WATER RESOURCES RESEARCH, VOL. 41, W02002, doi:10.1029/2004WR003682, 2005 Upscaling discrete fracture network simulations: An alternative to continuum transport models S. Painter Center for Nuclear Waste Regulatory Analyses, Southwest Research Institute, San Antonio, Texas, USA V. Cvetkovic Royal Institute of Technology, Stockholm, Sweden Abstract [1] Particle tracking through stochastically generated networks of discrete fractures provides an alternative to the conventional advection-dispersion description of transport in fractured rock. However, discrete fracture network simulations are computationally intensive and usually limited to small scales. An approach for direct upscaling of discrete fracture simulations is described. Trajectories for nonreacting tracer particles are first extracted from relatively small discrete fracture network simulations. Tracer-rock interaction is represented by also calculating a cumulative reactivity parameter along each trajectory. The residence time/reactivity information is then used in a Monte Carlo simulation to construct artificial particle trajectories of any length, thereby achieving the upscaling objective. In its simplest form the procedure has the form of a random walk evolving in a two-dimensional space. Tests using site-specific and generic networks show that it is necessary to modify the random walk to produce sequential correlation along the trajectories. We achieve this by using a discrete state Markov process to direct the random walk. The procedure is computationally efficient, easily implemented, and compares well with the network simulations. Received 25 September 2004; accepted 10 November 2004; published 2 February 2005. Keywords: fracture networks, retention, transport. Index Terms: 1832 Hydrology: Groundwater transport; 1869 Hydrology: Stochastic hydrology. Citation: Painter, S., and V. Cvetkovic (2005), Upscaling discrete fracture network simulations: An alternative to continuum transport models, Water Resour. Res., 41, W02002, doi:10.1029/2004WR003682. Copyright 2005 by the American Geophysical Union. Article Map Abstract 1. Introduction 2. Theory 2.1. Transport Model 2.2. Random Walk Representation for τ and B 2.3. Relationship to CTRW 3. Monte Carlo Simulation 3.1. Random Walk (RW) 3.2. Random Walk Directed by a Markov Process (MDRW) 4. Numerical Tests 4.1. Nonreacting Travel Time and Cumulative Reactivity 4.2. Transport 5. Example Simulations at the Geosphere Scale 6. Discussion 7. Conclusions 8. Acknowledgment References Figures Figure 1. Figure 2. Figure 3. Figure 4. Figure 5. Figure 6. Tables Table 1. Half-Life, Diffusion Coefficient, and Distribution Coefficients for Three Radionuclides Used as Examples in This Study Table 2. Peak of the Impulse Response γ* and Total Mass Released μ for Three Radionuclides, as Generated With the RW and MDRW Upscaling Methods Compared With Those From DFN Simulations Introduction [2] In hydrological applications involving low-permeability formations, flow and associated advective transport in the host rock are often not significant, and interconnected networks of rock fractures form the primary pathways for release of toxic chemical or radioactive wastes to the accessible environment. In particular, fractures provide the most plausible pathways to the biosphere for nuclear waste buried at depth, and therefore it is crucial to have accurate and practical tools for modeling radionuclide movement through fracture networks. [3] When applied to sparsely fractured rock, the conventional advective-dispersion equation (ADE) description of transport falls short of predicting many behaviors observed in field or numerical studies. For example, tracers are often transported over considerable distances in highly localized channels in contrast to the more uniform dispersion predicted by the ADE [National Research Council, 1996; Dverstorp et al., 1992; Neretnieks, 1993; Tsang and Neretnieks, 1998]. Dispersivities appear to vary with the scale of the observation, which is in direct conflict with the Fickian transport assumption underlying the ADE [e.g., National Research Council, 1996]. Numerical simulations also show that fracture networks display different effective porosities for transport in different directions [Endo et al., 1984] and highly non-Gaussian breakthrough curves [Andersson and Dverstorp, 1987; Schwartz and Smith, 1988; Berkowitz and Scher, 1997, 1998; Painter et al., 2002]. These and other related behaviors are not captured in the ADE model and may be due to the poorly connected nature of many natural fracture networks. In general, the quality of the representation provided by the ADE is better for highly and uniformly fractured rock as compared with sparsely fractured rock, and may be questionable for the sparsely fractured rock favored for geological disposal. [4] Discrete fracture network (DFN) models provide an alternative for situations where the ADE is inadequate, and have been used in a variety of studies in two dimensions [Long et al., 1982; Dershowitz, 1984; Endo et al., 1984; Robinson, 1984; Smith and Schwartz, 1984] and three dimensions [e.g., Shapiro and Andersson, 1985; Elsworth, 1986; Andersson and Dverstorp, 1987; Cacas et al., 1990; Long et al., 1992]. The DFN simulations avoid the volume averaging required for traditional equivalent continuum models and can generally represent a wider range of transport phenomena. DFN models play an important role in fundamental conceptual model evaluations, but site-specific applications are generally limited to near-field scales (50–100 m) due to the large computational demands. Many applications, particularly those involving geological disposal of high-level nuclear waste, involve spatial scales of a few hundred meters to a few kilometers in three dimensions. Thus there is a need for methods to bridge the gap between the spatial scale at which DFN simulations are tractable and the repository geosphere scale. [5] One approach for bridging this gap in scales is to perform DFN simulations for a subdomain of the field-scale domain to deduce parameters for a simpler and less computationally demanding model (such as a continuum model) for use at the field scale. The implicit assumption is that the fracture network model used in the DFN simulation is representative of the fracture network of the larger domain. This general approach has been referred to as the hybrid approach [National Research Council, 1996] and can also be thought of as an upscaling problem. For example, Long et al. [1982] used the results from a large number of DFN simulations to show that flow in some (but not all) DFNs could be described by equivalent hydraulic conductivity tensors. These permeability tensors are then available for use in a full field-scale simulation of flow using conventional continuum flow models. [6] Hybrid models have also been used for solute transport in fractured rock. Schwartz and Smith [1988] collected statistics on particle velocities in DFN simulations, fitted lognormal and gamma distributions to the velocity distributions, and then sampled these fitted distributions in a random walk Monte Carlo simulation. The approach eliminates the need to assign numerical values to a dispersion tensor, and accurately reproduced spreading patterns in two-dimensional DFNs. Berkowitz and Scher [1997, 1998] also collected statistics on particle velocities from DFNs and fitted a model distribution. Instead of using the fitted distribution in a Monte Carlo simulation, they used the continuous time-random walk (CTRW) formalism and obtained semianalytical results for plume evolution and important insights into anomalous (non-Fickian) transport. Although not explicitly addressing fractured rock, recent work on CTRW [e.g., Berkowitz et al., 2002; Cortis et al., 2004] add to its potential utility as a hybrid model for transport in fractured rock. Extending an approach originally proposed by Williams [1992, 1993], Benke and Painter [2003] used the results from DFN simulations to deduce parameters in a linear Boltzmann transport equation and then used the Boltzmann equation as a continuum model for transport at the field scale. This hybrid Boltzmann approach also successfully reproduced spreading patterns and breakthrough curves calculated from DFN simulations. [7] The hybrid models just discussed considered advective transport only and neglected retention processes that occur primarily in the porous matrix. While purely advective transport is a logical starting point for development of alternative models, mass exchanges between fractures and matrix combined with retention processes (diffusion, sorption) in the matrix are important for field applications and tend to dominate over advective transport at the time and spatial scales relevant for nuclear waste disposal. However, mass exchange between matrix and fractures is partly controlled by the local advective velocity [Cvetkovic et al., 1999]. Because of this strong coupling between hydrodynamics and mass retention, models for field-scale transport in fractured rock need to consider both the nonclassical phenomenology of advective transport in random fracture networks and mechanistic models for solute retention in the host rock. Dentz and Berkowitz [2003] recently extended the CTRW approach to account for retention process that can be described as a multitude of first-order trapping rates that do not vary spatially. Kosakowski et al. [2001] used breakthrough curves measured from a tracer test in a fractured till to fit CTRW waiting time distributions, an approach that implicitly and phenomenologically incorporates retention mechanisms. Using this approach, they successfully extrapolated (upscaled) from a travel distance of 2.1 m to a travel distance of 3.6 m. [8] In this paper, we describe a new approach for field-scale simulation of transport in fractured rock based on upscaling the results of particle tracking on discrete fracture network simulations. Specifically, we show how relevant information from nonreacting particle tracking in relatively small DFN simulations can be extracted and used in a Monte Carlo random walk at the geosphere scale. New aspects of this work are as follows. First, retention in the rock matrix is described within a stochastic Lagrangian formalism, an approach that honors the hydrodynamic controls on retention while accommodating a wide variety of retention models with random spatial variations in retention parameters. Second, two new Monte Carlo procedures for generating the controlling Lagrangian parameters at the field scale are described. Both make direct use of information extracted from DFN simulations without fitting a distributional model. Third, we show that sequential correlation (persistence) along the trajectory is necessary to accurately reproduce the breakthrough curves. When this persistence is taken into account, the upscaling method accurately reproduces the results of DFN simulations and provides a practical and easily implemented alternative to continuum transport models. Theory 2.1. Transport Model [9] Consider a hypothetical solute source located in a fractured rock volume. Steady state groundwater flow driven by a regional hydraulic gradient carries solutes toward a monitoring boundary located downstream. Diffusive mass exchange with the host rock and sorption in the host rock delay the downstream movement. We are interested in the time-dependent mass flux at this monitoring boundary. We adopt a Lagrangian perspective and consider multiple meandering transport pathways (trajectories) that connect each source location to the monitoring boundary. In general, the flow velocities, fracture apertures, and possibly retention properties fluctuate along each trajectory. [10] Recent theory [Cvetkovic et al., 1998, 1999; Cvetkovic and Haggerty, 2002; Cvetkovic et al., 2004] has shown that transport with a wide range of retention models can be represented in a generic compact form in the Laplace domain. Specifically, it has been shown that the fundamental solution (the impulse response function) for a Lagrangian representation of transport with retention in fractured rock can be represented as where λ is the radionuclide decay constant, τ is the water residence time along the trajectory, is the Laplace transform of the memory function, and B is a cumulative reactivity parameter that integrates retention properties along the trajectory. Equation (1) is rather general and includes the widely used retention models with spatial variability in retention parameters as special cases by appropriate selection of the memory function and cumulative reactivity. For example, limited diffusion, unlimited diffusion, and multirate linear transfer models are included as special cases. [11] The groundwater residence time can be written in integral form as where x is the coordinate in the direction of mean flow, Vx is the velocity component in that direction, and y = η(x), z = ζ(x) defines the particle trajectory. The cumulative reactivity B is dependent on the choice for retention model, is highly correlated to τ, and generally integrates retention properties along the trajectories. [12] The function γl(t; τ, B) is the conditional impulse response, the time-dependent discharge at a monitoring boundary located at distance l from a dirac-δ input for given values of τ and B. In the situation of no decay, it can be thought of as an arrival time distribution at the monitoring boundary. In general, it represents the discharge along a single trajectory characterized by τ and B, and is not observable. In applications, unconditional quantities are needed and can be obtained by averaging with respect to f(τ, B), the joint probability density for τ and B. For example, the unconditional impulse response is Under fully ergodic conditions γl(t) represents the breakthrough curve at the monitoring boundary. Under nonergodic conditions, the breakthrough curve will vary from realization to realization, and γl(t) represents the expected value. Other convenient measures of transport can be defined [Cvetkovic et al., 2004] for the nonergodic or ergodic situations. In either case, the density f(τ, B) is needed to calculate meaningful quantities. 2.2. Random Walk Representation for τ and B [13] The two Lagrangian quantities τ and B characterize transport and retention along a given trajectory. Specifically, given a value for τ and B, we can calculate the conditional timedependent discharge from equation (1). Given a distribution for τ and B, we can calculate the unconditional discharge and measures of uncertainty in the predicted discharge. Thus the challenge is to obtain a model or algorithm for the joint distribution of τ and B. [14] Consider a discretization of the trajectory into a number of jumps, with each jump corresponding to transit through an individual fracture. Let Δ1, Δ2, denote identically distributed random variables that model the change in the x coordinate for the jumps. Similarly, let Δτ1, Δτ2, and ΔB1, ΔB2, denote random variables modeling the change in τ and B, respectively. In general, Δi, Δτi and ΔBi are correlated. [15] After n jumps, the x position is L(n) = Δi, and the τ and B values are τ(n) = Δτi and B(n) = ΔBi. The random walk is to be executed until the particle hits a monitoring boundary. The number of jumps required to hit the monitoring boundary is a random variable Nl = min {n: L(n) ≥ l} and the corresponding τ and B are The stochastic process {τ(Nl), B(Nl)}l>0 governs the desired distribution of τ and B for a given l. In words, the process is a running sum of correlated random variables that is stopped when one of these (L) reaches a predefined cutoff value l. [16] The stochastic process defined by (3) with an additional assumption of independent jumps has been studied by Meerschaert et al. [2002], who develop limit theories for the process in the situation of power law tails in the distribution of jumps [see also Becker-Kern et al., 2004]. If the number of terms in the sums in equation (3) was fixed, then (3) would be a simple random walk. However, Nl is a random variable, and the stochastic process is thus a random walk subordinated by the random process governing Nl [Meerschaert et al., 2002]. Note that x position in (3) is assuming the role played by time in the process studied by Meerschaert et al. [2002]. In particular, the particle executes the random walk until it reaches a specified monitoring boundary, not a specified time level. In other words, our walk is developing in the two-dimensional τ, B space with the x position triggering the termination of the walk. The position in the direction of mean flow, x in (3) is not required to be strictly increasing as a time would be. [17] At this point, we have made no statement about the distribution of the jumps. An obvious choice would be to assume independent jumps. However, careful review of DFN simulation results clearly show correlation between successive segments along the trajectory [Painter et al., 2002; Benke and Painter, 2003]. A particle that is in a high-velocity segment is more likely to find itself in a high-velocity segment in the subsequent segment due to conservation of flux at the fracture intersections. We use a Markov-type approximation to reproduce the sequential correlation, wherein each jump is correlated to its previous jump. There are many ways that this could be implemented in practice. A particularly convenient method to generate this sequential correlation is to give each particle an internal state that changes from segment to segment according to a discrete state Markov process. The selection of the internal state will be discussed in Section 3; for now we denote the set of possible states as S. The internal state of the particle is then used to alter the distribution of jumps. Specifically, let S1, S2, S3, denote the sequence of internal states, which we assume to be a discrete state Markov process governed by the transition matrix A. Usual rules for discrete state Markov processes apply: Amn is probability for transitioning from state m to state n, Amn = Pn where Pn is probability for state n, Amn = 1. The distribution for each jump is assumed to be dependent on the internal state only and has conditional density denoted by f(Δ1, Δτ1, Δβ1|S1). The unconditional distribution for the first jump is then The joint distribution for the first and second jump is given by and the joint density for all jumps along the trajectory can be written as Such a model is an example of a subordination process with the Markov process directing the random walk and can also be thought of as a random walk with an internal state. We refer to the process as a Markov-directed random walk (MDRW). [18] If we assume only one state, then the jumps become independent and identically distributed: This is referred to as the random walk (RW) in the following to underscore the independent nature of the jumps. The MDRW model equation (3) with (6) is the primary focus in this paper; the RW model (3) with (7) is also considered for comparison purposes. 2.3. Relationship to CTRW [19] In the special case of advection only (no retention), ΔBi = 0, the time domain representation of equation (1) is γl(t; τ) = exp[−λτ]δ(t − τ), and the unconditional response becomes γl(t) = fτ(t)exp[−λt] where fτ( ) is the travel time distribution for a nonreacting, nondecaying tracer. In addition, (6) becomes Making the further assumption of independent jumps, Equation (3) with (9) is equivalent to the CTRW that Berkowitz and Scher [1997, 1998] applied to advective transport in fractured rock, with f(Δ, Δτ) corresponding to the CTRW waiting time density (denoted by ψ(t, l) in the notation of Berkowitz and Scher [1997, 1998]). [20] It should be noted, however, that the CTRW framework contains additional flexibility that Berkowitz and Scher [1997, 1998] did not use in their study of advective transport in fracture networks. Retention in the CTRW framework is included in a very different way than in our model. Specifically, retention is incorporated into the definition of the waiting density in CTRW, as opposed to our Lagrangian approach, which effectively decomposes the problem into one of determining nonreacting trajectories and then using equation (1) and (2) to incorporate the effects of retention. Kosakowski et al. [2001] determined the waiting time distribution by fitting breakthrough curves, an approach that phenomenologically incorporates retention into the waiting time distribution. Although not specifically addressing fractured rock, Dentz and Berkowitz [2003] incorporate mechanistic retention models in the situation of no spatial variability in the mobile-immobile exchange rates. In fractured rock, the rates of exchange between fractures and matrix are linked to the local velocity [Cvetkovic et al., 1999]; hence a spatially variable velocity necessarily implies spatially variable exchange rates. V. Cvetkovic and S. Painter (Tracer transport with transition and exchange disorder in random media, submitted to Physical Review E, 2004) recently showed how general mechanistic models with spatially variable exchange rates can be incorporated into the CTRW waiting time density, thus achieving the same as equation (7). Extensions of CTRW that incorporate correlation between one jump and the next jump in the sequence similar to equation (8) also exist [see, e.g., Hughes 1995, and references therein] but have not been applied to transport in the subsurface. Berkowitz and Scher [1998] proposed to artificially extend the particle displacement to mimic correlation along the trajectory, but they did not test this approximation nor did they specify how to estimate the required extension. In principle, these various refinements and extensions of the CTRW formalism could be combined with retention models appropriate for fractured rock to represent the same processes as our model, but this has not yet been demonstrated. We choose instead a simple Monte Carlo approach based on equation (3) with (6) or (7). Monte Carlo Simulation [21] For calculations, two steps remain: (1) determining the joint distribution of individual fracture values Δ, Δτ, ΔB appropriate for a given site, and (2) determining the global distribution of {τ, B} given the individual fracture distribution. [22] At present, DFN simulation is the only viable method for determining the joint distribution of individual fracture values. In this approach, realizations of the DFN at the near-field scale are generated using standard fracture network modeling tools, taking into account as much sitespecific information as possible to constrain the network properties. Site-appropriate hydraulic boundary conditions are applied and the resulting flow solved. Particles are then tracked through the DFN velocity field. For each fracture segment traversed by each particle, the {Δ, Δτ, ΔB} triplet is recorded. The set of these triplets represents Monte Carlo samples from the joint distribution. [23] The DFN simulations need to be large enough to minimize boundary effects that may cause the jump statistics near the particle source to be different from those of the bulk DFN. In the site-specific and generic simulations described later in this paper, such boundary effects were observed but disappear quickly with distance from the source due to randomizing effects of fracture intersections. In practice, it is easy to verify that the domain in large enough by dividing the domain into upstream and downstream halves and verifying that the jump statistics are not too different in the two halves. [24] Note that an assumption of ergodicity has not been required. We simply require access to the ensemble distribution of jumps. If the DFN simulations are large enough to ensure ergodicity in the particle trajectories, then a single DFN simulation with multiple trajectories could be used to generate the ensemble statistics. In the opposite extreme of fully nonergodic conditions, a single trajectory from each of multiple realizations would be used. In practice, the situation is likely to be intermediate between these two extremes, but experience suggests the situation to be closer to the ergodic situation due to strong randomizing effects of fracture intersections. In the numerical examples described later in this paper, multiple trajectories in a relatively small number of realizations are used. [25] The results of the DFN simulation can then be used to upscale the individual fracture values to obtain global {τ, B}distributions. The procedure varies depending on what stochastic process is assumed for the {τ, B}; that is, it depends on whether equation (6) or equation (7) is used. 3.1. Random Walk (RW) [26] If the individual jumps forming the running sums in equation (3) are assumed to independent (i.e., governed by equation (7)) then the stochastic process {τ(Nl), B(Nl)}l>0 is similar to the process studied by Meerschaert et al. [2002]. Our Monte Carlo procedure to simulate this process is to form a particle trajectory by repeatedly selecting values at random from the library of {Δ, Δτ, ΔB} triplets and forming the running sums in (3). A trajectory is terminated once the x value exceeds l, indicating that the monitoring surface of interest has been reached. The {τ, β} at that point becomes one Monte Carlo sample from the global distribution. Because this procedure is very fast, it can be executed for large l, and is not limited by the same computational constraints as DFN simulations. Thus it can be used to upscale the results of DFN simulations, provided the fracture network model used in the DFN simulations is representative of conditions throughout the larger region of interest. 