National Academies Press: OpenBook

Ground Water at Yucca Mountain: How High Can It Rise? (1992)

Chapter: Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?

« Previous: Might Increased Rainfall Cause Flooding of the Proposed Repository?
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

4

Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?

SUMMARY OF TERTIARY VOLCANIC HISTORY OF THE REGION

The late Tertiary geologic history of southwestern Nevada has been dominated by volcanism and the consequent deposition of volcanic flows and tuffaceous rocks. Yucca Mountain, like most surrounding ranges, is composed dominantly of a series of well-studied, Miocene ashflow tuff units and silicic volcanic rocks. The tuffs erupted between 17 and 7 Ma from multiple caldera complexes located throughout the region (Christiansen et al., 1977; Byers et al., 1976). The uppermost unit of these widespread eruptive tuff deposits exposed on Yucca Mountain is the Tiva Canyon member of the Timber Mountain caldera complex, dated at about 12.5 Ma.

Concurrent with the silicic episode, large-volume olivine basalt lava flows erupted between 11 and 8 Ma near and within some of the silicic centers in the Yucca Mountain region (Crowe et al., 1983a and 1986). In roughly the same time period (6-10 Ma), other basalt and basaltic andesite flows erupted that appear unrelated to the silicic centers (the “older rift basalts” of Crowe et al. (1986)). These basaltic flows probably mark the beginning of “modern” basin and range extension in this region of southern Nevada. Modern basin and range extension refers to the stress and faulting regime, still active today, which is responsible for the formation of the modern range and basin blocks. It is generally dated between 10 and 7 Ma throughout the northern Basin and Range region (see, e.g., Stewart, 1978; Zoback and Thompson, 1978; Zoback et al., 1981).

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

The next and last-recorded volcanism in the Yucca Mountain region involved flows and cinder cones of young basalt that erupted in Crater Flat, which borders Yucca Mountain on the west, between 3.7 and 0.13 Ma (the “younger rift basalts” of Crowe et al. (1986)). The alignment of cones suggests they were fed from dikes as is commonly the case elsewhere (see, e.g., Nakamura, 1977). These younger basalts interfinger with and are overlain by unconsolidated to semi-consolidated clastic (fragmental) material deposited in a variety of environments in response to extensional normal faulting.

STYLE AND SIZE OF A LIKELY INTRUSION IN THE YUCCA MOUNTAIN REGION

We concur with the conclusion reached by previous researchers (Crowe et al., 1983b; DOE, 1988) that the possibility of future large silicic caldera-forming eruptions can be dismissed from consideration in the Yucca Mountain area. The Miocene caldera complexes formed because of their location in a continental intra-arc setting close to, and overlying, an active east-directed subduction zone along the western continental margin. The subsequent evolution of the San Andreas transform fault system (see, e.g., Atwater, 1970) replaced the subduction zone along much of the western margin and changed the tectonic regime in Nevada. The current southernmost boundary of active subduction below western North America is the Mendocino triple junction at approximately 40°N, more than 300 km north of the latitude of Yucca Mountain. The present day volcano-tectonic regime throughout the northern Basin and Range province, which includes most of Nevada and the western half of Utah, is extensional, characterized by low-volume basaltic eruptions, such as those in Crater Flat (see, e.g., Christiansen and Lipman, 1972).

The age of the basaltic volcanism in Crater Flat remains the subject of considerable debate. Isotopic ages between 3.8 and 0.3 Ma have been obtained from the cinder cone centers in Crater Flat by Turrin and Champion, (1991). However, Wells et al. (1990) have argued that a nearby basaltic center, the Lathrop Wells cone located about 20 km south of Yucca Mountain, may be as young as 20 ka based on geomorphic and pedogenic characteristics as well as on the scatter of isotopic ages. Wells et al. (1990) further suggest that this center contains at least three discrete and temporally separate eruptive events that may have occurred over time spans of 1–10 ka, based on mapping of stratigraphic relations of tephra (volcanic debris) units here and elsewhere in the Basin and Range (Crowe et al., 1989). In contrast, 40Ar/39Ar age dating of two separate flow units in the Lath-

