0% found this document useful (0 votes)
223 views36 pages

Saatsaz

This document discusses groundwater modeling and its concepts, classifications, and applications for solute transport simulation. It begins by explaining that groundwater models can help hydrogeologists understand and predict groundwater systems. It then discusses some common applications of groundwater models, including characterization, prediction, engineering, management, and teaching. The document notes that groundwater model development requires simplifications and assumptions to represent the system adequately while balancing complexity. Simpler models may be less accurate than more complex models that account for heterogeneity.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
223 views36 pages

Saatsaz

This document discusses groundwater modeling and its concepts, classifications, and applications for solute transport simulation. It begins by explaining that groundwater models can help hydrogeologists understand and predict groundwater systems. It then discusses some common applications of groundwater models, including characterization, prediction, engineering, management, and teaching. The document notes that groundwater model development requires simplifications and assumptions to represent the system adequately while balancing complexity. Simpler models may be less accurate than more complex models that account for heterogeneity.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 36

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/344459959

GROUNDWATER MODELING AND ITS CONCEPTS, CLASSIFICATIONS, AND


APPLICATIONS FOR SOLUTE TRANSPORT SIMULATION IN SATURATED POROUS
MEDIA

Preprint · October 2020


DOI: 10.13140/RG.2.2.11923.71200/1

CITATIONS READS

0 1,951

2 authors:

Masoud Saatsaz Saeid Eslamian


Institute for Advanced Studies in Basic Sciences Isfahan University of Technology
30 PUBLICATIONS   194 CITATIONS    686 PUBLICATIONS   9,108 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

The Effect of Dust on the Suspended and Bed Load of Karkheh River View project

Understanding river channel transmission losses in near real-time View project

All content following this page was uploaded by Masoud Saatsaz on 17 August 2021.

The user has requested enhancement of the downloaded file.


GROUNDWATER MODELING AND ITS
CONCEPTS, CLASSIFICATIONS, AND
APPLICATIONS FOR SOLUTE TRANSPORT
SIMULATION IN SATURATED POROUS MEDIA
Masoud Saatsaz*
Department of Earth Sciences, Institute for Advanced Studies in Basic
Sciences (IASBS), Zanjan, Iran
Saeid Eslamian
Department of Water Engineering, College of Agriculture, Isfahan University of
Technology, Isfahan, Iran

Saatsaz, M., & Eslamian, S. (2020). Groundwater Modeling and Its Concepts,
Classifications, and Applications for Solute Transport Simulation in Saturated
Porous Media. In Eslamian & Eslamian, Advances in Hydrogeochemistry
Research (2020), Nova Science Publishers, Inc.

ABSTRACT
Numerical modeling of flow and solute transport has received increased
attention during the last three decades. Groundwater models can offer the most
accurate representation, characterization, and prediction of the hydrogeologic
processes. This chapter discusses (i) the basic concept of groundwater
modeling, (ii) the classification of groundwater models (general view), (iii)
the application of solute transport models, (iv) the identification of
hydrogeochemical processes, (v) the governing equation of solute transport,
(vi) the guide for the development of coupled flow and solute transport
numerical models, (vii) the applicability of various kinds of models to simulate
the fate and transport of the major ions, (viii) the inverse estimation of solute
transport properties, and (ix) the evaluation of approaches for modeling
groundwater contamination (specific view). One clear conclusion of this study
is that groundwater models should consider a series of hypothesizes,
assumptions, constraints, and simplifications based on the model objective,
modeling techniques, model reliability, solution complexity, data availability,

* Corresponding Author address


Email: [email protected]
2 Masoud Saatsaz

and characteristics of the modeled area. Although numerical models provide


a more comprehensive vision of the groundwater systems under complicated
conditions, these models are not suitable when the input data are not updated,
reliable, sufficient, and precise. As the groundwater flow velocity mainly
controls the magnitude of the movement, storage, and change in the solute
concentration, an accurate definition of the flow system is of great
significance. This chapter highlights the advection-dispersion equation suffers
from several conceptual limitations, especially in large scale heterogeneous
systems. To overcome these difficulties, advanced groundwater models can
solve complex problems, either one- or multidimensional, that often involve
multiple chemical species, diverse transport processes, material heterogeneity,
complex boundary conditions, and time-varying stresses.

Keywords: Finite difference method, Finite element method, Groundwater,


Hydrogeochemical processes, Saturated porous media, Numerical solute
transport model

Introduction

Men have always tried to improve their abilities in presenting the truth.
Depending on system complexity, techniques, sources (data, budget, and time), user
experience, and needs, the task may be easy or difficult. In this way, modeling is a
practical way used to define, represent, quantify, analyze, test, document, teach or
advertise something else using physical objects, texts, symbols, graphs, equations,
structures, computer codes, and a variety of other things. Scientific modeling is a
scientific process, the aim of which is to generate an abstract, conceptual
framework, and graphical or mathematical representation of an object under
examination. According to Schwarz et al. (2009), a scientific model is a
representation of a particular phenomenon (states, objectives, and events) that
abstracts and simplifies it to allow one to make explanations and predictions.
Modern science offers a growing collection of methods, techniques, theories, tools,
and facilities about all kinds of specialized scientific modeling (Prior 2017).
Generally, scientific models at best are not exact replicas; these are
approximations of the systems and objects that they represent (Pallant and Lee
2015). To make models more realistic, the modeling process has recently become
a problematic task for model designers. They routinely follow several formalized
and standardized steps, allowing models to be faster to set up, more secure to run,
more accurate to present, more comfortable to understand, and more flexible to
meet the needs. Practical issues in developing a scientific model consist of (i) the
establishment of model objectives, (ii) data gathering, classification, and
Groundwater modeling and its concepts, classifications, and applications for 3
solute transport simulation in saturated porous media
preparation, (iii) selection of the system’s main components, functions, behaviors,
and problems, (iv) identification and documentation of constraints, limitations, and
assumptions within the simulation model, and (v) validation of the simulation
results (Gray and Rumpe 2016).This chapter aims to provide a brief but
comprehensive explanation of principles, concepts, theories, and applications of
flow and solute transport modeling. This research answers (i) what are the main
processes controlling solute transport; (ii) how the mathematical equations describe
these processes; (iii) how numerical methods solve the governing equation; and (iv)
how the accuracy of the model is measured. It is important to note that the focus is
on non-reactive miscible solutes that pass through a saturated porous medium.

Groundwater Modeling

Groundwater resources are hidden systems and dynamic; their quality and
quantity are often hard to measure and understand. The state of any hydrogeological
system can change over time. In this regard, model designers and operators are
repetitively developing new scientific models and solution techniques to evaluate
these systems. Groundwater models can help hydrogeologists explain the behavior
of the whole or a part of a groundwater system, understand the processes, make
predictions, make management decisions, and test a hypothesis (Figure. 1). Some
of the typical groundwater model applications are:
(i) To improve hydrogeological understanding by analysis of data and
existing literature (characterization and planning);
(ii) To evaluate the aquifer behavior through simulation (prediction);
(iii) To identify practical solutions to meet objectives (engineering construction
and design);
(iv) To optimize designs for maximum performance, minimum costs and
environment conservation (control and optimization);
(v) To assess input, output, and water balance factors (water availability
assessment);
(vi) To estimate the predevelopment conditions of an aquifer system
(development and enhancement);
(vii) To predict the behavior of the groundwater system under different
hydrogeological stresses (management and decision-making);
(viii) To quantify the sustainable and safe yield (water allocation policies);
(ix) To analyze the sensitivity and optimization of parameters (uncertainty or
risk analysis);
(x) To visualize the behaviors, processes and ongoing and upcoming events of
a groundwater system (scaling and representation);
4 Masoud Saatsaz

(xi) To point out the strengths and weaknesses of other research and evaluating
other studies and theories through modeling (learning and teaching).

Groundwater Model Simplification

Similar to other models, the development of groundwater models requires the


consideration of a series of hypotheses, assumptions, constraints, and
simplifications to ensure that an idealization or abstraction sufficiently expresses
an idea. Simplifications may decrease the accuracy of the model. For example,
when the migration of chemicals from a point-source is simulated and predicted by
a simple model that assumes homogeneous groundwater conditions, the results may
be less accurate than those of a heterogeneous groundwater system. As another
example, a regional scale model with a 100 × 100 m cell size would not account for
preferential solute transport through the relatively small open framework gravel
lenses (Dommisse 2015). The range of assumptions and constraints of a
groundwater model depends upon the model objective, modeling techniques, model
reliability, solution complexity, data availability, and characteristics of the modeled
area. According to Beven (2009), a model should not be more complicated than is
needed to make sufficiently useful predicts of the observations. In more complex
groundwater models, a more detailed understanding of the groundwater system is
necessary. Hence, a logical relation between the data availability and model
complexity should be considered for better presentation. Selecting complex
systems, when there is a shortage of information, would not end in desirable results
and would not show the real world. In such situations, operators should revise their
expectations of the model’s ability based on data availability and accessibility. The
trade-offs between model complexity and simplicity are discussed by Konikow
(2011) and Hill (2006).
Groundwater models can be classified based on the result’s quality into the
board, detailed, and detailed, coupled with predictive modeling classes (Middlemis
et al. 2001). The first is basic models and low-cost models, fit for a baseline
evaluation, limited planning, and data gap identification. As these models do not
need significant information to develop, they are not appropriate for evaluating
complex conditions. The second models are relatively complex, requiring more data
and a better understanding of the groundwater system. These models are generally
suitable for characterizing suspected groundwater issues. The third models (also
known as aquifer simulators) are too specific, time-consuming, and costly. These
models can predict the changes in groundwater bodies and their interactions with
surface water on a large scale (aquifer or watershed). Their outputs can be used to
Groundwater modeling and its concepts, classifications, and applications for 5
solute transport simulation in saturated porous media
support comprehensive groundwater management and protection plans (Payne and
Woessner 2010).

Figure 1. The development and application of the modeling in the analysis of


groundwater problems, decision making, hypothesis testing (Adapted from
Konikow et al. 1996)

Groundwater Model Classification

Groundwater models can be classified using a variety of independent methods.


Based on structure, groundwater models are routinely classified into three
categories, including physical, analog, and mathematical. Mathematical models are
divided into two main categories, including deterministic and stochastic (statistical
models). The deterministic models can be analytical or numerical.

Physical Groundwater Models

Physical models (also known as porous media models or mocked-up models)


such as sand-box, bench-scale, parallel-plate, and viscous fluid models are an
abstracted and scaled-down laboratory representation of a groundwater system. The
earliest works on groundwater modeling were physical models that appeared in
European literature. The Darcy’s experiments (1855) describing the flow of fluids
6 Masoud Saatsaz

through porous media can be regarded as the first works in groundwater modeling
(Schweizer 2015; Landmeyer 2011; Simmons 2008). By applying this pioneering
model, Darcy could experimentally demonstrate the law of water flow through
saturated sand (now known as Darcy's law). Physical models can be currently
applied to simulate seepage, artificial recharge, dispersion, and seawater intrusion,
in situations that little is known about the processes and dynamics that act in a
groundwater system. The most significant problem in physical models is finding a
proper scale, so that be able to represent a real groundwater system in the most
effective way possible. Besides, a small number of situations can be described by
these models, mostly supporting the hydraulic properties of the porous media.
Another shortcoming associated with physical groundwater models is that the setup
of such models is relatively costly and time-consuming. Despite all the
disadvantages, these models can be used as an appropriate educational tool in the
hands-on learning of the basic concepts of groundwater hydrology. A description
of the pioneering works related to numerical modeling of the groundwater in porous
media is provided by Pinder and Bredehoeft (1968).

Analog Groundwater Models

Analog models utilize the correspondence between Darcy’s Law and other laws
of physics, such as Fourier's law of heat transfer, Ohm's law of electricity, Navier-
Stokes equation of fluid motion, and Fick's law of solute diffusion to describe the
dynamics of a groundwater system (Delleur 2010).
Membrane, thermal, conductive liquid, and electrical models are common
forms of groundwater analog models. For example, to simulate aquifer storage
(capacitance) and groundwater flow (resistance), electric analog models used a
network of capacitors and resistors, arranged in a grid to approximate aquifer
geometry. Here, the analogy between the electrical flow in a material conductor
(Ohm’s law) and the water flow in a porous media (Darcy’s law) can be stated
mathematically as (Thangarajan 2007):

𝑑𝑉
𝐼 = −𝜎 𝑑𝑥
(Ohm’s law) (1)

where 𝐼 is the electrical current per unit cross-sectional area, 𝜎 is the specific

𝑑𝑉
electrical conductivity of a substance, and 𝑑𝑥 is the voltage (potential) gradient 𝑉 .
Groundwater modeling and its concepts, classifications, and applications for 7
solute transport simulation in saturated porous media
𝑑ℎ
𝑞 = −𝐾 𝑑𝑥
(Darcy’s law) (2)

where 𝐼 is the Darcy velocity, and is similar to the electrical current stated in
equation (1). The hydraulic conductivity 𝐾 is alike the electrical specific
conductivity 𝜎, and hydraulic head ℎ corresponds to the electric potential.

Mathematical Groundwater Models

The behavior of a groundwater system can be mathematically formulated and


represented by a set of equations, variables, and numbers constituting a
mathematical model. These equations and related parameters are formulated based
on the geometrical and hydrogeological characteristics of the groundwater system,
spatial and temporal boundary conditions, initial conditions, and stresses of the
system. In Figure 2, a schematic shows the connections between the elements of a
simplified solute transport numerical model.
As mentioned, mathematical models are divided into deterministic and
stochastic (statistical) models. Stochastic models are provided with possible
outcomes and solutions based on spatial variabilities and probabilities of
occurrence. Deterministic models are established based on the cause and effect
relationships (one process, a cause, contributes to the production of another state,
an effect) of known systems and processes. These types of models are typical for
identifying local and regional groundwater systems. The governing equations of
deterministic models can be solved with analytical (if the model is relatively
simple), numerical, or a mixture of analytical and numerical techniques.

Figure 2. Schematic showing a simplified solute transport numerical mathematical


model
8 Masoud Saatsaz

Analytical Groundwater Models

The first analytical solution is provided by Dupuit (1863) to describe the


steady-state groundwater flow to pumping wells in confined or unconfined aquifers
(Renard 2005). Twenty-three years later, in 1886, Dupuit's approach was extended
and generalized by an Austrian engineer, Philipp Forchheimer. The Theis solution
(1935), as one of the commonly used analytical solutions in groundwater hydrology
that considers different types of aquifers, flow regimes, stresses, and boundary
conditions. Through the ages, analytical methods have become standard tools for
explaining the dynamics of groundwater quantity and quality. Also, these models
are occasionally applied before a numerical model starts. These models have their
disadvantages concerning mathematical assumptions that may not occur naturally.
Another problem is that they do not allow for simulating spatial and temporal
variations in various parameters simultaneously. Hence, these models are often
inadequate to mimic a heterogeneous subsurface system closely.

Numerical Models

In the late 1960s, following the development of computers and digital


technology, numerical models were widely adapted to provide a more
comprehensive vision of the groundwater systems under complicated conditions.
With numerical modeling, hydrogeologists were directly able to approximate the
solutions of complicated differential equations of the system by converting them to
discrete equations and dissecting the domain into meshes or grids. In numerical
models, dissection of a region and discretization of differential equations are
generally accomplished by approximation approaches like finite difference, finite
element, and finite volume boundary. Each of these approaches can be applied
solely or combined with other methods to reduce the computational complexity of
numerical models. Considering the applicability, ease of use, performance, and
computational speed, each method has its advantages and disadvantages. Among
them, however, finite difference and finite element methods are particularly the
most preferred approaches for solving groundwater flow and transport equations
(Wang et al. 2019).

Finite Difference Method

The finite difference methods are perhaps the best known among the numerical
techniques used to solve a set of partial differential equations based on the geometry
of the system, boundary conditions, and initial conditions. This method was first
Groundwater modeling and its concepts, classifications, and applications for 9
solute transport simulation in saturated porous media
introduced by Richardson in 1910 and was improved later in 1927 (Raats and
Knight 2018). In this method, a structured grid is divided into regular-shaped small
cells or grid points (nodes), the derivatives of the partial differential equations are
approximated in space, concerning to the differences between values of variables
at adjacent cells or nodes. The nodes represent the specific values of all
computational properties of the system (e.g., hydraulic head, solute concentration,
effective porosity, etc.) as functions of time. All nodes are defined in a body-fitted
coordinate system and labeled consecutively by a set of indices (i, j, and k along
the x, y, and z directions, respectively). The distance between neighboring cells or
nodes can be varied. The continuous medium can be discretized either into a block
or mesh centered node (point centered) domain. To solve a finite difference
equation, it is necessary to start with the initial distribution of heads or solute
concentration and compute their values at later time instants. The relation between
initial (observed) and calculated locations of particls in a finite-difference grid is
shown in Figure 3.
In the finite difference method, the solution is obtained by an iterative process.
An iterative process is an on-going process conducted for calculating the desired
result through a repeated cycle of operations. An iterative process should come
closer to the ideal outcome as the number of iterations increases, i.e., it should be
convergent. Some of the prevalent iterative numerical methods have been referred
by Thangarajan (2007), that are (i) Alternating Direction Implicit (ADI) Procedure
(Peaceman and Rachford, 1955); (ii) Successive Over-Relaxation (SOR) method
(Young, 1971); (iii) Modified Iterative Alternating Direction Implicit Method
(Prickett and Lonnquist, 1971); (iv) Strongly Implicit Procedure (SIP) (Stone,
1968).

Figure 3. A hypothetical finite difference grid depicting the relation between flow
lines and computed flow paths (Adapted from Konikow and Bredehoeft 1978)
10 Masoud Saatsaz

Finite Element Method

The finite element method, in its presently known form, was first theorized by
Turner et al. (1956) and used to solve civil engineering and construction problems
(Zeinkiewicz 1971). In this method, the grid is discretized into a network of nodes
that form the vertices of irregularly/regularly-sized triangular elements. The
elements and nodes are used for the interpolation of the field parameters such as the
hydraulic head and solute concentration. The elements may have different spatial
dimensions, orientations, and sizes. Where the size of triangles decreases (i.e., a
dense array of fine elements), more details is required. In the finite element method,
systematic approaches, such as the Weighted Residual Method, are needed to
generate the finite element equations. The Galerkin method is the most common
method of the finite element method for solving groundwater problems. First, an
approximate solution based on the behavior of the dependent variable of the system
is assumed. The obtained solution is often selected to satisfy the boundary
conditions of the model (Alexandrov 2004). The residual is the difference between
the actual and assumed solution. The Galerkin method minimizes the weighted sum
of the residuals of the differential equation for each of the finite elements until it
reaches the best approximate solution over the domain.
The main disadvantages of the finite element methods over finite difference
methods are their complexity to formulate, difficulty to solve, and time required to
use. Conversely, these methods have more flexibility than finite difference methods
to approximate a geometrically irregular mesh. However, a large amount of
calculation and storage memory for software simulation is necessary to
approximate an unconditionally irregular scheme.

Finite Volume Method

In the finite volume models, the small volume surrounding each node point on
a mesh is measured based on the differential form of the conservation equations a
control volume. This method relies primarily on the concept that the mass of a
dependent variable that diverges out from a control volume (or a finite volume) is
equal to the mass that passes across the boundary. When the entire flux is integrated
over the area and time, the total variation of the mass in each control volume can
be approximated (Rapp 2016). It means the change of the mass inside a control
volume plus the total mass fluxes through the boundary should be zero (Neill and
Hashemi 2018). Like in the case of the finite element method, this method is very
flexible, it can be easily implemented for irregular and complex geometries.
Groundwater modeling and its concepts, classifications, and applications for 11
solute transport simulation in saturated porous media
Another advantage is that the approach can enjoy an axisymmetric non-body fitted
coordinate system that allows the grid lines to be non-orthogonal, where edges are
not aligned with the horizontal and vertical axes. Further details on these methods
may be found in the following references: Scesi and Gattinoni (2012); Constanda
(2016); Howell (2018); Argyros et al. (2018); Dobrushkin (2017); Moatamedi and
Khawaja (2018).

Flow and Solute Transport Models

All mentioned above have confirmed the practice of groundwater modeling has
a lengthy history and comes in many different forms. At first, however,
groundwater models focused on the quantitative aspects of groundwater.
Groundwater flow models have been routinely used to calculate the rate, velocity,
and direction of groundwater flow within and across the boundaries of an aquifer
system (Mandle 2002). The outputs of a groundwater flow model are the hydraulic
heads and groundwater fluxes that are in equilibrium with the physical and
hydrogeological conditions defined for the model. Over time, the focus of
hydrogeologists has shifted from quantitate to qualitative issues. Groundwater
solute transport models have been extensively developed to simulate the
distribution and fate of various chemical species in different hydrogeological
settings since the mid-1960s. These models could simulate the flux and
concentration of solutes concerning the hydrogeological, geochemical, and
spatiotemporal conditions. Since the magnitude of the movement, storage, and
change in the concentration (transformation) of solutes are mainly controlled by the
groundwater flow velocity (the flow distance for a given time), an accurate
definition of the flow system is crucial to simulate solute transport problems.
In general, the term “solute transport model” is used to describe models that
simulate the ‘movement, storage, and transformations’ of a soluble chemical
species (contaminant). In a solution, a solute constitutes only a small portion of a
solvent (Bear and Cheng 2016). Since the most accessible groundwater resources
are in the alluvial plains, the modeling of solute transport through porous media has
received a lot of attention. According to Lage and Narasimhan (2000), a porous
medium is “a region in space comprising of at least two homogeneous material
constituents, presenting identifiable interfaces between them at a resolution level,
with at least one of the constituent remaining fixed or slightly deformable”.
12 Masoud Saatsaz

Application of Solute Transport Models

Solute transport modeling in porous media deals with hydrogeological,


hydrological, climatological, agricultural, industrial, geoecological, geographical,
environmental, hydraulic engineering, water economics, and health issues. The
specific and regulatory purposes of these models in hydrogeology include:
(xii) To analyze the mass balance of chemicals;
(xiii) To interpret concentration data;
(xiv) To describe the bulk and large scale motion of chemicals (advection or
convection); or random, small scale motion of chemicals (diffusion);
(xv) To assess the degree of mixing or dilution along the flow path (longitudinal
dispersion) and perpendicular to the flow path (transverse dispersion);
(xvi) To predict chemical processes and the changes in chemical concentration
taking place by dissolution, retardation, and degradation (e.g., adsorption,
precipitation, hydrolysis, cation exchange, bioaccumulation, biodegradation, redox
reaction);
(xvii) To develop a monitoring strategy and risk assessment;
(xviii) To estimate the cleanup time using remediation plans;
(xix) To manage and protect groundwater quality

Migration of Solutes in Groundwater

Contaminants can be mobilized and released into the groundwater from various
sources. The movement of soluble chemicals in the saturated zone of a groundwater
system represents a miscible displacement process, where a distinct interface
doesn’t exist in the fluid. The solubility of a chemical species is the most significant
solute characteristic (Ebrahim 2013), defined as the maximum mass of the chemical
species that can be dissolved in a unit quantity of the solvent under specified
conditions. The solubility of a substance depends on the nature and amount of solute
and solvent, presence of other ions, as well as the pH of the solution, temperature,
and pressure (Saatsaz et al. 2013a). Chemical components with low solubility are
often referred to as immiscible with water. Dense and light non-aqueous phase
liquids (DNAPLs and LNAPLs) are examples of liquid solution contaminants that
do not dissolve in or easily mix with water. A more detailed description of NAPLs
can be obtained from, for example, Müller and Sedláčková (2003); Mayer and
Hassanizadeh (2005); Rubin et al. (2013); Bharagava (2017).

Typical solute transport models are advection, advection-dispersion, and


advection-dispersion-chemical/biological reaction models (Bobba 2012). The
Groundwater modeling and its concepts, classifications, and applications for 13
solute transport simulation in saturated porous media
movement of solutes through groundwater flow, where there is no change in the
solute concentration with distance, is mainly described by advection models.
According to Darcy’s law of groundwater flow, advection is attributed mostly to
the effective porosity, hydraulic conductivity, and hydraulic gradient of the
medium.

Hydrogeochemical Processes and their Models

Typical solute transport models are advection, advection-dispersion, and


advection-dispersion-chemical/biological reaction models (Bobba 2012).
Advection-dispersion models describe the transport of solutes when solute
concentration is changed in time and space due to diffusion and dispersion. These
models have been developed based on the diffusion dispersion concept. The
dispersion term describes a mechanical mixing process caused due to the
heterogeneity of the medium (Pandey et al. 2018). This process allows solutes to
spread in both longitudinal and transverse directions. The spreading rate depends
upon the saturation degree and the value of water velocity in porous media, the rate
of solute mixing because of convergence and divergence of flow paths, and
molecular diffusion (Leij and van Genuchten 2002). Diffusion is a gradual
movement of individual solutes from one part with higher concentration to another
section with lower concentration. This process is carried by random molecular
motions due to the concentration gradient, with or without movement of water
(Demirel 2013). In the aquitard sediments, in which the hydraulic conductivity
values are relatively low, the advective proportions are negligible. In such a
condition, the diffusion effect is usually modeled because it becomes the controlling
factor of mass transport.
Advection-dispersion-chemical/biological reaction models are inherently more
complex. These models are frequently used to quantify the change in the
concentration of contaminants when chemical or biological reactions incorporate
into advection-dispersion processes. Other main mechanisms within the
groundwater systems that affect the transport, persistence, and fate of chemicals
include:
(i) Retardation is the act or result of delaying by which a solute becomes attached
to sediments and the geologic rack material and removed from the water. Often the
sorption is a reversible process. When solutes desorb, contamination plumes are
retarded, rather than solutes being permanently retained by the aquifer solids (Rivett
et al. 2006).
14 Masoud Saatsaz