3.2. Random Walk Directed by a Markov Process (MDRW) [27] To account for persistence along the trajectory, the Markov directed random walk defined by equations (3) and (6) can be used. The Monte Carlo procedure for simulating the process is similar to that described in section 3.1 except that we give each particle an internal state, and then use this state to control the selection of the next {Δ, Δτ, ΔB} triplet in the sequence. The selection of the internal state is discussed in the following paragraph. We record, in addition to the {Δ, Δτ, ΔB} triplets, the internal state for each segment of each particle trajectory in the DFN simulation. When executing the Monte Carlo simulation, the state of the current segment is used to constrain the selection of the next {Δ, Δτ, ΔB} triplet. Specifically, if the current segment has state designated K, then only those segments that have K as the preceding state are eligible for selection as the next segment. Using this algorithm, the sequence of internal states represent a discrete state Markov process governed by the transition matrix A (see section 2.2), which directs the random walk simulation of the particle trajectory in the τ, B space. With this sampling method, it is not necessary to explicitly construct the A matrix from the DFN simulations, although it is easily accomplished [Benke and Painter, 2003]. [28] The definition of internal states requires some further discussion. Given that the goal is to create some sequential correlation in the sequence of {Δτ, ΔB} selected, a logical choice would be to discretize the{Δτ, ΔB} space, and use this discretization to define the internal state. This would have the desired effect of correlating each {Δτ, ΔB} pair with the preceding pair in a classic Markov-type approximation, provided such correlation is manifest in the DFN trajectories. This general approach has been used successfully in a fully discretized sense by Benke and Painter [2003]. A closely related but simpler approach is to use the particle speed to define the internal state. Because {Δτ, ΔB} are both closely related to local speed, this choice has a similar effect of building in correlation along the trajectory. We take this latter approach. Specifically, we discretize the particle speed into a small number of classes and use these speed classes as the internal particle state. Unless otherwise noted, we use 20 classes and partition the range of observed speeds so that each class has the same number of observations. Numerical Tests [29] Site specific and generic three-dimensional DFN simulations were used to test the upscaling methods. The site-specific DFN and particle tracking simulations [Outters, 2003] were designed to mimic the fracture network near the Äspö hard Rock Laboratory, Sweden and used the FracMan [Dershowitz et al., 1998] software with the MAFIC module [Miller et al., 1994] for transport. The computational domain was a 100 m × 100 m × 100 m cube. Each of the 20 realizations contained about 20000 disk shaped fractures tessellated into triangular finite elements. In each of the generated realizations, generic boundary conditions were prescribed so as to obtain a globally unidirectional flow. Once the flow field was computed, a large number of inert particles were released from a 50 m × 50 m square on the upstream boundary and traced though the network, assuming perfect mixing at each intersection. [30] The generic simulations [Painter et al., 2002] were also computed in three dimensions, but used a simplification that approximates flow between any two fracture intersections as onedimensional pipe flow instead of using a finite element mesh on each fracture plane. Similar approximations have been used previously [e.g., Outters and Shuttle, 2000]. For example, it is available as one option in the MAFIC software [Miller et al., 1994]. Despite the simplified representation of flow on each fracture plane, this approximate model does represent some threedimensional effects. Specifically, it represents the coordination number (average number of neighbors for each node) correctly for three-dimensional systems, and is thus considered to be adequate for the purposes of testing the upscaling algorithm. Nevertheless, in applications full finite element discretization of the fracture plane is to be preferred. [31] Two sets of generic simulations were performed, one using a 30 m × 30 m × 30 m domain and one using a 90 m × 30 m × 30 m domain. For each set, 100 realizations were created. The fracture network model was identical in the two sets and used disk-shaped fractures with an isotropic model for fracture orientation. Fracture radii follow a lognormal distribution with log variance of 0.25 and geometric mean of 1 m. Fracture apertures were also selected from a lognormal distribution (log variance of 0.25 and geometric mean of 100 micron). The density of fracture centers was 0.0185 m−3, corresponding to 500 fractures for the smaller size and 1500 for the larger simulations. [32] In order to proceed with the numerical tests, we need to specify a retention model. The unlimited diffusion model [e.g., Neretnieks, 1980] is the most widely used model for transport in sparsely fractured rock and is the logical first model to consider for the Äspö site. In this case, the memory function is = s−1/2 and the impulse response function in the time domain is where H is the Heaviside function. The cumulative reactivity for this model can be written as B = ( DR)1/2 where , D, R, b are the matrix porosity, effective diffusion coefficient, retardation factor, and fracture half aperture, respectively. For simplicity, we assume , D, R are spatially constant and known. In this case, variability in B arises from variability in aperture and velocity along the pathway, B = ( DR)1/2 β = ( DR)1/2 ∑ βi, and the new variable β = captures this variability. This assumption of constant retention properties is reasonable given that little information on retention property variability is available. However, we emphasize that spatial variability in the retention properties can easily be accommodated in the upscaling procedure, if adequate site-specific information is available. 4.1. Nonreacting Travel Time and Cumulative Reactivity Figure Figure 1 1. Cumulative distribution (CD) and complementary cumulative distribution (CCD) of global τ and β from discrete fracture network simulations compared with upscaling results based on a random walk (RW) and Markovdirected random walk (MDRW) simulation. [33] The first step is to compare the distributions of τ and β from the DFN simulations with those obtained from Monte Carlo upscaling simulations. Marginal distributions of τ and β for the Äspö reference case are compared in Figure 1. The data points represent the results of the DFN simulation, the dashed lines are from the random walk (RW) upscaling procedure, and the solid lines are the result of the Markov-directed random walk (MDRW). In both upscaling results, we took single-fracture distributions generated from FracMan/MAFIC and attempted to reproduce the global τ and β distributions. The RW upscaling method generally reproduces the distributions of τ and β in the right tail and in the bulk of the distribution, but generally predicts slightly higher values of τ and β in the left tail. The MDRW produces a better fit to the left tail because it better captures the weak correlation in velocity along the particle trajectories, which reduces overall travel time and cumulative reactivity in the leading edge of the distribution. Figure 2. Cross plots of nonreacting travel time τ and cumulative reactivity parameter β (left) from discrete fracture network simulations and (right) upscaled from the single-fracture statistics by the Markov-directed random walk algorithm. The high correlation between τ and β is reproduced by the upscaling method. [34] Cross plots of τ and β obtained from the MDRW and the DFN simulations are compared in Figure 2. The similarity between the two cross plots demonstrates that the upscaling method reproduces the strong correlation between τ and β. Figure 3. Detail of cumulative distribution for global β from discrete fracture network simulations compared with upscaling results based on a random walk (RW) and random walk directed by 2-state, 5-state, and 20-state Markov processes. The mismatch between the Markov-directed random walk (MDRW) and the DFN simulations improves as the number of states is increased, but most of the improvement occurs when going from one state (the RW situation) to two states in the directing process. [35] The MDRW results used in the comparisons in Figures 1 and 2 were obtained by using the local speed to define the particle's internal state as described in Section 3. The particle speed was binned into 20 speed classes in this case. An interesting issue to address is how the results vary as the number of speed classes is varied from 1 (the RW case) to 20. This sensitivity is shown in Figure 3. In going from one bin to two bins, there is a relatively large decrease in the mismatch between the DFN and the upscaling results. Comparatively smaller improvements are noted when going from two to five bins and from 5 to 20 bins. Even with 20 bins, there are still some differences between the DFN and the upscaling results. One possible cause of this mismatch is longer-range persistence in the sequence that is not captured with the Markov-type approximation. Another possible cause is weak nonstationarity, especially near the source region. Whatever the cause, the upscaled results appear to provide a reasonable approximation. This adequacy of the fit is explored in the next subsection. [36] The upscaling algorithms were also tested on an Äspö DFN simulation that had fracture density reduced by a factor of two compared with the reference case. Results (not shown) were similar to the reference case. The only significant differences between the upscaling simulations and the DFN simulations were in the left tail of the distribution, where the MDRW performs better than the RW. Interestingly, both the RW and the MDRW perform better in the lower density case as compared with the reference case. One possible explanation for this better performance is that there is some residual longer-range correlation along the trajectory that is not capture by the MDRW, and that this longer-range correlation is stronger in the denser networks. This issue does, however, require further study. [37] In the site-specific simulations just described, an assumption of perfect mixing at the fracture intersections was used. Other assumptions for particle redistribution at intersections are possible (i.e., streamline routing), and the transport consequences of the mixing assumption have been studied [e.g., Küpper et al., 1995; Park et al., 2001]. For the Äspö site-specific simulations, the perfect mixing and streamline routing assumptions produce nearly identical {τ, β} distributions [Cvetkovic et al., 2004]. The choice of mixing assumption is not expected to affect the performance of the upscaling method, but this needs to be verified using DFN particle tracking simulations with streamline routing. Figure 4. Cumulative distribution (CD) and complementary cumulative distribution (CCD) of global τ and β from generic discrete fracture network simulations compared with upscaling results based on a random walk (RW) and Markov-directed random walk (MDRW) simulation. In this example, statistics extacted from 30 × 30 × 30 m DFN simulations were used in the two upscaling algorithms to predict the distributions at the 90 × 30 × 30 m scale. [38] In the final test of the ability to upscale the τ and β distributions, generic DFN simulations are considered. In this situation, we start with DFN simulations at the 30 × 30 × 30 m scale and attempt to use single-fracture distributions derived from these to reproduce larger simulations (90 × 30 × 30 m). Results are shown in Figure 4. As with the site-specific simulations, the RW produces significant errors in the left tail of the distribution. In the generic simulations, the RW also has significant errors in the right tail of the distribution. The MDRW reproduces the DFN simulations well across the entire distribution, with the only significant errors occurring in the extreme right tail of the distribution which is highly uncertain because of the very few samples there. 4.2. Transport [39] Figures 1, 2, 3, and 4 suggest that the MDRW upscaling procedure provides reasonably accurate approximations to the τ and β distributions. To assess the adequacy of the approximations in a more quantitative way, we need go beyond the travel time/cumulative reactivity distributions and consider transport. To make this assessment, we focus on the Äspö reference case and consider the unconditional response equation (2). Three radionuclides were considered with properties as shown in Table 1. Figure 5. Unconditional discharge (breakthrough curves) for three radionuclides as calculated from DFN simulations and the two upscaling methods. The travel distance is 100 m in this example. Recall that γ(t) is the breakthrough corresponding to a dirac-δ input of unit strength and is thus dimensionless. [40] The normalized unconditional response γ(t) (breakthrough) as calculated from the DFN, RW and MDRW representations of the τ and β distributions are shown in Figure 5 for the three radionuclides. The RW representation underestimates the breakthrough curve by a significant amount for all three radionuclides (as much as a factor of 10 for Cs). In all cases the MDRW representation performs much better, underestimating γ(t) by about 25% at the worst point and by much smaller amounts for most of the curves. This error is likely to be acceptable for applications. [41] Some more compact measures of transport can also be compared. Specifically, the mass fraction released μ, which is obtained by integrating equation (1) over all times, and the peak value for γ(t), denoted γ* are listed in Table 2. Both upscaling methods represent μ better than γ*, but even for γ* the MDRW values are within 25% of the DFN values. Example Simulations at the Geosphere Scale Figure after 6. Normalized unconditional discharge for two radionuclides upscaling to the 500 m scale. DFN simulations at this scale are not feasible because of computational limitations. Figure 2 [42] The computational requirements for the MDRW algorithm are smaller than those of a full DFN simulation by many factors of ten, and once a small-scale DFN simulation has been completed the MDRW can be used to extrapolate to larger scales. An example is shown in Figure 6. In constructing Figure 6 the MDRW algorithm was executed for l = 500 m. The resulting τ, β joint distribution was then used to construct γ(t) for 126Sn and 129I. The γ(t) for 135Cs was very small (10−18) and is not plotted. Recall that γ(t) is the breakthrough corresponding to a dirac-δ input of unit strength, and is thus dimensionless. [43] The results in Figure 6 demonstrate that it is feasible to use direct upscaling of DFN simulations to estimate geosphere-scale transport. Note that DFN simulations at the geosphere scale are not feasible due to computational limitations. The 100 m × 100 m × 100 m DFN simulations required a few days to complete. Relevant scales for the geosphere are 500–1000 m and larger. DFN simulations at the 1000 m scale would increase the memory requirements by a factor of 103, and the computational requirements would increase by a much larger factor because the simulation time in such simulations increases nonlinearly with the number of fractures. Discussion [44] Our approach for upscaling the τ.B distribution to field scales considers that particles hop from fracture intersection to fracture intersection. This conceptualization is intuitive and has been used in other studies of transport in fractured rock. [45] One fundamental difference between this work and previous works that used a similar conceptualization is in how retention is incorporated. Specifically, we use the random walk to determine the distribution for the auxiliary variables τ.B. This determination is but one step in the calculation of breakthrough curves; the joint distribution f(τ, B) must then be combined with a mechanistic model for the retention mechanisms (see (2)). This approach allows mechanistic understanding of retention processes to be incorporated in a direct way. Once the τ.B distribution is calculated, multiple retention models can then be evaluated. This latter feature is particularly convenient for nuclear waste repository studies, which often require that alternative models be evaluated. [46] Another new aspect of this work is that we use individual segment properties directly from a DFN simulation in a Monte Carlo simulation, without fitting a model distribution. Schwartz and Smith [1988] also use a Monte Carlo simulation, but they first fit a model distribution and do not address retention. The advantage of our Monte Carlo option is that it is extremely simple to implement. Only a few lines of code are required, and the simulation executes very quickly. Of course, Monte Carlo approaches do not yield the same level of insight that can be obtained through analytical methods, and we regard the approach as a practical tool that can be used to estimate transport at a given site. For more generic insights into the transport process, methods like CTRW and its extensions are available. [47] The Monte Carlo method is also able to include correlation between successive segments in the trajectory through the fractured rock mass. This correlation is clear from our generic and site-specific DFN simulations. The results in Figure 5 demonstrate that large errors are introduced if this sequential correlation is neglected. The particular variants of CTRW that have been used to date to study transport in fractured rock are based on an assumption that the segments along the trajectory are mutually independent; correlation between successive segments in the trajectory is not included. Variants of CTRW that do include this sequential correlation have been developed in the physics literature [e.g., Hughes, 1995], but application of such methods to transport in fractured rock has yet to be demonstrated. Conclusions [48] In conclusion, direct upscaling of DFN simulations provides an alternative to site-scale continuum transport models. The suggested procedure is to first perform small-scale DFN simulations utilizing site-specific information on the fracture network, and then use the results collected from these DFN simulations in a Monte Carlo calculation