Skip to main content
Log in

Incorporation of conceptual and parametric uncertainty into radionuclide flux estimates from a fractured granite rock mass

  • Original Paper
  • Published:
Stochastic Environmental Research and Risk Assessment Aims and scope Submit manuscript

Abstract

Detailed numerical flow and radionuclide simulations are used to predict the flux of radionuclides from three underground nuclear tests located in the Climax granite stock on the Nevada Test Site. The numerical modeling approach consists of both a regional-scale and local-scale flow model. The regional-scale model incorporates conceptual model uncertainty through the inclusion of five models of hydrostratigraphy and five models describing recharge processes for a total of 25 hydrostratigraphic–recharge combinations. Uncertainty from each of the 25 models is propagated to the local-scale model through constant head boundary conditions that transfer hydraulic gradients and flow patterns from each of the model alternatives in the vicinity of the Climax stock, a fluid flux calibration target, and model weights that describe the plausibility of each conceptual model. The local-scale model utilizes an upscaled discrete fracture network methodology where fluid flow and radionuclides are restricted to an interconnected network of fracture zones mapped onto a continuum grid. Standard Monte Carlo techniques are used to generate 200 random fracture zone networks for each of the 25 conceptual models for a total of 5,000 local-scale flow and transport realizations. Parameters of the fracture zone networks are based on statistical analysis of site-specific fracture data, with the exclusion of fracture density, which was calibrated to match the amount of fluid flux simulated through the Climax stock by the regional-scale models. Radionuclide transport is simulated according to a random walk particle method that tracks particle trajectories through the fracture continuum flow fields according to advection, dispersion and diffusional mass exchange between fractures and matrix. The breakthrough of a conservative radionuclide with a long half-life is used to evaluate the influence of conceptual and parametric uncertainty on radionuclide mass flux estimates. The fluid flux calibration target was found to correlate with fracture density, and particle breakthroughs were generally found to increase with increases in fracture density. Boundary conditions extrapolated from the regional-scale model exerted a secondary influence on radionuclide breakthrough for models with equal fracture density. The incorporation of weights into radionuclide flux estimates resulted in both noise about the original (unweighted) mass flux curves and decreases in the variance and expected value of radionuclide mass flux.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  • Aban IB, Meerschaert MM, Panorska AK (2006) Parameter estimation methods for the truncated Pareto distribution. J Amer Stat Assoc 101:270–277

    Article  CAS  Google Scholar 

  • Akaike H (1974) A new look at statistical mdoel identification. IEEE Trans Automat Contr AC-19:716–722

    Article  Google Scholar 

  • Andersson J, Dershowitz B, Hermanson J, Meier P, Tullborg E-L, Winberg A (2002) Final report of the TRUE block scale project. 1. Characterization and model development, TR-02-13. Swedish Nuclear Fuel and Waste Management Co. (SKB), Stockholm, Sweden

  • Ballou L (1979) Field permeability measurements, Waste Isolation Projects: FY 1978, UCRL-50050-78. Lawerence Livermore National Laboratory, Livermore, California

  • Bandis SC, Makurat A, Vik G (1985) Predicted and measured hydraulic conductivity of rock joints. In: Proceedings of the international symposium on fundamentals of rock joints, Björdkliden, Norway, September 15–20

  • Bechtel Nevada (2006) A hydrostratigraphic model and alternatives for the groundwater flow and contaminant transport model of Corrective Action Unit 97: Yucca Flat-Climax Mine, Lincoln and Nye Counties, Nevada, DOE/NV/11718-1119

  • Belcher WR (ed) (2004) Death Valley regional ground-water flow system, Nevada and California—hydrgeologic framework and transient ground-water flow model. U.S. Geological Survey Scientific Investigations Report 2004–5205

  • Belcher WR, D’Agnese FA, O’Brien GM (2004) Introduction, Chapter A. In: Belcher WR (ed) Death Valley regional ground-water flow system, Nevada and California—hydrgeologic framework and transient ground-water flow model. U.S. Geological Survey Scientific Investigations Report 2004–5205

