the Isabella anomaly coincide with low Vp/Vs values with peak perturbations of −2%, yet they do not covary spatially. The P-wave inversion imaged the Isabella anomaly as a unimodal eastward-plunging body. However, the volume of that Isabella anomaly contains three separate bodies as defined by varying Vp/Vs values. High speeds, regionally average Vp/Vs values (higher than the other two anomalies), and lower anisotropy characterize the core of the Isabella anomaly. The western and shallowest part has high wave speeds and lower Vp/Vs values than the surrounding mantle. The eastern and deepest part of the anomaly also contains high speeds and lower Vp/Vs values but exhibits higher anisotropy. We considered combinations of varying temperature, Mg content (melt depletion), or modal garnet to reproduce our observations. Our results suggest that the displaced garnet-rich mafic root of the Mesozoic Sierra Nevada batholith is found in the core of the Isabella anomaly. If remnant oceanic lithosphere exists within the Isabella anomaly, it most likely resides in the shallow, westernmost feature. Within the Sierra Nevada, the highest upper-mantle anisotropy is largely contained within the central portion of the range and the adjacent Great Valley. Anisotropy along the Sierra crest is shallow and confined to the lithosphere between 20 and 40 km depth. Directly below, there is a zone of low anisotropy (from 170 to 220 km depth), low velocities, and high Vp/Vs values. These features suggest the presence of vertically upwelling asthenosphere and consequent horizontal flow at shallower depths. High anisotropy beneath the adjacent western foothills and Great Valley is found at ∼120 km depth and could represent localized mantle deformation produced as asthenosphere filled in a slab gap.
APA, Harvard, Vancouver, ISO, and other styles
42

Whitney, John W., Ralph R. Shroba, S. G. Wells, L. D. McFadden, C. E. Renault, and B. M. Crowe. "Comment and Reply on "Geomorphic assessment of late Quaternary volcanism in the Yucca Mountain area, southern Nevada: Implications for the proposed high-level radioactive waste repository"." Geology 19, no. 6 (1991): 661. http://dx.doi.org/10.1130/0091-7613(1991)019<0661:caroga>2.3.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Rouer, O., H. Lapierre, C. Coulon, and A. Michard. "New petrological and geochemical data on mid-Paleozoic island-arc volcanics of northern Sierra Nevada, California: evidence for a continent-based island arc." Canadian Journal of Earth Sciences 26, no. 12 (December 1, 1989): 2465–78. http://dx.doi.org/10.1139/e89-210.

Full text
Abstract:
The mid-Paleozoic volcanics of northern Sierra Nevada consist of the Sierra Buttes rhyolites, the Taylor basalts and andesites, and the Keddie Ridge basalt–latite–rhyolite suite. The Sierra Buttes calc-alkaline rhyolites display strong light rare-earth element enrichment and negative εNd values. The Taylor basalts and andesites in the northern Hough and Genesee blocks exhibit calc-alkaline affinities (REE rare-earth element patterns highly enriched in LREE), whereas in the southern Hough block they are tholeiitic (flat rare-earth element patterns). The abundance of silicic lavas, the low εNd values of both the Sierra Buttes and Taylor volcanics and the δ18O values of the Sierra Buttes rhyolite and Bowman Lake trondjhemite provide evidence that the northern Sierra Nevada island arc was continent based. The Keddie Ridge differentiated volcanics, characterized by high Zr, Y, Nb, K, and light rare-earth elements, are geochemically similar to a shoshonite suite. Their eruption at the end of the mid-Paleozoic volcanic episode suggests a reversal of subduction, uplift, and block faulting in the island arc.The mid-Paleozoic volcanics of the northern Sierra Nevada are thought to represent the remnant of a mature island arc because calc-alkaline rocks predominate over tholeiitic ones, the lavas display a K enrichment with time, and the volcanics are evolved in their isotopes, compared with rocks erupted in young or primitive island arcs.
APA, Harvard, Vancouver, ISO, and other styles
44

Coble, Matthew A., and Gail A. Mahood. "Geology of the High Rock caldera complex, northwest Nevada, and implications for intense rhyolitic volcanism associated with flood basalt magmatism and the initiation of the Snake River Plain–Yellowstone trend." Geosphere 12, no. 1 (December 23, 2015): 58–113. http://dx.doi.org/10.1130/ges01162.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Tilling, R. I. "Volcanism and associated hazards: the Andean perspective." Advances in Geosciences 22 (December 14, 2009): 125–37. http://dx.doi.org/10.5194/adgeo-22-125-2009.