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

rop Wells volcanic center yields arithmetic mean of ages of 183 ±21 and 144 ±35 ka (Turrin et al., 1991). On the basis of this dating and as yet unpublished K/Ar dates, Turrin et al. (1991) conclude that there were two eruptive events at Lathrop Wells, dated at 136 ±8 ka and 141 ±9 ka. They speculate that the time interval between flows may be less than 100 years because field mapping and paleomagnetic data indicate remanent magnetization directions only a few degrees apart for the two flow units. Differences in remanent magnetization directions can be accounted for by secular (temporal) variation of the earth 's magnetic field, the rates of which have been calibrated in other volcanic fields at approximately 4° per 100 years. Thus, they interpreted the nearly identical remanent directions in the two Lathrop Wells flow units to imply a short duration (<100 years) of eruptive activity. However, this interpretation of the paleomagnetic data is controversial. Because both remanent directions are very similar to the time-averaged geomagnetic field in the study area, these directions could represent equally well eruptions separated by 100 years, 10 ka, 100 ka, or 1 Ma.

The geologic record thus indicates that the only likely style of intrusion in the Yucca Mountain region in the lifetime of the repository is a low-volume basaltic dike intrusion. To place constraints on the size and geometry of such intrusions we can look at the occurrence and distribution of other Quaternary basaltic dikes, flows, and cinder cones throughout the northern Basin and Range province. Most of this young basaltic activity occurs within the modern basins, with many of the eruptive centers occurring along basin margins, particularly localized near range-bounding faults (Smith et al., 1990). In virtually all cases, including the basaltic centers in Crater Flat, the alignment of cinder cones, fissures, and other eruptive centers trends N to NNE. This is the direction of the regional maximum horizontal principal stress throughout the Quaternary, as established by earthquake focal mechanisms and fault striation studies (Zoback, 1989). The N to NNE trend thus appears to coincide with a principal stress plane, as expected theoretically (Nakamura, 1977; Nakamura et al., 1978). There is little or no shear stress across these planes. Models in which basaltic volcanism is localized along (or controlled by) major crustal shear zones have little or no corroborating analogs, either theoretically or in field examples.

Some basalt intrusion has occurred within what are now modern range blocks. A dike exposed along the Solitario Canyon normal fault where it cuts into the Yucca Mountain range block NW of the proposed respository site has been dated at 10 Ma (C. Fridrich, pers. comm., 1991). This dike is probably part of the the “older rift basalt”

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

sequence that erupted in the early phases of modern basin and range extension. Models that fit ground magnetic highs located further to the south adjacent to Solitario Canyon include a vertical zone of increased magnetization which could be interpreted as a basaltic dike (Bath and Jahren, 1985); however, there is no surface expression of such a dike in this region. Bath and Jahren (1985) did report a personal communication from R. B. Scott, USGS, that weathered basalt fragments had been found at a lower elevation and about 0.8 km south of the magnetic anomaly. However, despite excellent and abundant outcrops (especially scarp slopes of ridges) and core from five continuously-cored drill holes, no dikes younger than 10 Ma have been found within the Yucca Mountain block. (See also R.F. Marvin, written comm., 1980 in Carr et al. (1986)). However, the drill holes were vertical as are most of the dikes.

Dikes produced during the most recent episode of basaltic volcanism near Yucca Mountain are typically 2 m or less in width (Bruce Crowe, pers. comm. in Carrigan et al., 1990). The maximum width of the 10 Ma dike exposed along the Solitario Canyon fault mentioned above was also 2 m (Chris Fridrich, pers. comm., 1991). Theoretical consideration of elastic strains in surrounding rock in response to normal stresses caused by dike widening constrains dike dimensions (Davis, 1983). A brittle crust 10-15 km thick should limit dike width in the upper crust to 2-4 m (Carrigan et al., 1990). Thus, theoretical considerations, as well as structural and geologic evidence from the Yucca Mountain region and the surrounding Basin and Range province, suggest a likely geometry for basaltic dike intrusion in the upper crust in the vicinity of the repository to be a 2-4 m wide (or less) near-vertical zone of intrusion of magma initially at about 1200°C.