  • Beven KJ, Binley AM (1992) The future of distributed models: model calibration and uncertainty prediction. Hydrol Process 6:279–298. doi:10.1002/hyp.3360060305

    Article  Google Scholar 

  • Beven KJ, Freer J (2001) Equifinality, data assimilation, and uncertainty estimation in mechanistic modeling of complex environmental systems using the GLUE methodology. Hydrol. Process 6:279–298

    Article  Google Scholar 

  • Botros F, Hassan AE, Reeves DM, Pohll G (2008) On mapping fracture networks onto continuum. Water Resour Res 44:WO8435. doi:10.1029/2007WR006092

    Article  Google Scholar 

  • Bowen S, Finnegan DL, Thompson JL, Miller CM, Baca PL, Olivaas LF, Goffrion CG, Smith DK, Goishi W, Esser BK, Meadows JW, Namboodiri N, Wild JF (2001) Nevada Test Site radionuclide inventory 1951–1992, LA-13859-MS. Los Alamos National Laboratory, Los Alamos, New Mexico

  • Cook AM, Myer LR, Cook NGW, Doyle FM (1990) The effect of toruosity on flow through a natural fracture, Rock mecdhanics contributions and challenges. In: Hustrulid WA, Johnson GA (eds) Proceedings of the 31st U.S. symposium on rock mechanics. A.A. Balkema, Rotterdam

  • de Dreuzy J-R, Davy P, Bour O (2001) Hydraulic properties of two-dimensional random fracture networks follwoing a power law length distribution. 1. Effective connectivity. Water Resour Res 37(8)2065–2078

    Google Scholar 

  • Dershowitz WJ, Hermanson S, Follin S, Mauldon M (2000) Fracture intensity measures in 1-D, 2-D, and 3-D at Äspö, Sweden. In: Proceedings of the fourth North American rock mechanics symposium, Pacific Rocks 2000, vol 4, Seattle, Washington, pp 849–853

  • Doherty J (2000) PEST—model-independent parameter estimation—user’s manual, 4th edn. Watermark Numerical Computing

  • Faure G (1977) Principles of isotope geology. John Wiley and Sons, New York

    Google Scholar 

  • Federal Facility Agreement and Consent Order (1996, as amended) Agreed to by the State of Nevada, the U.S. Department of Energy, and the U.S. Department of Defense

  • Fisher R (1953) Dispersion on a sphere. Proc R Soc Lond Ser A 217:295–305

    Article  Google Scholar 

  • Franks SW, Beven KJ, Gash JHC (1999) Multi-objective conditioning of a simple SVAT model. Hydrol Earth Syst Sci 4:477–789

    Article  Google Scholar 

  • Guimerà J, Carrera J (2000) A comparison of hydrualic and transport parameters measured in low-permeability fractured media. J Contam Hydrol 41:261–281

    Article  Google Scholar 

  • Gustafson G, Fransson A (2005) The use of the Pareto distribution for fracture transmissivity assessment. Hydrogeol J. doi:10.1007/s10040-005-0440-y

  • Harbaugh AW, Banta ER, Hill MC, McDonald MG (2000) MODFLOW-2000, the U.S. Geological Survey Modular Ground-Water Model—user guide to modularization concepts and the ground-water flow process. U.S. Geological Survey Open–File Report 00-92

  • Hassan AE, Mohamed MM (2003) On using particle tracking methods to simulate transport in single-continuum and dual continua porous media. J Hydrol 275:242–260

    Article  CAS  Google Scholar 

  • Hendricks Franssen HJWM, Gómez-Hernández JJ (2002) 3D inverse modelling of groundwater flow at a fractured site using a stochastic continuum model with multiple statistical populations. Stoch Environ Res Risk Assess 16:155–174. doi:10.1007/s00477-002-0091-7

  • Hurvich CM, Tsai C-L (1989) Regression and time series model selection in small sample, Biometrtika 76:99–104

    Google Scholar 

  • Isherwood D, Harrar J, Raber E (1982) Characterization of Climax granite ground water, UCRL-53309. Lawrence Livermore National Laboratory, Livermore, CA