Full text
Abstract:
Abstract. Andean volcanism occurs within the Andean Volcanic Arc (AVA), which is the product of subduction of the Nazca Plate and Antarctica Plates beneath the South America Plate. The AVA is Earth's longest but discontinuous continental-margin volcanic arc, which consists of four distinct segments: Northern Volcanic Zone, Central Volcanic Zone, Southern Volcanic Zone, and Austral Volcanic Zone. These segments are separated by volcanically inactive gaps that are inferred to indicate regions where the dips of the subducting plates are too shallow to favor the magma generation needed to sustain volcanism. The Andes host more volcanoes that have been active during the Holocene (past 10 000 years) than any other volcanic region in the world, as well as giant caldera systems that have produced 6 of the 47 largest explosive eruptions (so-called "super eruptions") recognized worldwide that have occurred from the Ordovician to the Pleistocene. The Andean region's most powerful historical explosive eruption occurred in 1600 at Huaynaputina Volcano (Peru). The impacts of this event, whose eruptive volume exceeded 11 km3, were widespread, with distal ashfall reported at distances >1000 km away. Despite the huge size of the Huaynaputina eruption, human fatalities from hazardous processes (pyroclastic flows, ashfalls, volcanogenic earthquakes, and lahars) were comparatively small owing to the low population density at the time. In contrast, lahars generated by a much smaller eruption (<0.05 km3) in 1985 of Nevado del Ruiz (Colombia) killed about 25 000 people – the worst volcanic disaster in the Andean region as well as the second worst in the world in the 20th century. The Ruiz tragedy has been attributed largely to ineffective communications of hazards information and indecisiveness by government officials, rather than any major deficiencies in scientific data. Ruiz's disastrous outcome, however, together with responses to subsequent hazardous eruptions in Chile, Colombia, Ecuador, and Peru has spurred significant improvements in reducing volcano risk in the Andean region. But much remains to be done.
APA, Harvard, Vancouver, ISO, and other styles
46

De Angelis, M., J. Simões, H. Bonnaveira, J. D. Taupin, and R. J. Delmas. "Volcanic eruptions recorded in the Illimani ice core (Bolivia): 1918–1998 and Tambora periods." Atmospheric Chemistry and Physics 3, no. 5 (October 16, 2003): 1725–41. http://dx.doi.org/10.5194/acp-3-1725-2003.

Full text
Abstract:
Abstract. Acid layers of volcanic origin detected in polar snow and ice layers are commonly used to document past volcanic activity on a global scale or, conversely, to date polar ice cores. Although most cataclysmic eruptions of the last two centuries (Pinatubo, El Chichon, Agung, Krakatoa, Cosiguina, Tambora, etc.) occurred in the tropics, cold tropical glaciers have not been used for the reconstruction of past volcanism. The glaciochemical study of a 137 m ice core drilled in 1999 close to the summit of Nevado Illimani (Eastern Bolivian Andes, 16°37' S, 67°46' W, 6350 m asl) demonstrates, for the first time, that such eruptions are recorded by both their tropospheric and stratospheric deposits. An 80-year ice sequence (1918-1998) and the Tambora years have been analyzed in detail. In several cases, ash, chloride and fluoride were also detected. The ice records of the Pinatubo (1991), Agung (1963) and Tambora (1815) eruptions are discussed in detail. The potential impact of less important regional eruptions is discussed.
APA, Harvard, Vancouver, ISO, and other styles
47

Scuderi, Louis A. "Tree-Ring Evidence for Climatically Effective Volcanic Eruptions." Quaternary Research 34, no. 1 (July 1990): 67–85. http://dx.doi.org/10.1016/0033-5894(90)90073-t.

Full text
Abstract:
AbstractRingwidth variations from temperature-sensitive upper timberline sites in the Sierra Nevada show a marked correspondence to the decadal pattern of volcanic sulfate aerosols recorded in a Greenland ice-core acidity profile and a significant negative growth response to individual explosive volcanic events. The appearance of single events in the mid-latitude tree-ring record, in connection with ice-core evidence from the arctic and historical records from the Mediterranean, indicates that the majority of these events represent climatically effective volcanic eruptions, producing temperature decreases on the order of 1°C for up to 2 yr after the initial eruption. Clusters of climatically effective volcanic events may serve as a trigger to glaciation and are consistently associated with lowered ringwidths and late-Holocene glacier advance in the Sierra Nevada. The tree-ring record strongly suggests forcing of solar radiation receipt and temperatures by increased volcanic aerosols, especially during the Recess Peak advances and Matthes (Little Ice Age) advances from 1400 to 1850 A.D. Intervals with an absence of significant volcanic aerosol production or historically documented eruptive activity correspond to intervals of significantly increased indexed ringwidth values, minimal numbers of severe annual negative ringwidth anomalies, and an absence of glacial deposits in the southern Sierra Nevada.
APA, Harvard, Vancouver, ISO, and other styles
48

Huggel, Christian, Jorge Luis Ceballos, Bernardo Pulgarĺn, Jair Ramírez, and Jean-Claude Thouret. "Review and reassessment of hazards owing to volcano–glacier interactions in Colombia." Annals of Glaciology 45 (2007): 128–36. http://dx.doi.org/10.3189/172756407782282408.