A recent analysis of seismicity in regions of young volcanism cited the Crater Flat area as one example of a number of young basaltic centers characterized by a paucity of background seismicity and low rates of recent normal faulting (Parsons and Thompson, 1991). Parsons and Thompson suggest that ongoing regional extensional strain accumulation is released in these areas primarily by expansion of the crust due to dike intrusion, rather than by normal faulting. Thus, the present-day lack of seismicity down to very small magnitude ranges (M≃0) in the vicinity of Yucca Mountain (Gomberg, 1991) may be an indication of either active dike intrusion at depth or the fact that the regional strain accumulation was largely relieved by the Quaternary basaltic volcanic events. Possible evidence for molten rock below Crater Flat is discussed later in this chapter.

Estimating the accumulated extensional strain since the last basaltic episode (20 ka or 140 ka), using an average Quaternary extension-

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

al rate of 0.01 mm/yr (based on estimates of Quaternary offset of faults in the Yucca Mountain region, (Scott, 1990), results in an accumulated extension on the order of 0.2 m (for a 20 ka event) or 1.4 m (for a 140 ka event). This strain could be relieved either by extensional faulting or by dike intrusion. A 1.4 m dike width is consistent with the other estimates of basaltic dike widths in the Yucca Mountain region as discussed above. It should be noted, however, that not all dike intrusions accommodating regional strain necessarily reach the surface or the near-surface. Dikes tend to propagate subhorizontally in the seismogenic (brittle, elastic) crust in the depth range of about 2-10 km (Halls and Fahrig, 1987; Rubin and Pollard, 1987).

MODELS OF WATER TABLE RISE ACCOMPANYING DIKE INTRUSION

Two independent types of models of water level changes associated with dike intrusion have been investigated for Yucca Mountain. Both models assume intrusion along a vertical zone beneath the repository with the top of the intrusion located 250-500 m below the water table. Kuiper (1991) computed the potential water table rise due to the increased temperature effect of the intrusion using both scoping calculations and a two-phase (steam and water) flow model. In an alternate approach, Carrigan et al. (1990) focused solely on the elastic strain effect of a dike intrusion that results in a region of dilation just above the level of dike intrusion and a region of contractional strain centered just below the dike top. Pore-water pressure enhancement was computed from the volumetric strain, and the excess pore-water pressure was used to drive the flow.

The computed response of the water table for these two very different models was less than 25 m for the thermal model of Kuiper and generally 4-6 m for the poro-elastic model of Carrigan et al. Kuiper assumed a 10 km long, 100 m wide vertical disk-shaped zone of temperature rise (related to dike intrusion) with a 1-hour period of temperature increase of 300°C, and included the effects of subsequent cooling. His maximum increase in water table level of less than 25 m was strongly dependent on both the assumed width and the magnitude of the temperature increase related to the zone of intrusion. Thin basaltic dikes intrude into the near surface at approximately 1200°C and cool rapidly. Unfortunately discussion of Kuiper's model was available only in an abstract in which the exact intrusion dimensions assumed to produce the 100 m wide zone of a 300°C temperature increase were not given.

Carrigan et al. (1990) investigated the effect on the water table of

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

intrusion of both a 2 m and a 4 m wide dike beneath the repository. Using uniform permeability values, they found a maximum rise of the water table of 4 m for a 2 m wide dike (6 m for the 4 m wide dike) over a 1 km wide region centered on the dike, occurring 8.6 days after the intrusion. For an assumed vertical permeability that was everywhere 10 times greater than the horizontal permeability, they found a peak of 6.5 m in less than one day for the 2 m dike. They also examined the effect of a tenfold to a thousandfold increase in the vertical permeability (101 to 104) of a local 100 m wide zone centered on the dike, such as might result from cracking induced by the intrusion. Sharp peaks in the water table rise were predicted for the large vertical permeability enhancement. The maximum water table rise of 12 m for the 2 m wide dike (14 m for the 4 m dike) was associated with a 103 increase in vertical permeability. Further increases in vertical permeability resulted in no further increase in the water table level, suggesting that any further rise was limited by the volume of ground water available to be channeled into the zone of enhanced permeability.