    Google Scholar 

  • Kashyap RL (1982) Optimal choice of AR and MA parts in autoregressive moving models. IEEE Trans Pattern Anal Mach Intell 4:99–104

    Article  Google Scholar 

  • Kemeny J, Post R (2003) Estimating three-dimensional rock discontinuity orientation from digitial images of fracture traces. Comput Geosci 29:65–77

    Article  Google Scholar 

  • LaBolle E, Fogg G, Thompson AFB (1996) Random-walk simulation of solute transport in heterogeneous porous media: local mass-conservation and implementation methods. Water Resour Res 32:583–593

    Article  Google Scholar 

  • LaBolle E, Quastel J, Fogg G, Gravner J (2000) Diffusion processes in composite porous media with their integration by numerical walks: generalizated stochastic differential equations with discontinuous coefficients. Water Resour Res 36:651–662

    Article  Google Scholar 

  • Liu HH, Bodvarsson GS, Pan L (2000) Determination of particle transfer in random walk methods for fractured porous media. Water Resour Res 36:707–713

    Article  Google Scholar 

  • Mardia KV, Jupp PE (2000) Directional statistics. Wiley, New York

    Google Scholar 

  • Munier R (2004) Statistical analysis of fracture data adopted for modeling discrete fracture networks—version 2. Rep. R 04-66. Swedish Nuclear Fuel and Waste Management, Co. (SKB). Stockholm, Sweden

  • Murray WA (1980) Permeability testing of fractures in Climax stock granite, NTS. In: Repository sealilng field testing workshop, Santa Fe, New Mexico, September 18–19

  • Murray WA (1981) Geohydrology of the Climax stock granite and surrounding rock formations, NTS, UCRL-53138. Lawrence Livermore National Laboratory, Livermore, CA

  • Neuman SP (2003) Maximum likelihood Bayesian averaging of uncertain model predictions. Stoch Environ Res Risk Assess 17:291–305. doi:10.1007/s00477-003-0151-7

    Article  Google Scholar 

  • Paillet FL (1988) Flow modeling and peremability estimation using borehole flow logs in heterogeneous fractured formations. Water Resour Res 34(5):997–1010

    Article  Google Scholar 

  • Pohlmann K, Pohll G, Chapman J, Hassan AE, Carroll R, Shirley C (2004) Modeling to support groundwater contaminant boundaries for the Shoal underground nuclear test. Desert Research Institute, Division of Hydrologic Sciences Publication No. 45184

  • Pohlmann K, Ye M, Reeves D, Zavarin M, Decker D, Chapman J (2007) Modeling of groundwater flow and radionuclide transport at the Climax mine sub-CAU, Nevada Test Site, DOE/NV/26383-06. Nevada Site Office, National Nuclear Security Administration, U.S. Department of Energy, Las Vegas, NV

  • Reeves DM, Benson DA, Meerschaert MM (2008a) Transport of conservative solutes in simulated fracture networks. 1. Synthethic data generation. Water Resour Res 44:W05401. doi:10.1029/2007WR006069

    Article  Google Scholar 

  • Reeves DM, Benson DA, Meerschaert MM, Scheffler H-P (2008b) Transport of conservative solutes in simulated fracture networks. 2. Ensemble solute transport and the correspondence to operator-stable limit distributions. Water Resour Res 44:W05410. doi:10.1029/2008WR006858

    Article  Google Scholar 

  • Reeves DM, Benson DA, Meerschaert MM (2008c) Influence of fracture statistics on advective transport and implications for geologic repositories. Water Resour Res 44:W08405. doi:10.1029/2007WR006179

    Article  Google Scholar 

  • Renshaw CE (1999) Connectivity of joint networks with power law length distributions. Water Resour Res 35(9):2661–2670

    Article  Google Scholar 

  • Russell CE, Minor T (2002) Reconnaissance estimates of recharge based on an elevation-dependent cloride mass-balance approach, DOE/NV11508-37. Nevada Site Office, National Nuclear Security Administration, U.S. Department of Energy, Las Vegas, NV