Full text
Abstract:
AbstractThe Cordillera Central in Colombia hosts four important glacier-clad volcanoes, namely Nevado del Ruiz, Nevado de Santa Isabel, Nevado del Tolima and Nevado del Huila. Public and scientific attention has been focused on volcano–glacier hazards in Colombia and worldwide by the 1985 Nevado del Ruiz/Armero catastrophe, the world’s largest volcano–glacier disaster. Important volcanological and glaciological studies were undertaken after 1985. However, recent decades have brought strong changes in ice mass extent, volume and structure as a result of atmospheric warming. Population has grown and with it the sizes of numerous communities located around the volcanoes. This study reviews and reassesses the current conditions of and changes in the glaciers, the interaction processes between ice and volcanic activity and the resulting hazards. Results show a considerable hazard potential from Nevados del Ruiz, Tolima and Huila. Explosive activity within environments of snow and ice as well as non-eruption-related mass movements induced by unstable slopes, or steep and fractured glaciers, can produce avalanches that are likely to be transformed into highly mobile debris flows. Such events can have severe consequences for the downstream communities. Integrated monitoring strategies are therefore essential for early detection of emerging activity that may result in hazardous volcano–ice interaction. Corresponding efforts are currently being strengthened within the framework of international programmes.
APA, Harvard, Vancouver, ISO, and other styles
49

De Angelis, M., J. Simões, H. Bonnaveira, J. D. Taupin, and R. J. Delmas. "Volcanic eruptions recorded in the Illimani ice core (Bolivia): 1918–1998 and Tambora periods." Atmospheric Chemistry and Physics Discussions 3, no. 3 (May 15, 2003): 2427–63. http://dx.doi.org/10.5194/acpd-3-2427-2003.

Full text
Abstract:
Abstract. Acid layers of volcanic origin detected in polar snow and ice layers are commonly used to document past volcanic activity on a global scale or, conversely, to date polar ice cores. Although most cataclysmic eruptions of the last two centuries (Pinatubo, El Chichon, Agung, Krakatoa, Cosiguina, Tambora, etc.) occurred in the tropics, cold tropical glaciers have not been used for the reconstruction of past volcanism. The glaciochemical study of a 137 m ice core drilled in 1999 close to the summit of Nevado Illimani (Eastern Bolivian Andes, 16°37' S, 67°46' W, 6350 m a.s.l.) demonstrates, for the first time, that such eruptions are recorded by both their tropospheric and stratospheric deposits. An 80-year ice sequence (1918–1998) and the Tambora years have been analyzed in detail. In several cases, ash, chloride and fluoride were also detected. The ice records of the Pinatubo (1991), Agung (1963) and Tambora (1815) eruptions are discussed in detail. Less important eruptions located in the Andes are also recorded and may also disturb background aerosol composition on a regional scale.
APA, Harvard, Vancouver, ISO, and other styles
50

Picard, M. "Remembering First Oil in Nevada." Earth Sciences History 28, no. 2 (November 5, 2009): 161–74. http://dx.doi.org/10.17704/eshi.28.2.3568120856325474.

Full text
Abstract:
In June 1954 Nevada became the twenty-ninth oil-producing state in the United States (Picard 1955). Interestingly, production was from volcanic rocks from the open-hole interval 6,450 to 6,730 ft (1,966 to 2,051 m) in the Oligocene Garrett Ranch volcanics, an unexpected reservoir in the kind of rocks rarely productive anywhere in the world. The pour-point (65-80° F) and gravity (26-29° API) of the crude were high, similar to oils found in the Eocene Green River Formation of the Uinta Basin, northeast Utah. Cumulative production in the field through September 1978 was 3.3 million barrels of oil. An early estimate of ultimate primary reserves was four million barrels of oil (Bortz and Murray, 1979). The trap is a faulted truncated wedge of Oligocene and Cretaceous-Eocene rocks with a top seal of impermeable valley fill, a bottom seal of Paleozoic rocks, and an east-side seal formed by a basin boundary fault and impermeable Paleozoic rocks. The new field in Railroad Valley of east-central Nevada, finally totaling fourteen producing wells, was called Eagle Springs after the locality and the name of the discovery well drilled by the Shell Oil Company. Twenty-two years after the Eagle Springs discovery a larger oil field, Trap Spring, was discovered by Northwest Exploration Company less than ten miles west of Eagle Springs, in Tertiary ash-flow tuffs. Two hundred dry holes had been drilled in Nevada between the two discoveries. In 1982, six years after the Trap Spring discovery, Amoco Production Company drilled the first well outside of Railroad Valley at Blackburn field on the east side of Pine Valley in Eureka County. Blackburn, a structural trap above a Tertiary low-angle extensional fault, produces from Devonian reservoirs. In 1983, Northwest Production brought in the Grant Canyon field about 10 mi (6 km) south of Eagle Springs. The oil reservoir of Devonian carbonates there is entrapped in a ‘buried-hill’. The discovery in 2004 of the Covenant field in Central Utah, because of similarities to large oil fields in the thrust belt of Wyoming and Utah and some resemblance to the Nevada fields of the Great Basin, ignited a frenzy of leasing which still goes on when land is available. Located along the thrust-belt (hingeline), Covenant produces oil from the Jurassic Navajo Sandstone that apparently originated in the Paleozoic.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography