Effects of large-scale heterogeneity and temporally varying hydrologic processes on estimating immobile pore space: A mesoscale-laboratory experimental and numerical modeling investigation

https://doi.org/10.1016/j.jconhyd.2021.103811Get rights and content

Highlights

  • Impermeable barriers and density dependence increase estimated immobile porosity.

  • Decreasing dispersivity leads to increased estimated immobile porosity.

  • Immobile porosity estimates were insensitive to changes in system flow rates.

Abstract

The advection-dispersion equation (ADE) often fails to predict solute transport, in part due to incomplete mixing in the subsurface, which the development of non-local models has attempted to deal with. One such model is dual-domain mass transfer (DDMT); one parameter that exists within this model type is called immobile porosity. Here, we explore the complexity of estimating immobile porosity under varying flow rates and density dependencies in a large-scale heterogeneous system. Immobile porosity is estimated experimentally and using numerical models in 3-D flow systems, and is defined by domains of comparatively low advective velocity instead of truly immobile regions at the pore scale. Tracer experiments were conducted in a mesoscale 3-D tank system with embedded large impermeable zones and the generated data were analyzed using a numerical model. The impermeable zones were used to explore how large-scale structure and heterogeneity affect parameter estimation of immobile porosity, assuming a dual-porosity model, and resultant characterization of the aquifer system. Spatially and temporally co-located fluid electrical conductivity (σf) and bulk apparent electrical conductivity (σb)—using geophysical methods—were measured to estimate immobile porosity, and numerical modeling (i.e., SEAWAT and R3t) was conducted to explore controls of the immobile zones on the experimentally observed flow and transport. Results showed that density-dependent flow increased the hysteresis between measured fluid and bulk electrical conductivity, resulting in larger interpreted immobile pore-space estimates. Increasing the dispersivity in the model simulations decreased the estimated immobile porosity; flow rate had no impact. Overall, the results of this study highlight the difficulty faced in determining immobile porosity values in field settings, where hydrogeologic processes may vary temporally. Our results also highlight that immobile porosity is an effective parameter in an upscaled model whose physical meaning is not necessarily clear and that may not align with intuitive interpretations of a porosity.

Introduction

Non-Fickian or “anomalous” solute transport, characterized by features such as early arrival times, elongated tailing, and concentration rebounding, are not adequately described by the classic advection-dispersion equation (ADE) (e.g., Sudicky et al., 1985; Benson et al., 2000; Kosakowski et al., 2001; Berkowitz et al., 2006; Culkin et al., 2008). Some studies attribute the advection-dispersion equation's (ADE) inability to capture such experimental observations to the presence of small-scale zones, or domains (i.e., relative to the measurement scale) that are “immobile” with respect to bulk flow where advection controls the transport (e.g., Ceriotti et al., 2019). These immobile domains are the result of the natural physical- and chemical-property variability of porous media across a range of scales and flow rates (Harvey et al., 1994; Haggerty and Gorelick, 1995; Liu and Kitanidis, 2012; Bouquain et al., 2012). Their importance can be explicitly quantified in terms of a local Peclet number, a dimensionless number that represents the ratio of advective and diffusive transport rates. The accessibility of solutes to immobile domains is thought to be velocity dependent and can be inferred from the analysis of the Damkohler (DaI) number, a dimensionless number that relates the physical mass transfer rate of solutes between the mobile and immobile domains to the advective transport rate. Mass transfer between the two domains can be assumed to be instantaneous for DaI >> 1 and the mobile and immobile domains treated as single domain. For values of DaI << 1, advection dominates and the contribution of the immobile pore space to total transport may be small, even though long tails may still exist (Bahr and Rubin, 1987). Quantifying the contributions and volume of the immobile pore space is important as they likely control the presence of secondary contaminant sources that render remediation processes inefficient and slow (e.g., Haggerty and Gorelick, 1994; Harvey et al., 1994; Wilson, 1997).