(ii) Filtration is a process to separate components based on their relative particle


sizes. This process affects particulate contaminants (contaminants consist of
particles suspended in the water) rather than dissolved solutes.
(iii) Precipitation is the removal of dissolved and suspended solids from solution
by the formation of insoluble compounds through chemical means.
(iv) Cation exchange is the interchanging between cations in the solution and
cations on the surfaces of clay particles or organic colloids.
(v) Biodegradation is the transformation of a substance into new compounds
through biochemical reactions or the actions of microorganisms (Insam and
Bertoldi, 2007). Biodegradation is a complex process by which an aquatic system
can depollute chemicals (Ney et al. 1981). In general, soluble solutes are more
biodegradable than those with low water solubility (Powell 2017).
(vi) Hydrolysis occurs in saturated and unsaturated zones of the groundwater
systems so that a molecule of water is added to a substance to alter the chemical's
structure and break down its bonds. The rate of hydrolysis reaction s depends on
many factors such as temperature, pH, solubility, adsorption or absorption, and the
redox (oxidation-reduction) potential (Richens 1997).
(vii) Dissociation or ionization is the separation of a compound into separated anion
and cation when the chemical becomes electrically charged through the redox
reaction (IUPAC 2014).
(viii) Bioaccumulation is the gradual accumulation of chemicals into the living
tissue of an organism, such as plants or animals (Alexander 1999). As living
organisms are not generally found in groundwater, the bioaccumulation process in
in groundwater is negligible.. A simplified graphic of the described
hydrogeochemical processes for the one-dimensional case is given in Figure 4.
Groundwater modeling and its concepts, classifications, and applications for 15
solute transport simulation in saturated porous media

Figure 4. The effect of the advection (A), advection and dispersion/diffusion (B)
advection, dispersion/diffusion, and retardation (C), and advection,
dispersion/diffusion, retardation and degradation (D), on the concentration of
chemical species along a flow path (Adapted from Merkel et al. 2005)

Further information on physicochemical processes that control the fate and


transport of chemicals in groundwater is provided by Freeze and Cherry (1979);
Stumm and Morgan (1996); Bedient et al. (1999); Fetter (2018); Kovarik (2000);
Schwartz and Zhang (2003); Appelo and Postma (2004); Merkel et al. 2005; Clark
(2015); Scozzari and Dotsika (2016); Tikhomirov (2016); Dassargues (2018).

Governing Equations
The form of the governing equations for three-dimensional flow in saturated
porous media is a function of the properties of the water (i.e., density and viscosity),
the features of the porous media (i.e., intrinsic permeability, grain size, shape
factor), and the gradient of the hydraulic head. Darcy’s law represents a generalized
form as follows::
16 Masoud Saatsaz

𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝑑ℎ
𝜕𝑥
(𝐾𝑥𝑥 𝜕ℎ) + 𝜕𝑦 (𝐾𝑦𝑦 𝜕ℎ ) + 𝜕𝑧 (𝐾𝑧𝑧 𝜕ℎ) − 𝑄 = 𝑆𝑠 𝑑𝑡 (3)
𝑥 𝑦 𝑧

where ℎ is hydraulic head [L], 𝑆𝑠 is specific storage that is equal to the volume of
water released from the storage per unit change in head per unit volume of porous
material (flow area × Thickness) [L-1], and 𝐾𝑥𝑥 , 𝐾𝑦𝑦 , and 𝐾𝑧𝑧 [LT] are hydraulic
conductivities along the x, y, and z axes. The term 𝑄 is the volumetric water flux
(recharge/discharge) per unit volume of porous media (sources and sinks) [T-1]
(Karatzas 2017):

𝑄 = ∑𝑟𝑖=1 𝑄𝑖 𝛿 (𝑥 − 𝑥𝑖 )(𝑦 − 𝑦𝑖 )(𝑧 − 𝑧𝑖 ) (4)

where 𝛿 is the number of sources and sinks along the x, y, and z axes, and 𝑄𝑖 is the
volumetric injection/discharge rate [L3T-1] at point (𝑥𝑖 , 𝑦𝑖 , and 𝑧𝑖 ). The value of 𝑄𝑖
in withdrawal/discharging points is positive, and for injection/recharging points is
negative.

Darcy’s velocity is the volumetric flow rate (volume per unit time) through a unit
surface area of porous media that is related to hydraulic conductivity and hydraulic
gradient, as follows:

𝜕
𝑞𝑥 = − (𝐾𝑥𝑥 𝜕ℎ) (5)
𝑥

𝜕
𝑞𝑦 = − (𝐾𝑦𝑦 𝜕ℎ ) (6)
𝑦

𝜕
𝑞𝑧 = − (𝐾𝑧𝑧 𝜕ℎ ) (7)
𝑧

𝜕 𝜕 𝜕
where − (𝜕ℎ), − (𝜕ℎ ), and − ( 𝜕ℎ) (dimensionless) are the hydraulic gradients
𝑥 𝑦 𝑧

along the x, y, and z axes.

Darcy’s velocity is not the actual velocity of groundwater through the voids. Hence,
to calculate the actual velocity (seepage velocity), Darcy’s velocity is divided by
porosity (Bear and Verruijt 2012):

𝐾𝑥𝑥 𝜕ℎ
𝑞𝑥 = − ( 𝜑 𝜕𝑥
) (8)
Groundwater modeling and its concepts, classifications, and applications for 17
solute transport simulation in saturated porous media
𝐾𝑦𝑦 𝜕ℎ
𝑞𝑦 = − ( ) (9)
𝜑 𝜕𝑦

𝐾𝑧𝑧 𝜕ℎ
𝑞𝑧 = − ( ) (10)
𝜑 𝜕𝑧

where 𝜑 is the porosity of the medium (dimensionless).

The specific storage in Equation (3) can be written as:

𝑆
𝑆𝑠 = (11)
𝑑

where 𝑆 is storage coefficient (dimensionless). Hence, Equation (3) can be


rewritten as:

𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝑆 𝑑ℎ
𝜕𝑥
(𝐾𝑥𝑥 𝜕ℎ) + 𝜕𝑦 (𝐾𝑦𝑦 𝜕ℎ ) + 𝜕𝑧 (𝐾𝑧𝑧 𝜕ℎ) − 𝑄 = 𝑑 𝑑𝑡 (12)
𝑥 𝑦 𝑧

The above equation is equivalant to:

𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝑑ℎ
𝜕𝑥
(𝑇𝑥𝑥 𝜕ℎ) + 𝜕𝑦 (𝑇𝑦𝑦 𝜕ℎ ) + 𝜕𝑧 (𝑇𝑧𝑧 𝜕ℎ) − (𝑄 ′) = 𝑆 𝑑𝑡 (13)
𝑥 𝑦 𝑧

where 𝑇 is the transmissivity (L2T-1), and 𝑄′ is the volumetric water flux per unit
area [LT-1].

Advection and dispersion are the typical mass transport mechanisms for non-
reactive solutes. The advective transport of a chemical through porous media is
given by (Bear 1988, Freeze and Cherry 1979, Scheidegger 1961):

𝐽𝐴 = 𝑞. 𝑐 (14)

where 𝐽𝐴 is the advective mass flux of chemical species [ML-2T-1], 𝑐 is the solute
concentration per unit volume [ML-3], and 𝑞 is Darcy’s velocity [LT-1].