POSSIBLE DEEP (LOWER CRUSTAL) MAGMA CHAMBERS IN THE YUCCA MOUNTAIN REGION

Analysis of far-traveled earthquake waves (P-waves) passing nearly vertically through the crust and upper mantle beneath Yucca Mountain and surrounding regions (Evans and Smith, 1992) shows no evidence of a low velocity feature that would suggest a volume of molten rock (or magma chamber) beneath Yucca Mountain. The minimum size of a feature identifiable in that study is 4 km across. However, a volume of slightly lower velocity material extending upward from the Moho (the boundary between the earth's crust and mantle) to within 12 km of the surface was found directly to the west, beneath Crater Flat. The seismic P-wave velocity in this anomalous volume beneath Crater Flat is about three percent lower than material beneath Yucca Mountain in the same depth range. The standard errors of this type of velocity analysis are about 0.4-0.5 percent; the minimum believable velocity change is therefore on the order of 1 percent.

The three percent lower velocity anomaly identified under Crater Flat is considerably smaller in magnitude than velocity anomalies associated with known large-scale high silica magma chambers, such as Long Valley, where velocity decreases of six to eight percent are typically found (Iyer, 1988). The possibility that molten rock is present beneath Crater Flat deserves further study.

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

To establish the presence (or absence) of molten rock currently beneath Crater Flat, a broader and higher resolution analysis would be required. Seismic reflection profiling techniques tuned for identifying the presence of “fluids” should also be included. Such a study would need to include the determination of the velocity of shear (S) waves, which are much more sensitive to the presence of fluids or melts, in addition to P-waves, as well as attenuation studies, which investigate the rate of change of the amplitude of the waves, and can detect the presence of fluids or melts.

PROBABILISTIC ASSESSMENT OF THE LIKELIHOOD OF A BASALTIC DIKE INTRUSION

Several assessments of the probability of a basaltic dike intrusion at Yucca Mountain are available. All rely on dividing the probability assessment into two steps: (1) P1, the probability of basaltic activity in the general region of the proposed repository, including Lathrop Wells and Crater Flat (called the Crater Flat Volcanic Zone by Crowe), and (2) P2, the conditional probability that, given volcanic activity in the area, the assumed basaltic dike would be close enough to affect the proposed repository. The total probability of volcanic activity affecting Yucca Mountain is calculated as the product of the two.

Crowe and his co-workers (1983, 1986, 1991) have estimated P1 to be on the order of 10−6 per year, based on four clusters of events in 3.7 Ma. If the seven identified events in the region are used, the rate is twice this. If volcanic activity in the region is mature and will wane in the future, as suggested by Crowe (1991), the assumption of a stationary rate of occurrence is conservative for future occurrences.

The calculation of P2 depends on the geometry of the proposed repository with respect to the volcanic field, and on the spatial characteristics governing the location of vents and dikes (the size and ratio of length to width of the volcanic field). The trend of the major axis of vents or feeders of contemporaneous eruptive centers within a volcanic field can be estimated from paleovolcanic events; this trend is NNE for the two main eruptive centers in the Crater Flat Volcanic Zone. The trend of dikes is generally perpendicular to the direction of least principal horizontal stress, which is WNW-ESE in the region. Sheridan (1990) developed an illustrative calculational model of these effects and estimated P2 at about 8 × 10−3. Crowe (1991) estimated P2 at about 10−2.

Combining these two probabilities (P1 and P2) leads to a total probability of about 10−8 per year for the occurrence of a basaltic dike that would affect the proposed repository at Yucca Mountain. This is the thresh-

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

old probability for consideration of events that might affect a waste repository in EPA regulations, and is indeed a small number.

Some researchers (e.g., Smith et al., 1990; Trapp, 1989) use other areas within which to define probabilities of occurrence, based either on geometrical areas (Smith et al.) with some consideration of the present regional stress regime or on interpretation of aeromagnetic data (Trapp). Any associated calculations that assume that future volcanic vents are unrelated in space to existing vents have less credibility, in the panel's view. Furthermore, where probabilities are assessed only qualitatively (Smith et al., 1990), comparison to quantitative calculations is not possible. While the uncertainty may be several orders of magnitude in the probability values described here, the panel concludes that the probability of volcanic intrusion into the proposed repository is low.

CONCLUSIONS

Geologic observations indicate that the form of intrusion most likely to affect the repository is a near-vertical dike of basaltic composition, with a width of 2-4 m. Modeling of the pore-elastic effects of such a dike intruded directly beneath the repository (and with a top less than 1 km below the bottom of the repository) results in maximum water table rises of less than 10 m in both an isotropic case, in which rock properties are the same in all directions, and in a case where the vertical permeability is 10 times the horizontal permeability. Concentrating the water rise in a 100 m wide vertical highly permeable zone directly above the dike resulted in a maximum rise of 12-14 m when vertical permeabilities were enhanced 103-104 within the vertical zone. Thus, the panel concludes that the elastic effect of dike intrusion would result in raising the water table no more than a few tens of meters.

Models of the thermal effects on water table rise are highly dependent on the total amount of heat energy released as a result of the intrusion. The modeled 100 m wide zone of 300°C temperature increase produced a water table rise of less than 25 m; however, the number and exact configuration of dikes within this zone was not specified. For a single dike of 2-4 m width, the modeled 300°C temperature rise over a width of 100 m is clearly an overestimate of the heat energy. As an upper bound, considering energy equivalence, the maximum amount of energy available in a 4 m wide dike intruded at an effective temperature of 1400°C (which includes effects of the latent heat of crystallization) is approximately equivalent to a temperature rise of about 60°C, not 300°C, over a 100 m wide

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

zone. In the opinion of the panel, therefore, a 25 m rise in water table is clearly a conservative upper bound estimate for the expected form of intrusion in the Yucca Mountain region.

It is the panel's view that the low probability of a dike forming close enough to affect the proposed repository, combined with the small effect that a basaltic dike intrusion at Yucca Mountain would have on the water table, means that the threat of volcanic effects is sufficiently low that volcanic intrusions can be discounted as potentially disruptive events.

RECOMMENDATIONS

Unfortunately, modeling of the coupled thermal and pore-elastic problem of the effects of dike intrusion on the water table has not been addressed. While the thermal effects may dominate, the panel notes that this modeling could establish the maximum water table excursion due to dike intrusion.

The study plan for the characterization of volcanic features calls for drilling core holes to investigate aeromagnetic anomalies that may represent either buried volcanic centers or intrusive rocks (probably basaltic in composition). No mention is made of studying the core to determine the extent, if any, of hydrothermal activity (temperature variations in adjacent rocks, fluid inclusion studies, hydrothermal alteration products) that may have been induced by intrusion or extrusion of a body of volcanic rock. The panel recommends this type of investigation be added to the work plan.

Further teleseismic studies for better definition of the low velocity zone beneath Crater Flat should be carried out. A combination of P- and S-wave velocity and attenuation studies should constrain both the nature and source (e.g., possible fraction of melt if present) of this velocity anomaly. Although the panel does not consider it likely that either a larger volume or a different style of igneous intrusion will be found that is inconsistent with the recent geologic record, it would be prudent to follow up with an evaluation of this anomaly to ascertain its significance.

REFERENCES

Atwater, T. 1970. Implications of plate tectonics for the Cenozoic evolution of North America. Geological Society of America Bulletin 81: 3513-3536.

Bath, G. D., and C. E. Jahren. 1985. Investigation of an aeromagnetic anomaly on west side of Yucca Mountain, Nye County, Nevada U.S. Geological Survey Open File Report 85-459: 24 pp.

Byers, F. M., W. J. Carr, P. P. Orkild, W. D. Quinivan, and K. A. Sargent. 1976. Volcanic

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

suites and related cauldrons of the Timber Mountain-Oasis Valley caldera complex, southern Nevada. U. S. Geological Survey Professional Paper 919. 70 pp.