One of the simplest groups of models capable of representing immobile pathways are dual-domain mass transfer (DDMT) models, developed as a modification of the ADE. These types of models represent the non-equilibrium conditions between two domains (Coats and Smith, 1964; van Genuchten and Wierenga, 2010; Ma and Zheng, 2011), which cannot be captured through the use of a retardation coefficient, which assumes instantaneous equilibrium between mobile and immobile regions in the ADE. Lab-scale (mm to m) (e.g., Coats and Smith, 1964; Swanson et al., 2012, Swanson et al., 2015), field-scale (m to km) (e.g., Feehley et al., 2000; Zheng et al., 2011; Briggs et al., 2018) and numerical studies (e.g., Haggerty and Gorelick, 1995; Wheaton and Singha, 2010; Liu and Kitanidis, 2012; Baratelli et al., 2014) have shown that DDMT models can match observations better than the traditional ADE under a variety of conditions. DDMT models have even been successfully applied in systems where velocity differences between the mobile and immobile domains are small (e.g., Baratelli et al., 2011) and in fractured rock where diffusion was considered negligible (e.g., Becker and Shapiro, 2000).

DDMT conceptually treats one domain as immobile with respect to flow; the mass transfer between mobile and immobile domains is generally modeled as a first-order exchange process mimicking diffusive mass transfer, although more complex higher-order exchange processes can also readily be represented (e.g., Haggerty and Gorelick, 1995; Carrera et al., 1998). The resulting transport of a solute through the mobile and immobile domains can be described according to:θmCmt=θmDmCmθmvCmRwhere θm is the porosity of the mobile domain [−], Cm is the dissolved species concentration in the mobile domain [kg/m3], v is the velocity in the mobile domain [m/s], and Dm is the hydrodynamic dispersion tensor [m2/s]. R is a source/sink term, where Rim is an exchange (reaction) term [kg/m3s] that is a component of R, which for a single-rate DDMT model is expressed as:Rim=θimαCmCimwhere θim is the porosity of the immobile domain [−], Cim is the dissolved species concentration in the immobile domain [kg/m3], and α is a first-order single-rate constant [1/s] and, along with other parameters within these transport models, is largely treated as a fitting parameter. In reality, concentration within “immobile” pore space may just have low advective velocities compared to the primary flowpaths. While some previous work has shown that these mass transfer parameters may be correlated to hydraulic conductivity or fracture length scales (e.g., Benson et al., 2001; Reeves et al., 2008), the physical meanings are generally considered to be difficult to correlate to anything meaningful in the field (e.g., Zhang et al., 2007; Willmann et al., 2008). Additionally, it has been recognized that velocity may control the estimated mass-transfer parameters (e.g., Haggerty and Gorelick, 1995) and they therefore display local Peclet number dependency (Ceriotti et al., 2019). While immobile zones are often conceptualized at the pore scale, field observations have indicated that they can be caused by large-scale heterogeneity as well (e.g., Briggs et al., 2018).

Recently, geophysical methods have been used to help constrain the mass-transfer parameters in Eq. (2)—specifically θim and α—during ionic tracer tests in the field given co-located fluid (σf) and bulk electrical conductivity (σb) measurements (e.g., Singha et al., 2007; Hampton et al., 2019; MahmoodPoor Dehkordy et al., 2018, MahmoodPoor Dehkordy et al., 2019). The σf measurements preferentially sample well-connected pore spaces (i.e., the mobile domain), whereas σb measurements are thought to collectively capture both the mobile and immobile domains (Singha et al., 2007); the σb measurements provide a way to “see” into the immobile domain and estimate the parameters from Eq. (2) at particular locations. In a single-domain medium, co-located σf and σb measurements are expected to respond similarly to an ionic tracer. In a dual-domain medium on the other hand, co-located σf and σb will deviate from one another due to differences in the solute loading and unloading rate(s) between the mobile and immobile domains; this difference allows these key mass-transfer parameters to be estimated.

While many column experiments have been used to calculate the parameters from Eq. (2) (e.g., Swanson et al., 2012, Swanson et al., 2015; Briggs et al., 2014), there are currently no larger-scale, controlled experiments reported in the literature that robustly explore the estimation of immobile porosity in the presence of large-scale heterogeneity. Here, we similarly look to explore estimates of θim in a mesoscale 3-D sand tank, with large impermeable blocks emplaced that produce a nonuniform flow field. These barriers produce low local velocities that result in tank-scale immobile zones. This work is motivated in part by that of Briggs et al. (2018), who hypothesized that their immobile porosity estimates were controlled by largely stagnant zones downgradient of large cobbles within glacially deposited lake sediments at their field site. We additionally explore the controls of density-dependent tracers, typically neglected in field tracer tests but clearly a process that affects transport in many settings (e.g., Beinhorn et al., 2005). Co-located measurements of σf and σb were collected at a variety of locations within the tank, and those data were then used with graphical and numerical models to estimate θim throughout the tank; inverted maps of σb were also used to image the transport of the tracer. A numerical model of the soil tank was also created as part of this work to support the experimental efforts. Flow velocities were varied in the model to control the global Peclet number to provide insight into how θim estimates and the contributions of advective and diffusive transport change in the presence of large-scale heterogeneities. As the generation of accurate data needed under controlled conditions is not feasible in the field, this study used an existing intermediate scale test tank where the boundary and initial conditions can be controlled and the soil conditions are accurately known.