For dispersive transport, the dispersive mass flux is considered equal to Fick’s first
law:
18 Masoud Saatsaz

𝐽𝐷 = −𝐷𝛻𝑐 (15)

where 𝐽𝐷 is the dispersive mass flux of a solute [ML-2T-1], and 𝐷 is the molecular
diffusion tensor [L2T-1]. As stated by this equation, the dispersive mass flux is a
function of the solute concentration gradient. The dispersion tensor can be indicated
based on the following components (Karatzas 2017):

𝛼𝐿 𝑞𝑥 2 +𝛼𝑇 𝑞𝑦 2 +𝛼𝑉 𝑞𝑧 2
𝐷𝑥𝑥 = |𝑞|
+𝐷 (16)

𝛼𝑇 𝑞𝑥 2 +𝛼𝐿 𝑞𝑦 2 +𝛼𝑉 𝑞𝑧 2
𝐷𝑦𝑦 = |𝑞|
+𝐷 (17)

𝛼𝑉 𝑞𝑥 2 +𝛼𝑇 𝑞𝑦 2 +𝛼𝐿 𝑞𝑧 2
𝐷𝑧𝑧 = |𝑞|
+𝐷 (18)

(𝛼𝑇− 𝛼𝑉 )(𝑞𝑥 ×𝑞𝑦 )


𝐷𝑥𝑦 = 𝐷𝑦𝑥 = |𝑞|
(19)

(𝛼𝐿− 𝛼𝑉 )(𝑞𝑦 ×𝑞𝑧 )


𝐷𝑧𝑦 = 𝐷𝑦𝑧 = |𝑞|
(20)

(𝛼𝐿− 𝛼𝑉 )(𝑞𝑥 ×𝑞𝑧 )


𝐷𝑥𝑧 = 𝐷𝑧𝑥 = |𝑞|
(21)

where 𝛼𝐿 is longitudinal dispersivity [L], 𝛼 𝑇 is transversal dispersivity [L]; 𝛼𝑉 is


vertical dispersivity [L], 𝐷𝑚 is the molecular diffusion coefficient in a saturated
porous medium [L-2T-1], and |q| is the average velocity of groundwater flow [LT-1].
The value of |q| is given by (Burnett and Frind 1987):

|𝑞| ≡ √𝑞𝑥 2 + 𝑞𝑦 2 + 𝑞𝑧 2 (22)

The convection-diffusion equation (transport equation) on the domain Ω is a partial


differential equation as follows (Babu et al. 1984):
Groundwater modeling and its concepts, classifications, and applications for 19
solute transport simulation in saturated porous media
𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕
{𝜕𝑥 (𝐷𝑥𝑥 𝜕𝐶 + 𝐷𝑥𝑦 𝜕𝐶 + 𝐷𝑥𝑧 𝜕𝐶 ) + 𝜕𝑦 (𝐷𝑦𝑦 𝜕𝐶 + 𝐷𝑦𝑥 𝜕𝐶 + 𝐷𝑦𝑧 𝜕𝐶 ) +
𝑥 𝑦 𝑧 𝑦 𝑥 𝑧
𝜕 𝜕𝐶 𝜕𝐶 𝜕𝐶 𝜕𝐶 𝜕𝐶 𝜕𝐶
𝜕𝑧
(𝐷𝑧𝑧 𝜕 + 𝐷𝑧𝑦 𝜕 + 𝐷𝑧𝑥 𝜕 ) − (𝑞𝑥 𝜕 + 𝑞𝑦 𝜕 + 𝑞𝑧 𝜕 ) + 𝑄(𝐶𝑊 − 𝐶) − 𝜑[1 +
𝑧 𝑦 𝑥 𝑥 𝑦 𝑧
𝑑𝐶
𝐸(𝐶)] 𝑑𝑡 = 0} (23)

where 𝐶 is the concentration of contaminant [ML-3] at the point (x, y, z) at time t,


𝐶𝑊 is the concentration of water entering the system [ML-3], 𝐸(𝐶) is a function
representing chemical adsorption properties.

The concentration of the withdrawn water (𝐶𝑊 ) at the discharging (pumping) point
of the groundwater system is assumed to be the same as the concentration of the
surrounding ambient water (𝐶). Thus, the term 𝑄(𝐶𝑊 − 𝐶) at all withdrawal points
becomes equal to zero. When the point is injecting (recharging), the concentration
of the injection water is considered equal to the 𝐶𝑊 component.

The conventional form of the advection-dispersion equation suffers from


conceptual shortcomings, the most known of which is that dispersivity is scale-
dependent and describes the heterogeneity. (Fetter 2018). Dispersivity is a function
of the domain size; it increases with increasing solute transport distance. Thus, the
dispersivity derived from small scale laboratory experiments may not be valid at
large scales (Willmann et al. 2008). Solute transport can be varied by the viscosity
and density differences between groundwater and leachate. Transport through a
homogeneous groundwater system can also be quite complex where horizontal
hydraulic gradient frequently changes by variation of hydrological stresses (e.g.,
extraction of groundwater by pumping wells or surface recharge), boundary
conditions, and the physical and chemical properties of the solute. In these cases,
analytical models and laboratory studied using sank tank models can improve the
assessment of the effect of changes in horizontal hydraulic gradient, and
groundwater draft and recharge on solute transport (Atlabachew et al. 2018). For
more detailed information about solute transport equations and their supporting
literature see, e.g., McDowell‐Boyer et al. (1986); Batu (2005); Woessner (2007);
Bear and Bachmat (2012); Selim (2014); Adelana (2015); Kelbaliyev et al. (2019).

Development of Solute Transport Numerical Models

The development of solute transport models in saturated porous media should


be performed in a logical sequence (Figure. 5). The fundamental steps required for
20 Masoud Saatsaz

the model development are: (i) the identification of the model objectives: (ii) data
collection: (iii) hydrogeological characterization: (iv) model conceptualization: (v)
code selection: (vi) model calibration: (vii) model verification: (viii) performing
post-processing tasks, and (ix) presentation of results (Saatsaz et al. 2013b). At first,
a conceptual model which outlines components and processes of the system should
be constructed based on available topographic, climatological, geological,
hydrogeological, hydrogeochemical, and land-use data. The next step is code
selecting, input parameters, and model design. Subsequently, calibration and
adjustment between field data (observed data) and model outputs (calculated data)
should be performed. Following the calibration, to sure about the reliability of the
model, the verification process (history matching) should be initiated. This process
analyzes the ability of the calibrated model to behave like the real system. After
completing calibration and validation, the model can be used for predicting the
future under applied scenarios. Upon performing post-processing tasks and
predictive simulations, the results of the model should be described clearly and
concisely..

Inverse Estimation of Solute Transport Properties

Proper estimation of parameter values in a solute transport model of a


groundwater system is often tricky, needs a detailed analysis of field measurements.
The task difficulty is likely because field measurements are often complicated,
time-consuming, expensive, and technique‐dependent. The hydrogeochemical
properties of an aquifer can vary due to the spatial heterogeneity and complex
chemical processes taking place in a groundwater system. Besides, field
measurements can be affected by data lacking, data quality, and data analysis error.
For instance, when multiple solutes with different diffusion coefficients are used to
assess the effects of diffusion and hydrodynamic dispersion in a porous medium,
the quality of data is so important, especially for models whose cell size is too small.
The majority of the problems discussed can be fixed by both forward and inverse
modeling techniques. Forward models mathematically optimize the system
processes based on specific parameters. Inverse models optimize the system’s
parameters based on the system’s operations. Inverse models are routinely applied
to quantify parameters which are hard, challenging, and time-consuming to measure
directly. The significant advantages of inverse modeling are (i) the precise
calculation of parameter values that yield the best fit model; (ii) evaluation of the
uncertainty of measurements, (iii) identifying data gaps, and needs; (iv) checking
the reliability of the model; and (v) documentation of matters such as model limits,
ignored by the manual calibration methods (Poeter and Hill 1997).
Groundwater modeling and its concepts, classifications, and applications for 21
solute transport simulation in saturated porous media
The practice of inverse modeling techniques along with numerical solutions for
estimating solute transport parameter values is well-acknowledged (Mishra and
Parker 1989; Abbasi et al. 2003; Giudici 2004; Vrugt 2008; Kirkham et al. 2019).
A robust approach for parameter value estimation is the nonlinear regression
method. In recent decades, the regression approach has been applied effectively to
optimize groundwater flow and solute transport models. In the regression analysis,
an objective function is applied to make an analogy between simulated values (e.g.,
hydraulic heads and solute concentrations), and observed values (prior values).
Parameter values that provide the best fit between observed and simulated data are
defined as those which yield the smallest amount of the objective function
(Tiedeman and Hill 2007). It means the goal, during regression analysis, is the
estimation of parameter values that minimize the objective function in a given error
tolerance. Preferably, a better fit is found as regression moves forward. One of the
widely used methods for minimizing the objective function is the weighted least-
squares objective function. This method reduces the objective function as follows
(Hill 2000):

𝑆 = ∑𝑁𝐷+𝑁𝑃𝑅
𝑖=1 𝜔𝑖 (𝑦𝑖 − 𝑦̂𝑖 )2 (24)

where ND is the total number of observation, NPR is the total number of prior data
equations, 𝑦𝑖 is the ith observed value (prior value), 𝑦̂𝑖 is the ith simulated value,
𝜔𝑖 is the weight for the ith observation, (𝑦𝑖 − 𝑦̂𝑖 ) is the residual for the ith observed
value, and 𝜔𝑖 (𝑦𝑖 − 𝑦̂𝑖 )2 is the square weighted residual, showing the regression
fitting as a matter of how the residuals are weighted. In most cases, weights are
determined with respect to the standard deviation of the error measured for the
difference between observation and simulation values. Further detailed information
on inverse modeling can be found in Rastogi (2007), Rastogi et al. (2010), Sun
(2013), and Bear and Cheng (2016).
22 Masoud Saatsaz

Figure 5. Flow chart classifying main steps involved in the development of a


solute transport numerical model
Groundwater modeling and its concepts, classifications, and applications for 23
solute transport simulation in saturated porous media
Classification of Solute Transport Numerical Models

Solute transport models can simulate simple, complex, or highly complex


processes. The earliest works on numerical solutions for solute transport were
simple but evidence-based in their design and performance, providing a solution to
the advection-dispersion equation for one-dimensional transport of a single,
nonreactive solute under isotropic and homogenous conditions in a steady-state
regime (Lewis et al. 1986). Later, several model designers have incorporated solute-
porous media interactions through the sources/sinks term in the governing mass-
balance equations. Consequently, mechanisms such as ion exchange, degradation,
and adsorption have been received closer attention, presenting new insights into the
simulation of single-solute transport. Since the early 1980s, numerical solution
techniques have been widely used to deal with multiple solutes involving in various
chemical reactions. At the regional scale, solute transport models can be
categorized based on components of the model, and their relations are categorized
as a transfer function, compartment, and advanced advection-dispersion models
(Orban 2009).