Carr, W. J., F. M. Byers, Jr., and P. P. Orkild. 1986. Stratigraphic and volcano-tectonic relations of Crater Flat tuff and some older volcanic units, Nye County, Nevada U. S. Geological Survey Professional Paper 1323. 28 pp.

Carrigan, C. R., G. C. P. King, and G. E. Barr. 1990. A scoping study of water table excursions induced by seismic and volcanic events Lawrence Livermore National Laboratory Report UCRL-ID-105340. 41 pp.

Champion, D. E. 1991. Volcanic episodes near Yucca Mountain as determined by paleomagnetic studies at Lathrop Wells, Crater Flat, and Sleeping Butte, Nevada In Proceedings of Second International Conference on High-Level Radioactive Waste Management Las Vegas, Nevada, pp. 61-67.

Christiansen, R. L., and P. W. Lipman. 1972. Cenozoic volcanism and plate-tectonic evolution of the western United States. Vol. II Late Cenozoic, Philosophical Transactions of Royal Society of London 271: 249-284.

Christiansen, R. L., P. W. Lipman, W. J. Carr, F. M. Byers, P. P. Orkild, and K. A. Sargent. 1977. Timber Mountain-Oasis Valley caldera complex of southern Nevada Geological Society of America Bulletin 88: 943-956.

Crowe, B.M. 1986. Volcanic hazard assessment for disposal of high-level radioactive waste. In Active Tectonics. National Academy Press, pp. 247-260.

Crowe, B. M. 1991. Probability Calculations, Presentation to Nuclear Waste Technical Review Board. March 1.

Crowe, B., D. T. Vaniman, and W. J. Carr. 1983a. Status of volcanic hazards studies for the Nevada Nuclear Waste Storage Investigations Los Alamos National Laboratory Report LA-9325-MS. 47 pp.

Crowe, B., S. Self, D. Vaniman, R. Amos, and F. Perry. 1983b. Aspects of potential magmatic disruption of a high-level radioactive waste repository in southern Nevada Journal of Geology 91: 259-276.

Crowe, B. M., K. H. Wohletz, D. T. Vaniman, E. Gladney, and B. Bower. 1986. Status of volcanic hazards studies for the Nevada Nuclear Waste Storage Investigations. Los Alamos National Laboratory Report LA-9325-MS-2. 101 pp.

Crowe, B. M., B. D. Turrin, S. G. Wells, L. D. McFadden, C. E. Renault, F. V. Perry, C. D. Harrington, and D. E. Champion. 1989. Polycyclic volcanism: A common eruption mechanism of small basaltic centers in the western USA (abs.). New Mexico Bureau of Mines Bulletin 131: 63.

Davis, P. M. 1983. Surface deformation associated with a dipping hydrofracture. Journal of Geophysical Research 88: 5826-5834.

Evans, J. R., and M. Smith. 1992. Taleseismic tomography of the Yucca Mountain region: volcanism and tectonism In Proceedings of the International High-Level Radioactive Waste Conference Las Vegas, Nevada. April.

Halls, H.C. and W. F. Fahrig, eds. 1987. Mafic Dike Swarms. Geological Association of Canada. Special Paper 34.

Iyer, H. M. 1988. Seismological detection and delineation of magma chambers beneath intraplate volcanic centers in western U.S.A. In Modeling of Volcanic Processes. C. Y. King and R. Scarpa, eds. Friedr. Vieweg & Sohn, Braunschweig/Wiesbaden, Germany, pp. 1-56.

Kuiper, L. K. 1991. Water-table rise due to magma intrusion beneath Yucca Mountain, EOS 72: 121.

Nakamura, K. 1977. Volcanoes as possible indicators of tectonic stress orientation— principal and proposal Journal of Volcanology and Geothermal Research 2: 1-16.

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×

Nakamura, K., K. H. Jacob, and J. N. Davies. 1978. Volcanoes as possible indicators of tectonic stress orientation—Aleutians and Alaska. Pure and Applied Geophysics 115: 87-112.

Parsons, T., and Thompson, G. A. 1991. The role of magma overpressure in suppressing earthquakes and topography: worldwide examples. Science 253: 1399-1402.