Section snippets

Experimental methods

A 429 × 244 × 36 cm intermediate-scale synthetic aquifer at the Center for Experimental Study of Subsurface Environmental Processes (CESEP) was used for the tracer testing and geophysical tracking (Fig. 1). The tank was packed with dry Unimin Granusil #70 sand around four impermeable plastic barriers to create stagnant zones. The sand used in these experiments was previously characterized experimentally to have a hydraulic conductivity of 0.864 cm/min (Sakaki and Illangasekare, 2007) and was

Numerical modeling

Because the tank experiments were time consuming—requiring months of time to prepare and execute—exploring how velocity affects the estimates of immobile porosity was conducted with numerical models. Water flow and solute transport were simulated in six tracer scenarios (Table 1) and coupled with electrical flow modeling in R3t (Binley, 2019), which was also used to invert the experimental data to create tomograms of tracer movement. Hydraulic gradients matching those of the tank experiment and

Results and discussion

When barriers were included in the flow and transport models, the magnitude of peak concentrations from the numerical simulations matched observed data in the upper portion of the synthetic aquifer as the dispersivity increased at the selected wells (14, 19, 20, 25, and 36), both with and without density dependence considered; however, the model often underpredicted the magnitude of σf at the bottom of the synthetic aquifer (Fig. SI-1). This observation likely indicates that the vertical

Conclusions

Barriers, velocity, and density-dependent flow control the observed hysteresis between σb versus σf in 3-D tank experiments and associated models, and thus the estimates of θim. Estimated θim was found to increase in numerical models in the presence of impermeable barriers, which produced stagnant zones with relatively slow flow velocities. Estimates of θim increased slightly with density dependence. However, there was no statistically significant difference in estimating θim under simulated

Declaration of Competing Interest

None.

Acknowledgments

Funding for the research presented in this manuscript was provided by the National Science Foundation (EAR-1446235). We thank Jackie Randell for her support of the tank work. Data and model results from this manuscript are available through CUAHSI's HydroShare at http://www.hydroshare.org/resource/2a2ccab0b4ff4c9f9be97a9aff4f0b27.

References (61)

  • G.R. Barth et al.

    A new tracer-density criterion for heterogeneous porous media

    Water Resour. Res.

    (2001)
  • M.W. Becker et al.

    Tracer transport in fractured crystalline rock: evidence of nondiffusive breakthrough tailing

    Water Resour. Res.

    (2000)
  • D.A. Benson et al.

    The fractional-order governing equation of Levy motion

    Water Resour. Res.

    (2000)
  • D.A. Benson et al.

    Fractional dispersion, Lévy motion, and the MADE tracer tests

    Transp. Porous Media

    (2001)
  • B. Berkowitz et al.

    Modeling non-Fickian transport in geological formations as a continuous time random walk

    Rev. Geophys.

    (2006)
  • A. Binley

    R3t Manual

    (2019)
  • J. Bouquain et al.

    The impact of inertial effects on solute dispersion in a channel with periodically varying aperture

    Phys. Fluids

    (2012)
  • M.A. Briggs et al.

    Dual-domain mass-transfer parameters from electrical hysteresis: theory and analytical approach applied to laboratory, synthetic streambed, and groundwater experiments

    Water Resour. Res.

    (2014)
  • M.A. Briggs et al.

    Direct observations of hydrologic exchange occurring with less-Mobile porosity and the development of anoxic microzones in Sandy lakebed sediments

    Water Resour. Res.

    (2018)
  • J. Carrera et al.

    On matrix diffusion: formulations, solution methods and qualitative effects

    Hydrogeol. J.