Transfer Function Models

Transfer function models are particularly useful when no sufficient data exist
in the structure and functioning of the system. These models are mainly controlled
by a limited number of parameters describing mass fluxes and concentration values
only at the system input and output. In such models, transfer function (also called
impulse response) is the ratio of the output of a system to the input of a system. The
relation between the input and output of the system can be expressed as follows:

𝑌(𝑡) = 𝐹(𝑡) ∗ 𝑋(𝑡) (25)

where 𝑋(𝑡) is the input function, 𝐹(𝑡) is the transfer function, and 𝑌(𝑡) is the output
(response of the system) function. A non-parametric transfer function is not
explicitly described by a function or a mathematical model. It is obtained by
deconvolution of the system response 𝑌(𝑡) by the solicitation of 𝑋(𝑡) applied to the
system (Orban 2009). Some of transfer function models are constructed based on
more parameters describing the model stresses and responses. Such models need
time series with enough history to adjust the parameters. Further information about
the application of these types of models is provided by Maloszewski (2000), Carter
et al. (2002), and Rumynin (2011).
24 Masoud Saatsaz

Advanced Solute Transport Numerical Models

The advanced solute transport models, also known as extended solute transport
models or fully coupled (hybrid) models, are applied to solve complex problems,
either one- or multidimensional, that often involve multiple chemical species,
diverse transport processes, material heterogeneity, complex boundary conditions,
and time-varying stresses. HYDRUS (2D/3D), FEHM (Finite Element Heat and
Mass Transfer), MOC (Method of Characteristics), MOFAT (Multiphase Flow and
Multicomponent Transport), MIKE SHE, MODFLOW-2000 (Modular Three-
dimensional Flow Model), NAPL Simulator, SUTRA, MODPATH (Particle-
tracking for MODFLOW), and MT3DMS (Modular Transport, 3-Dimensional,
Multi-Species model) are among the most widely used advanced solute transport
models. Table 1 provides some examples of the advanced solute transport models.

Table 1. Simulator software for groundwater flow and transport modeling

Model Features
This model simulates water flow in the saturated zone; it can be linked with
MODFLOW chemical transport modules and particle tracking modules (for flow path
simulations)
MODPATH, the model is a 3-dimensional particle tracking software for flow path and pattern
path3D simulation; the model can be integrated with MODFLOW;
MT3D simulates 3-D advection, dispersion, sorption, and reaction; it can be linked
MT3D to MODFLOW, the model solves equations based on the finite-difference method;
it is suitable for both saturated and unsaturated zones.
Family of models that couples chemical transport with biodegradation. BIOPLUME
2 focuses the aerobic reactions that occure under oxygen-limiting conditions.
BIOPLUME
BIOPLUME 3 simulates the aerobic and anaerobic reactions with multispecies
1, 2 ,3
biodegradation kinetics. The model can be used to simulate the fate and transport
of hydrocarbons and electron acceptors.
A 3-Dimensional flow and solute transport model that simulates both saturated and
3DFEMFAT unsaturated media. The solution is based on the finite element method; it is suitable
for density-dependent flow and transport.
A 2- and 3-Dimensional software for both fractured media and porous media; it is
BIOF&T suitable for simulation under heterogeneity; the model can consider advection,
dispersion, sorption, reaction, and biodegradation.
CHEMFLO is a one-dimensional vadose zone screening model; its simulation is
CHEMFLO
based on the finite difference method.
This model is a finite element model useful to solve advection, dispersion, sorption,
CHEMFLUX and decay problems; it has the capability for automatic mesh size and time step
refinement.
Groundwater modeling and its concepts, classifications, and applications for 25
solute transport simulation in saturated porous media
MOFAT MOFAT can Simulate the LNAPL movement from the surface spill to the water
table.
This software simulates the movement of air, water, and soil through saturated and
UTCHEM unsaturated zones. It can simulate multiphase fluid flow as well as formation in
vadose and saturated zones.

Continued Table 1. Simulator software for groundwater flow and transport


modeling

Model Features
A 2- and 3-D software that is used for both saturated and unsaturated media, the
FEFLOW
model is suitable for salinity and heat-dependent transport.
FLONET/ This model is a 2-D flow and solute transport-retardation model that simulates based
TRANS on the finite element method.
A 3-D groundwater flow and solute transport software that used for a single solute
HST3D
with advection, dispersion, linear sorption, and first-order decay.
A 2-D finite-difference software that simulates groundwater flow and solute
transport under the influence of advection anddispersion, along with first-order
MOC
decay. This software is also useful to simulate the reversible equilibrium-controlled
sorption and reversible equilibrium-controlled ion exchange.
A 2-D software that is practical for simulating flow and solute transport in variably
SWMS 2D saturated media with irregular boundaries under the influence of dispersion, linear
sorption, zero-order production, and first-order decay.
A 2-D finite element program that is applied for solving flow and reactive solute
VSAFT2
transport in variably saturated porous media.
A multidimensional software that is used for variable-density flow and transport
SUTRA modeling of energy or dissolved substances in a saturated and unsaturated
media

CONCLUSION
Groundwater modeling is a powerful tool for evaluating the quantity and
quality of groundwater resources over time. The comparison among different types
of groundwater models shows numerical models are often more versatile, practical,
and reliable to cope with complex solutions that rarely have simple or
straightforward solutions. In this regard, selecting model complexity is a
challenging task that controls the performance of a numerical model. The
performance of a groundwater numerical model is also dependent on so many
criteria such as the definition of the conceptual model, model structure, quality of
input data, solution techniques, characteristics of the modeled area (spatial and
temporal resolutions), operator proficiency, etc.). Comparing with flow models, the
performance of solute transport models can be affected by more factors such as the
26 Masoud Saatsaz

domain size, viscosity and density variations, changes in hydrological stresses (e.g.,
extraction of groundwater by pumping wells or surface recharge), heterogeneity of
the medium, physical and chemical properties of the solute, and complex chemical
processes taking place in the groundwater system. Although most solute transport
models provide solutions for one- or two-dimensional transport of a single,
nonreactive solute, under isotropic and homogenous conditions, in a steady-state
regime, the advanced ones can be used to solve complex problems, either one- or
multidimensional, involving multiple chemical species, diverse transport processes,
material heterogeneity, complex boundary conditions, and time-varying stresses.
However, more investigations and computations are required to increase their
robustness and to decrease uncertainty in the results.

REFERENCES
Abbasi, F., Simunek, J., Feyen, J., Van Genuchten, M. T., & Shouse, P. J.
(2003). Simultaneous inverse estimation of soil hydraulic and solute transport
parameters from transient field experiments: Homogeneous soil. Transactions of
the ASAE, 46(4), 1085.

Adelana, S. (2015). Groundwater: Hydrogeochemistry, Environmental


Impacts, and Management Practices. Nova Publishers, USA.

Alexander (1999). Bioaccumulation, bioconcentration, biomagnification.


Environmental Geology, Encyclopedia of Earth Science. pp. 43–44. doi: 10.1007/1-
4020-4494-1_31.

Alexandrov, A. (2004). Computer aided design of electrical apparatus.


Avangard Prima, Sofia, Bulgaria.

Appelo, C. A. J., & Postma, D. (2004). Geochemistry, groundwater, and


pollution. CRC press. New York, USA.

Argyros, I. K., George, S., & Thapa, N. (eds.). (2018). Mathematical modeling
for the solution of equations and systems of equations with applications. Nova
Science Publishers, USA.
Groundwater modeling and its concepts, classifications, and applications for 27
solute transport simulation in saturated porous media
Atlabachew, A., Shu, L., Wu, P., Zhang, Y., & Xu, Y. (2018). Numerical
modeling of solute transport in a sand tank physical model under varying hydraulic
gradient and hydrological stresses. Hydrogeology Journal, 26(6): 2089-2113.