  • Schwarz G (1978) Estimating the dimension of a model. Annu Stat 6(2):461–464

    Article  Google Scholar 

  • Stigsson M, Outters N, Hermanson J (2001) Äspö Hard Rock Laboratory, Prototype repository hydraulic DFN model no. 2, IPR-01-39. Swedish Nuclear Fuel and Waste Management Co. (SKB), Stockholm, Sweden

  • U.S. Department of Energy (DOE) (2000a) United States Nuclear Tests, July 1945 through September 1992, DOE/NV-209 (Rev. 15). Nevada Operations Office

  • U.S. Department of Energy (DOE) (2000b) Investigation Plan for Corrective Action Unit 97: Yucca Flat/Climax Mine, Nevada Test Site, Nevada. DOE/NV-659. Nevada Operations Office

  • Vrugt JA, ter Braak CJF, Gupta HV, Robinson BA (2008) Equifinality of formal (DREAM) and informal (GLUE) Bayesian approaches in hydrologic modeling. Stoch Environ Res Risk Assess. doi:10.1007/s00477-008-0274-y

  • Wood ATA (1994) Simulation of the Von Mises distribution. Commun Stat-Sim 23(1):157–164

    Article  Google Scholar 

  • Ye M, Pohlmann KF, Chapman JB (2008) Expert elicitation of recharge model probabilities for the Death Valley regional flow system. J Hydrol 354:102–115. doi:10.1016/j.jhydrol.2008.03.001

    Article  Google Scholar 

  • Ye M, Pohlmann KF, Chapman JB, Pohll GM, Reeves DM (2010) A model-averaging method for assessing groundwater conceptual model uncertainty. Ground Water. doi:10.1111/j.1745-6584.2009.00633.x

  • Yow JL (1984) Geologic structure mapping database, Spend Fuel Test—Climax, Nevada Test Site, DE85006267

Download references

Acknowledgements

This research was supported by the U.S. Department of Energy, National Nuclear Security Administration Nevada Site Office under Contract DE-AC52-00NV13609 with the Desert Research Institute. Special thanks goes to the guest editor Dr. Yu-Feng Lin and Drs. Yonas Demissi, Abhishek Singh and Andrew F.B. Tompson for constructive comments that greatly improved the quality of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Donald M. Reeves.

Appendix: Fracture characterization

Appendix: Fracture characterization

Numerical modeling of fluid flow in fracture dominated subsurface flow regimes requires statistical analysis of fracture data for the determination of fracture properties, such as fracture sets and their mean orientation, length, spacing and distribution, density, and permeability of individual fractures or zones (e.g., Munier 2004; Reeves et al. 2008a). Fracture characterization at the Climax stock is based on the Spent Fuel Test—Climax (SFT-C) Geologic Structure Database (Yow 1984) that consists of data describing joints, faults and shear zones (sample population n = 2, 591) that were collected during fracture mapping efforts in tunnel drifts constructed for the Climax Spent Fuel Test.

1.1 Fracture set orientation

Analysis of fracture orientation statistics, according to standard spherical statistical tests (Mardia and Jupp 2000), reveals a total of seven fracture sets (Table 3 and Fig. 10). Of the seven fracture sets, Sets 1 through 6 fit a Fisher distribution (Fisher 1953):

$$ f(x)={\frac{\kappa \cdot{\sin}(x)\cdot{e}^{\kappa\cdot{\cos}(x)}}{e^{\kappa}-e^{-\kappa}}} $$
(7)

where the divergence, x (degrees), from a mean orientation vector is symmetrically distributed \((-{\frac{\pi}{2}}\leq x \leq{\frac{\pi}{2}})\) according to a constant dispersion parameter, κ. The Fisher distribution is a special case of the Von Mises distribution, and is similar to a normal distribution for spherical data (Mardia and Jupp 2000). The extent to which individual fractures cluster around a mean orientation is proportional to values of κ (i.e., higher values of κ describe higher degrees of clustering). Values of κ for natural rock fractures range between 10 and 300 (Kemeny and Post 2003). Stochastic simulation of Fisher random deviates is based on a method by Wood (1994). The occurrence of each fracture set is governed by the prior probabilities listed in Table 3. The distribution of fracture orientation for Set 7 is assumed uniform.