    (1998)
  • K.H. Coats et al.

    Dead-end pore volume and dispersion in porous media

    Soc. Pet. Eng. J.

    (1964)
  • R.A. Cox et al.

    A new total variation diminishing scheme for the solution of advective-dominant solute transport

    Water Resour. Res.

    (1991)
  • S.L. Culkin et al.

    Implications of rate-limited mass transfer for aquifer storage and recovery

    Ground Water

    (2008)
  • A.D. Eaton et al.

    5210 B. 5-Day biochemical oxygen demand (BOD) test

  • C.E. Feehley et al.

    A dual-domain mass transfer approach for modeling solute transport in heterogeneous aquifers: application to the macrodispersion experiment (MADE) site

    Water Resour. Res.

    (2000)
  • S.P. Friedman

    Soil properties influencing apparent electrical conductivity: a review

    Comput. Electron. Agric.

    (2005)
  • S.P. Friedman et al.

    Particle shape characterization using angle of repose measurements for predicting the effective permittivity and electrical conductivity of saturated granular media

    Water Resour. Res.

    (2002)
  • L.W. Gelhar et al.

    A critical review of data on field-scale dispersion in aquifers

    Water Resour. Res.

    (1992)
  • C. Geuzaine et al.

    Gmsh: a 3-D finite element mesh generator with built-in pre- and post-processing facilities

    Int. J. Numer. Methods Eng.

    (2009)
  • R.R. Goswami et al.

    Comparison of numerical techniques used for simulating variable-density flow and transport experiments

    J. Hydrol. Eng.

    (2011)
  • Cited by (10)

    • Quantification of solute transport in a fracture-matrix system using geoelectrical monitoring

      2023, Journal of Hydrology
      Citation Excerpt :

      However, accurate characterization and prediction of solute transport in fractured rock is particularly challenging owing to the extreme heterogeneities of flow velocity and matrix porosity (Dou et al., 2018; Chen et al., 2022; Viswanathan et al., 2022). Over the past several decades, the hydrological and environmental communities have extensively highlighted the complexity of the solute transport process in fractured rock system (Cardenas et al., 2007; Chen et al., 2021; Qian et al., 2011a; Rezaei et al., 2022; Rezaei et al., 2016; Wang et al., 2022; Zhan et al., 2009; Zhou et al., 2018; Zhou et al., 2017; Zhu et al., 2018). Many studies have shown that preferential flow in rock fracture promotes fast transfers, while diffusion from rock fracture into matrix can results in the occurrence of solute heavy tailing (Hyman et al., 2019).

    • Improving predictions of solute transport in a laboratory sandbox aquifer through high-resolution characterization with hydraulic tomography

      2022, Journal of Hydrology
      Citation Excerpt :

      Regardless of the vigorous debates over the validity of ADE in the groundwater research community, accurate characterization of subsurface heterogeneity has been frequently proven to be one key issue in improving predictions of solute transport (Salamon et al., 2007; Zheng et al., 2011; Illman et al., 2012; Guo et al., 2021). So far, numerous studies have been published to understand the benefits of high-resolution characterization of subsurface heterogeneity, as well as the need in consideration of geological information, for the reliable predictions of plume migration (e.g., Liu et al., 2010; Ronayne et al., 2010; Händel and Dietrich, 2012; Foster et al., 2021). Under controlled conditions, laboratory sandbox experiments have been conducted with prescribed heterogeneity patterns mimicking randomly distributed (e.g., Russell et al., 1996; Silliman et al., 2001; Levy and Berkowitz, 2003; Fernàndez-Garcia et al., 2004) or deterministically structured (e.g., Sudicky et al., 1985; Pedretti et al., 2016; Maina et al., 2018) K fields, to show the impacts of aquifer heterogeneity on solute migration and on the reproduction of BTCs.

    • Improved high-resolution characterization of hydraulic conductivity through inverse modeling of HPT profiles and steady-state hydraulic tomography: Field and synthetic studies

      2022, Journal of Hydrology
      Citation Excerpt :

      Accurate descriptions of hydraulic conductivity (K) distributions have long been acknowledged as critical for groundwater flow and contaminant transport related issues (Gelhar, 1986; Carrera et al., 2005; Sudicky et al., 2010; Foster et al., 2021).

    View all citing articles on Scopus
    View full text