Babu, D. K., Pinder, G. F., Niemi, A., Ahlfeld, D. P., & Stothoff, S. A. (1984).
Chemical Transport by Three Dimensional Ground Water Flows. Princeton
University, New Jersey, USA.

Batu, V. (2005). Applied flow and solute transport modeling in aquifers:


fundamental principles and analytical and numerical methods. CRC Press. New
York, USA.

Bear, J. (1988). Dynamics of Fluids in Porous Media, Dover, New York, USA.

Bear, J., & Verruijt, A. (2012). Modeling groundwater flow and pollution (Vol.
2). Springer Science & Business Media. Berlin, Germany.

Bear, J., & Bachmat, Y. (2012). Introduction to modeling of transport


phenomena in porous media (Vol. 4). Springer Science & Business Media,
Germany.

Bear, J., & Cheng, A. H. D. (2016). Modeling groundwater flow and


contaminant transport. Springer, New York, USA.

Bedient, P. B., Holder, A. W., Enfield, C. G., & Wood, A. L. (1999). Enhanced
remediation demonstrations at Hill air force base: introduction. Department of
Environmental Science and Engineering, Rice University, Houston, USA.

Beven, K. (2009). Environmental modeling: An uncertain future?. The


Cromwell Press. London, UK. ISBN 978-0-415-45759-0.

Bharagava, R. N. (2017). Environmental Pollutants and their Bioremediation


Approaches. CRC Press. New York, USA.

Bobba, A. G. (2012). Ground Water-Surface Water Interface (GWSWI)


modeling: recent advances and future challenges. Water resources management,
26(14): 4105-4131.
28 Masoud Saatsaz

Burnett, R. D., & Frind, E. O. (1987). Simulation of contaminant transport in


three dimensions: 2. Dimensionality effects. Water Resources Research, 23(4),
695-705.

Clark, I. (2015). Groundwater geochemistry and isotopes. CRC press. New


York, USA.

Constanda, C. (2016). Solution techniques for elementary partial differential


equations. Chapman and Hall/CRC. New York, USA.

Dassargues, A. (2018). Hydrogeology: groundwater science and engineering.


CRC Press. New York, USA.

Delleur, J. W. (2010). The handbook of groundwater engineering. CRC press.


Boca Raton, FL, USA, 992 pp.

Demirel, Y. (2013). Nonequilibrium thermodynamics: transport and rate


processes in physical, chemical and biological systems. Elsevier Science. pp: 792.
https://doi.org/10.1016/C2012-0-00459-0.

Dobrushkin, V. (2017). Applied Differential Equations with Boundary Value


Problems. Chapman and Hall/CRC. New York, USA.

Dommisse, J. P. (2015). Mathematical modeling of solute transport in a


heterogeneous aquifer. Dissertation for the degree of Master of Science. Victoria
University of Wellington, New Zealand.

Ebrahim, G. Y. (2013). Modeling groundwater systems: Understanding and


improving groundwater quantity and quality management. IHE Delft Institute for
Water Education. Delft, Netherlands.

Fetter, C. W. (2018). Applied hydrogeology. Waveland Press. Illinois, USA.

Freeze, R. A. & Cherry, J. A. (1979). Groundwater. Prentice-Hall Inc, USA.

Giudici, M. (2004). Inverse modeling for flow and transport in porous media.
Department of Earth Sciences, Geophysics Section, University of Milan, Milan,
Italy.
Groundwater modeling and its concepts, classifications, and applications for 29
solute transport simulation in saturated porous media
Gray, J., & Rumpe, B. (2016) Models in Simulation Software & Systems
Modeling. Retrieved from https://doi.org/10.1007/s10270-016-0544-y.

Hill, M. (2006). The practical use of simplicity in developing groundwater


models. Ground Water, 44(6): 775-781.

Hill, M. C. (2000). Methods and guidelines for effective model calibration. In


Joint Conference on Water Resource Engineering and Water Resources Planning
and Management, Minneapolis, Minnesota, USA. (pp. 1-10).

Howell, K. B. (2018). Ordinary Differential Equations: An Introduction to the


Fundamentals. CRC Press. New York, USA.

Insam, H., & De Bertoldi, M. (2007). Microbiology of the composting process.


In Compost Science and Technology, In LF. Diaz, M de Bertoldi, W Bidlingmaier,
E Stentiford (eds.). Waste management series (Vol. 8, pp. 25-48). Elsevier.

IUPAC. (2014). Compendium of Chemical Terminology, Gold Book Version


2.3.3. International Union of Pure and Applied Chemistry. Available at:
http://goldbook.iupac.org/PDF/D01817.pdf. Accessed 22 May 2018.

Karatzas, G. P. (2017). Developments on the modeling of groundwater flow


and contaminant transport. Water Resources Management, 31(10), 3235-3244.

Kelbaliyev, G. I., Tagiyev, D. B., & Rasulov, S. R. (2019). Transport


Phenomena in Dispersed Media. CRC Press. New York, USA.

Kirkham, J. M., Smith, C. J., Doyle, R. B., & Brown, P. H. (2019). Inverse
modeling for predicting both water and nitrate movement in a structured-clay soil
(Red Ferrosol). PeerJ, 6, e6002.

Konikow, L. (2011). The secret to successful solute transport modeling.


Ground Water 49(2): 144-159.

Konikow, L. F., & Bredehoeft, J. D. (1978). Computer Model of Two-


Dimensional Solute Transport and Dispersion in Ground Water. U.S. Geological
Survey Techniques of Water Resources Investigations.
30 Masoud Saatsaz

Konikow, L. F., Goode, D. J., & Hornberger, G. Z. (1996). A three-dimensional


method-of-characteristics solute-transport model (MOC3D). Water-Resources
Investigations Report, 96, 4267.

Kovarik, K. (2000). Numerical models in groundwater pollution. Springer


Science & Business Media, Germany.

Lage, J. L., & Narasimhan, A. (2000). Porous Media Enhanced Forced


Convection: Fundamentals and Applications. In: Vafai, K. (ed.) Handbook of
Porous Media, 8, 357-394.

Landmeyer, J. E. (2011). Introduction to phytoremediation of contaminated


groundwater: historical foundation, hydrologic control, and contaminant
remediation. Springer Science & Business Media.

Leij, F. J., & van Genuchten, M. T. (2002). Solute transport. Soil physics
companion, 189-248.

Lewis, F. M., Voss, C. I., & Rubin, J. (1986). Numerical simulation of


advective-dispersive multi-solute transport with sorption, ion exchange and
equilibrium chemistry. Water-Resources Investigations Report, 86, 4022.

Mandle, R. J. (2002). Groundwater modeling guidance. Michigan Department


of Environmental Quality, GMP, draft, 1, USA.

Mayer, A. S., & Hassanizadeh, S. M. (2005). Soil and groundwater


contamination: Nonaqueous phase liquids (No. 17). American Geophysical Union,
USA.

McDowell‐Boyer, L. M., Hunt, J. R., & Sitar, N. (1986). Particle transport


through porous media. Water Resources Research, 22(13), 1901-1921.

Merkel, B. J., Planer-Friedrich, B., & Nordstrom, D. K. (2005). Groundwater


geochemistry. A practical guide to modeling of natural and contaminated aquatic
systems, 2. Spring-Verlag Berlin Heidelberg, Germany. pp: 200.

Middlemis, H., Merrick, N., & Ross, J. (2001) Groundwater flow modeling
guideline. Murray Darling Basin Commission (MDBC) Report, Australia.
Groundwater modeling and its concepts, classifications, and applications for 31
solute transport simulation in saturated porous media
Mishra, S., & Parker, J.C. (1989). Parameter estimation for coupled unsaturated
flow and transport. Water Resources Research. 25:385–396.

Moatamedi, M., & Khawaja, H. A. (2018). Finite Element Analysis. CRC


Press. New York, USA.

Müller, F. K. Z. B. P., & Sedláčková, M. K. I. (2003). Contamination of soils


and groundwater by petroleum hydrocarbons and volatile organic compounds–Case
study: ELSLAV BRNO. Bulletin of Geosciences, 78(3), 225-239.

Neill, S. P., & Hashemi, M. R. (2018). Fundamentals of Ocean Renewable


Energy: Generating Electricity from the Sea. Academic Press. Cambridge,
Massachusetts, USA.

Ney, U., Schoberth, S. M., & Sahm, H. (1991). Anaerobic degradation of sulfite
evaporator condensate in a fixed-bed loop reactor by a defined bacterial consortium.
Applied microbiology and biotechnology, 34(6), 818-822.