Fig. 10
figure 10

Lower-hemisphere equal area projection of the poles to the mean orientation vectors for Sets 1 through 6; the remaining 15% of fractures are randomly oriented

1.2 Fracture length

Fracture length is perhaps the most important parameter for discrete fracture network investigations. The distribution of fracture length has been found in theoretical studies to control both network connectivity (Renshaw 1999; de Dreuzy 2001) and the spreading rate of solutes at the leading plume edge (Reeves et al. 2008b, c). For the Climax stock, fracture length is recorded for approximately 95% of the fractures contained in the SFT-C data base. Fracture lengths range from 0.006 to 40 m within this data set. The upper range in the data is misleading as fracture length values are restricted to drift length and orientation, i.e., the longest fracture is parallel to one of the drifts and approximately two-thirds of the total drift length (67 m). According to a maximum likelihood estimation method (Aban et al. 2006), fracture length data fit a Pareto power-law distribution (α = 1.6) that is truncated for the largest values (denoted as “TPL”) (Fig. 11). Again, the truncation in fracture length is artificial, and is a result of the drift orientation and length relative to the occurrence and orientation of fractures.

Fig. 11
figure 11

Mandelbrot plot of all fracture lengths (l along with best fit Pareto power-law (PL), truncated Pareto power-law (TPL) and exponential (exp) models). Note the decay of the largest fracture lengths (i.e., distributional tail) follows a strong power-law trend (linear in log-log space) with an abrupt truncation. The truncation is artificial and results from the length and orientation of the drifts. All fracture lengths are in meters

To honor the power-law trend observed for the fracture length data set (linear trend on Fig. 11), an upper truncated Pareto model is used to randomly assign fracture lengths (Aban et al. 2006):

$$ P(L\,>\,l)={\frac{\gamma^{\alpha}(l^{-\alpha}-\nu^{-\alpha})}{1-({\frac{\gamma} {\nu}})^{\alpha}}} $$
(8)

where \(L_{(1)},L_{(2)},\ldots,L_{(n)}\) are fracture lengths in descending order and L (1) and L (n) represent the largest and smallest fracture lengths, γ and ν are lower and upper fracture length cutoff values, and α describes the tail of the distribution. Specific values used to describe fracture lengths at the Climax stock include: α = 1.6, γ = 30 m, and ν = 1,000 m. The lower cutoff of 30 m is equal to two times the edge of a cell in the continuum grid (15 m) and the upper cutoff of 1,000 m is equal to one-third of the FCM domain in the x-direction. The purpose of assigning fracture length according to (8) is to restrict fracture length to a finite upper bound. There is no evidence, with the exception of kilometer-scale faults that bound the Climax stock, of large faults that potentially span the entire length (∼5 km) of the stock. Thus, we deem a classical Pareto power-law (denoted as “PL” in Fig. 11) as an inappropriate choice for fracture length.

1.3 Fracture density

It is not possible to directly measure the three-dimensional fracture density of a rock mass. Instead, three-dimensional density of discrete fracture networks is estimated from density measures of lower dimensions, i.e., one-dimensional fracture frequency from boreholes or tunnel drifts or two-dimensional density from outcrops or fracture trace maps. Fracture frequency in the SFT-C database, based on fracture mapping along tunnel drifts, is relatively high and ranges between 2.0 to 5.5 fractures per meter. However, the high fracture spacing in the SFT-C database is misleading as the frequency of “flowing” or conductive fractures is not considered.

Field observations for fractured rock masses indicate that rock volumes are often intersected by only a few dominant fractures and only approximately 10% (or less) of the total fracture population contributes to flow (Dershowitz et al. 2000). This implies that 90% of fractures in the SFT-C database are not connected to the hydraulic backbone. Two specific studies at Climax provide some insight into the frequency of open fractures. In a tunnel drift for the SFT-C experiments, a series of five boreholes extending 9 to 12 meters below the tunnel drift yielded permeability values typical of unfractured granite cores (Ballou 1979). This indicates that the borehole array, which is on the scale of a grid cell, only intersects either solid rock or rock with “healed” fractures. “Healed” fractures (i.e., veins) refer to fractures containing mineral precipitates (e.g., calcite) and are, therefore, not open to flow. Several instances of healed fractures were documented in the SFT-C database by Yow (1984)—these fractures were excluded from the frequency analysis when recorded. In the permeability test conducted by Isherwood et al. (1982), only 2 out of 10 (20%) fractures in a densely fractured zone were open and had permeability values higher than the background matrix.