to obtain transport results at the field scale. The approach avoids volume averaging and other assumptions inherent in the continuum approach. It preserves the highly non-Gaussian velocity statistics and the spatial correlation in velocity that are observed in DFN simulations. It also allows mechanistic models for retention processes to be incorporated directly, including the effects of spatial variability in retention properties. We demonstrated that these processes can be modeled in combination and at the geosphere scale with relatively modest computational effort. [49] Although not specifically addressed in the numerical tests presented here, variations in fracture aperture within each fracture can also be accommodated. Specifically, if internal variability in fracture apertures is included in the small-scale DFN simulations, then the effect of this variability will be embedded in the single-fracture statistics of travel time and cumulative reactivity, and will thus be carried forward into the upscaled results. DFN simulation with internal aperture variability has been demonstrated previously [e.g., Nordqvist et al., 1992, 1996], and is now available as an option in commercial software. For τ, internal variability is anticipated to be less important than fracture-to-fracture variability. However, internal variability may result in channeling and alter the cumulative retention parameter B. Additional study is required to assess the magnitude of the effect. [50] One key finding of this study is that correlation between successive segments along the particle trajectory is an important control on the breakthrough curves. Random walk models for transport in fractured rock need to incorporate this sequential correlation. The Monte Carlo algorithm described here achieves this in a simple and direct way. [51] Two straightforward modifications of the upscaling approach may be needed for applications. The method is based on the assumption that the jump statistics from the DFN simulations are representative of the jump statistics for the larger region of interest. This would not be true if the statistical properties of the networks vary significantly over the larger region of interest. Such nonstationarity is not uncommon and can be addressed in applications by simply dividing the larger region of interest into subregions with approximately constant network properties in each. For anisotropic networks, the jump statistics may also depend on the direction of macroscopic gradient relative to the principal directions of the network. If the direction of the macroscopic gradient varies significantly over the larger region of interest, similar subdivision may also be needed. This would, of course, require a separate set of DFN simulations to obtain the jump statistics in each modeled subregion. Acknowledgment [52] The authors thank Jan-Olof Selroos, the Swedish Nuclear Fuel and Waste Management Company (SKB), and the Southwest Research Institute Advisory Committee for Research for supporting this research. WATER RESOURCES RESEARCH, VOL. 41, W02002, doi:10.1029/2004WR003682, 2005 Figures Figure 1. Cumulative distribution (CD) and complementary cumulative distribution (CCD) of global τ and β from discrete fracture network simulations compared with upscaling results based on a random walk (RW) and Markov-directed random walk (MDRW) simulation. Figure 2. Cross plots of nonreacting travel time τ and cumulative reactivity parameter β (left) from discrete fracture network simulations and (right) upscaled from the single-fracture statistics by the Markov-directed random walk algorithm. The high correlation between τ and β is reproduced by the upscaling method. Figure 3. Detail of cumulative distribution for global β from discrete fracture network simulations compared with upscaling results based on a random walk (RW) and random walk directed by 2-state, 5-state, and 20-state Markov processes. The mismatch between the Markovdirected random walk (MDRW) and the DFN simulations improves as the number of states is increased, but most of the improvement occurs when going from one state (the RW situation) to two states in the directing process. Figure 4. Cumulative distribution (CD) and complementary cumulative distribution (CCD) of global τ and β from generic discrete fracture network simulations compared with upscaling results based on a random walk (RW) and Markov-directed random walk (MDRW) simulation. In this example, statistics extacted from 30 × 30 × 30 m DFN simulations were used in the two upscaling algorithms to predict the distributions at the 90 × 30 × 30 m scale. Figure 5. Unconditional discharge (breakthrough curves) for three radionuclides as calculated from DFN simulations and the two upscaling methods. The travel distance is 100 m in this example. Recall that γ(t) is the breakthrough corresponding to a dirac-δ input of unit strength and is thus dimensionless. Figure 6. Normalized unconditional discharge for two radionuclides after upscaling to the 500 m scale. DFN simulations at this scale are not feasible because of computational limitations. Citation: Painter, S., and V. Cvetkovic (2005), Upscaling discrete fracture network simulations: An alternative to continuum transport models, Water Resour. Res., 41, W02002, doi:10.1029/2004WR003682. Copyright 2005 by the American Geophysical Union. WATER RESOURCES RESEARCH, VOL. 41, W02002, doi:10.1029/2004WR003682, 2005 Tables Table 1. Half-Life, Diffusion Coefficient, and Distribution Coefficients for Three Radionuclides Used as Examples in This Studya Tracer t1/2, years D, m2/yr Kd, m3/kg 126 Sn 1.0e5 1.3e-6 1.0e-3 129 I 1.6e7 3.9e-6 0 135 Cs 2.3e6 1.3e-6 5.0e-2 a Read 1.0e5 as 1.0 × 105. Table 2. Peak of the Impulse Response γ* and Total Mass Released μ for Three Radionuclides, as Generated With the RW and MDRW Upscaling Methods Compared With Those From DFN Simulationsa γ* Tracer RW μ MDRW DFN RW MDRW DFN 126 Sn 7.1e-7 1.4e-6 2.0e-6 0.086 0.12 0.14 129 I 4.5e-4 7.4e-4 9.2e-4 0.97 0.97 135 Cs 2.6e-9 7.4e-9 1.1e-8 0.013 0.029 a 0.97 0.037 Scale is 100 m. Citation: Painter, S., and V. Cvetkovic (2005), Upscaling discrete fracture network simulations: An alternative to continuum transport models, Water Resour. Res., 41, W02002, doi:10.1029/2004WR003682. Copyright 2005 by the American Geophysical Union. Figure 3