Orban, P. (2009). Solute transport modeling at the groundwater body scale:


Nitrate trends assessment in the Geer basin (Belgium). Doctoral Dissertation.
University of Liège, Belgium.

Pallant, A., & Lee, H. S. (2015). Constructing scientific arguments using


evidence from dynamic computational climate models. Journal of Science
Education and Technology, 24(2-3): 378-395.

Pandey, A. K., Kumar, R., & Singh, M. K. (2018). A Solution to Advection–


Dispersion Equation for the Heterogeneous Medium Using Duhamel’s Principle.
In Amit Kumar Pandey, Rohit Kumar, Mritunjay Kumar Singh (eds.). Applications
of Fluid Dynamics (pp. 559-571). Springer, Singapore.

Payne, S. M., & Woessner, W. W. (2010). An Aquifer Classification System


and Geographical Information System‐Based Analysis Tool for Watershed
Managers in the Western US 1. JAWRA Journal of the American Water Resources
Association, 46(5): 1003-1023.
32 Masoud Saatsaz

Peaceman, D. W. & Rachford Jr., H. H. (1955). Numerical solution of the


parabolic and elliptic differential equations. Journal of the Society for Industrial
and Applied Mathematics. 3: 28-41.

Pinder, G. F., & Bredehoeft, J. D. (1968) Application of the digital computer


for aquifer evaluation. Water Resource Research, 4: 1069–1093.
doi:10.1029/WR004i005p01069.

Poeter, E.P., & Hill, M.C. (1997). Inverse modeling: A necessary next step in
groundwater modeling. Ground Water: 35(2): 250-260.

Powell, A. L. (2017). Water Resource Planning, Development and


Management. Nova Publishers, USA.

Prickett, T. A. & Lonnquist, C. G. (1971). A selected digital computer for


groundwater resource evaluation. Illinois State Water Survey Bull., 55: 62.

Prior, J. M. (2018). Contemporary Mission Theology: Engaging the Nations


(Essays in Honor of Charles E. Van Engen). Robert Gallagher & Paul Hertig (eds)
American Society of Missiology Series No. 53. Maryknoll, New York, Orbis
Books, USA.

Raats, P. A., & Knight, J. H. (2018). The Contributions of Lewis Fry


Richardson to Drainage Theory, Soil Physics, and the Soil-Plant-Atmosphere
Continuum. Frontiers in Environmental Science, 6, 13.

Rapp, B. E. (2016). Microfluidics: Modeling, Mechanics, and Mathematics.


William Andrew. Norwich, NY, USA.

Rastogi, A. K. (2007). Numerical groundwater hydrology. Penram


International Publication, India. pp: 981.

Rastogi, A. K., Aparna, B., & Sanjeeta, K. (2010). Modeling and Inverse
modeling in Groundwater Hydrology: Application of FEM and soft computing
optimization. Omniscriptum Publishing Group, Germany. pp: 412.

Renard, P. (2005). The future of hydraulic tests. Hydrogeology Journal, 13:


259–262.
Groundwater modeling and its concepts, classifications, and applications for 33
solute transport simulation in saturated porous media
Richens, D. T. (1997). The chemistry of aqua ions: synthesis, structure, and
reactivity: a tour through the periodic table of the elements. Wiley. ISBN 0-471-
97058-1.

Rivett, M., Drewes, J., Barrett, M., Chilton, J., Appleyard, S., Dieter, H. H.,
Wauchope, D. & Fastner, J., (2006). Chemicals: Health relevance, transport, and
attenuation, in protecting groundwater for health managing the quality of drinking-
water source. In Schmon, O., Howard, G., Chilton, J., Chorus, I., & World Health
Organization (eds.). Protecting groundwater for health: Managing the quality of
drinking-water sources. pp: 81-137. Recuperado de http://www. WHO.
int/iris/handle/10665/43186.

Rubin, H., Narkis, N., & Carberry, J. (Eds.). (2013). Soil and aquifer pollution:
non-aqueous phase liquids-contamination and reclamation. Springer Science &
Business Media.

Saatsaz, M., Sulaiman, W. N. A., Eslamian, S., & Mohammadi, K. (2013)a.


Hydrogeochemistry and groundwater quality assessment of Astaneh-Kouchesfahan
Plain, Northern Iran. International journal of water, 7(1-2), 44-65.

Saatsaz, M., Sulaiman, W. N. A., Eslamian, S., & Javadi, S. (2013)b.


Development of a coupled flow and solute transport modeling for Astaneh-
Kouchesfahan groundwater resources, North of Iran. International journal of water,
7(1-2), 80-103.

Scesi, L. T. G., & Gattinoni, P. (2012). Methods and Models to Determine the
Groundwater Flow in Rock Masses: Review and Examples.

Scheidegger, A. E. (1961). On the statistical properties of some transport


equations. Canadian Journal of Physics, 39(11), 1573-1582.

Schwartz, F. W., & Zhang, H. (2003). Fundamentals of Groundwater. John


Wiley & Sons. New York, USA. pp: 583.

Schwarz, C. V., Reiser, B. J., Davis, E. A., Kenyon, L., Acher, A., Fortus, D.,
Shwartz, Y., Hug, B., & Krajcik, J. (2009). Developing a learning progression for
scientific modeling: Making scientific modeling accessible and meaningful for
learners. Journal of Research in Science Teaching, 46(6): 632–654.
34 Masoud Saatsaz

Schweizer, B. (2015). Darcy's law and groundwater flow modeling Snapshots


of modern. Mathematics from Oberwolfach, no 7, Germany.

Scozzari, A., & Dotsika, E. (2016). Threats to the Quality of Groundwater


Resources. Springer Berlin Heidelberg, Germany.

Selim, H. M. (2014). Transport & fate of chemicals in soils: principles &


applications. CRC Press. New York, USA.

Simmons, C. T. (2008). Henry Darcy (1803–1858): Immortalized by his


scientific legacy. Hydrogeology Journal, 16(6), 1023. doi:10.1007/s10040-008-
0304-3.

Stone, H. L. (1968). Iterative Solution of Implicit Approximations of


Multidimensional Partial Differential Equations. SIAM Journal on Numerical
Analysis. 5 (3): 530-538. doi:10.1137/0705044.

Stumm, W., & Morgan, J. J. (1996). Aquatic Chemistry: Chemical Equilibria


and Rates in Natural Waters. Environmental Science and Technology. John Wiley
& Sons. New York, USA.

Sun, N. Z. (2013). Inverse problems in groundwater modeling (Vol. 6).


Springer Science & Business Media, Germany.

Thangarajan, M., (ed.). (2007). Groundwater: Resource evaluation,


augmentation, contamination, restoration, modeling, and management. Springer
Science & Business Media.

Tiedeman, C. R., & Hill, M. C. (2007). Model calibration and issues related to
validation, sensitivity analysis, post-audit, uncertainty evaluation and assessment
of prediction data needs. In Groundwater (pp. 237-282). Springer, Dordrecht,
Germany.

Tikhomirov, V. V. (2016). Hydrogeochemistry fundamentals and advances:


groundwater composition and chemistry, volume 1 for designers, owners and
operators. John Wiley & Sons Inc; ISBN-10: 1119160391.
Groundwater modeling and its concepts, classifications, and applications for 35
solute transport simulation in saturated porous media
Turner, M. J., Clogh, R. W., Martin, H. C., & Topp, L. J. (1956). Stiffness and
deflection analysis of complex structures. Journal of Aeronautical Science,
23(9):805–823, 854.

Vrugt, J. A., Stauffer, P. H., Wöhling, T., Robinson, B. A., & Vesselinov, V.
V. (2008). Inverse modeling of subsurface flow and transport properties: A review
with new developments. Vadose Zone Journal, 7(2), 843-864.

Wang, D., Wu, C., Huang, W., & Zhang, Y. (2019). Vibration investigation on
fluid-structure interaction of AP1000 shield building subjected to multi earthquake
excitations. Annals of Nuclear Energy, 126: 312-329.

Willmann, M. Carrera, J. Sanchez-Vila, X. Willmann, M., Carrera, J., &


Sánchez‐Vila, X. (2008). Transport upscaling in heterogeneous aquifers: What
physical parameters control memory functions?. Water Resources Research,
44(12).

Woessner, W. W. (2007). Applied Flow and Solute Transport Modeling in


Aquifers: Fundamental Principles and Analytical and Numerical Methods. CRC
Press Book. New York, USA.

Young, D. M. (1971). Iterative Solution of Large Linear Systems. Academic


Press, New York, USA.

Zienkiewicz, O.C. (1971). The Finite Element Method in Engineering Science.


McGraw-Hill, London, UK.

View publication stats

You might also like