Given the high level of uncertainty in fracture density, this parameter is determined in the fracture continuum model during calibration (Table 1). Refer to Sect. 3.2 for additional discussion.

1.4 Hydraulic conductivity

Only a handful of field-scale hydraulic conductivity (K) measurements exist for the Climax stock (on the order of 10−7 to 10−10 m/s) (Murray 1980, 1981). These values describe the bulk hydraulic conductivity of the fractured stock, and most likely underestimate variability in fracture K since these estimates are based on hydraulic testing over large open borehole intervals where properties of multiple fractures are averaged. This narrow range (3 orders of magnitude) is inconsistent with other studies of highly characterized fractured granite rock masses where values of fracture K encompass 5–8 orders of magnitude (Paillet 1998; Guimerà and Carrera 2000; Andersson et al. 2002; Hendricks Franssen and Gómez-Hernández 2002; Gustafson and Fransson 2005).

Instead of relying on a handful of effective permeability measurements to parameterize a probability distribution for fracture K, the distribution of mechanical fracture apertures in the SFT-C database is analyzed. Recorded aperture values are lognormally distributed with a standard deviation of 1.05 (not shown) and this value is used to describe the variability in the fracture K distributions. This value is identical to the standard deviation of the transmissivity distribution used by Stigsson et al. (2001) at the Äspö Hard Rock Laboratory. Though rock fracture hydraulic conductivity is proportional to mechanical aperture, correlations between mechanical and equivalent hydraulic apertures are often unreliable (Bandis et al. 1985; Cook et al. 1990); therefore, mean values of K are not computed from the mechanical aperture distribution. Aperture data are used only to gauge the suitability of a lognormal distribution for fracture K and to provide an estimate of standard deviation.

To maintain a constant conceptual model for radionuclide flux estimates (refer to Sect. 3.2 for more detail), fracture K distributions were held constant at values of −7 and −8 (the original values of K are in units of meters per second prior to log transformation). Both of these mean fracture K values are within the narrow range defined by Murray (1981). The higher mean value of −7 m/s is assigned to fractures that belong to fracture set 6 (32% of the fractures) (Table 3). Fractures in this set experience the least amount of compressive stress normal to their fracture walls, suggesting that these fractures are potentially more permeable than fractures oriented at other directions to the regional stress field. The remaining fracture sets (68% of the fractures) are assigned K values according to the lower mean K of −8. A log10 K standard deviation value of 1.05 is applied to all fracture sets.

1.5 Fracture porosity

The computation of cell velocity from Darcy flux values calculated in the FCM flow realizations requires values of fracture and matrix cell porosity. Constant values of 0.006 are assigned to matrix cells (Murray 1981). Equivalent porosity values of fracture cells are based on tracer test results from a similar fractured granite rock mass where equivalent porosity was found to be lognormally distributed within a range of 0.027 and 0.054 (Pohlmann et al. 2004). Given the hydraulic conductivity distribution for fractures at Climax, an empirical power-law relationship:

$$ n=0.04(K_{\rm fracture})^{0.25} $$
(9)

is used to correlate fracture cell porosity n with fracture hydraulic conductivity K fracture, while maintaining both the range and distribution of the porosity values reported by Pohlmann (2004).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Reeves, D.M., Pohlmann, K.F., Pohll, G.M. et al. Incorporation of conceptual and parametric uncertainty into radionuclide flux estimates from a fractured granite rock mass. Stoch Environ Res Risk Assess 24, 899–915 (2010). https://doi.org/10.1007/s00477-010-0385-0

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00477-010-0385-0

Keywords

Navigation