Rubin, A. M., and D. D. Pollard. 1987. Origins of blade-like dikes in volcanic rift zones. In Volcanism in Hawaii, R. W. Decker, and T. Wright, eds., U.S. Geological Survey Professional Paper 1350, pp. 1449-1470.

Scott, R. B. 1990. Tectonic setting of Yucca Mountain, southwest Nevada. In Basin and Range Extensional Tectonics near the Latitude of Las Vegas, Nevada, B. P. Wernicke, ed. Geological Society America Memoir 176: 251-282.

Smith, E. I., D. L. Feuerback, T. R. Naumann, and J. E. Faulds. 1990. The area of most recent volcanism near Yucca Mountain, Nevada. Implications for Volcanic Risk Assessment, In Proceedings, International High-Level Radioactive Waste Management Conference Las Vegas, Nevada, ANS/ASCE 1: 81-90.

Sheridan, M. F. 1990. Volcano Occurrences, in Demonstration of a Risk-Based Approach to High-Level Waste Repository Evaluation EPRI Report NP-7057, Palo Alto, California. October.

Stewart, J. H. 1978. Basin-range structure in western North America, a review. Geological Society of America Memoir 152: 1-13.

Trapp, J. 1989. Probability of volcanism at Yucca Mountain. Memorandum. United States Nuclear Regulatory Commission. July 14. 9 pp.

Turrin, B. D., and D. E. Champion. 1991. 40Ar/39Ar laser fusion and K-Ar ages from Lathrop Wells, Nevada and Cima, California: The age of the latest volcanic activity in the Yucca Mountain area In Proceedings of the Second International High-Level Radioactive Waste Conference Las Vegas, Nevada, pp. 68-75.

Turrin, B. D., D. E. Champion, and R. J. Fleck. 1991. 40Ar/39Ar age of the Lathrop Wells volcanic center, Yucca Mountain, Nevada Science 253: 654-657.

U. S. Department of Energy. 1988. Site Characterization Plan. Vol. 1 Chap. 1, Section 1.5.1. pp. 1-200-206.

Wells, S. G., L. D. McFadden, C. E. Renault, and B. M. Crowe. 1990. Geomorphic assessment of late Quaternary volcanism in the Yucca Mountain area, southern Nevada: Implications for the proposed high-level radioactive waste repository Geology 18: 549-553.

Zoback, M. L. 1989. State of stress and modern deformation of the Northern Basin and Range province Journal of Geophysical Research 94: 7105-7128.

Zoback, M. L., and G. A. Thompson. 1978. Basin and Range rifting in northern Nevada: Clues from a mid-Miocene rift and its subsequent offsets Geology 6: 111-116.

Zoback, M. L., R. E. Anderson, and G. A. Thompson. 1981. Cenozoic evolution of the state of stress and style of tectonism in the western United States Philosophical Transactions of the Royal Society of London, Series A, 300: 407-434.

Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 93
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 94
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 95
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 96
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 97
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 98
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 99
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 100
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 101
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 102
Suggested Citation:"Can an Igneous Intrusion Raise the Water Table to the Proposed Repository Level?." National Research Council. 1992. Ground Water at Yucca Mountain: How High Can It Rise?. Washington, DC: The National Academies Press. doi: 10.17226/2013.
×
Page 103
Next: Could a Nearby Earthquake Cause Flooding of the Proposed Repository? »
Ground Water at Yucca Mountain: How High Can It Rise? Get This Book
×
 Ground Water at Yucca Mountain: How High Can It Rise?
Buy Paperback | $50.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

The site of a proposed repository for high-level radioactive waste from the nation's nuclear power plants is not at risk of ground water infiltration, concludes this important book. Yucca Mountain, located about 100 miles northwest of Las Vegas, has been proposed as the site for permanent underground disposal of high-level radioactive waste from the nation's civilian nuclear power plants.

To resolve concerns raised by a Department of Energy (DOE) staff scientist concerning the potential for ground water to rise 1,000 feet to the level proposed for the repository, DOE requested this study to evaluate independently the past history and future potential of large upward excursions of the ground water beneath Yucca Mountain.

READ FREE ONLINE

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  7. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!