Saatsaz
Saatsaz
net/publication/344459959
CITATIONS READS
0 1,951
2 authors:
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
All content following this page was uploaded by Masoud Saatsaz on 17 August 2021.
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,
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).
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 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.
Numerical Models
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
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.
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).
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
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).
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)
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):
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
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
𝑆
𝑆𝑠 = (11)
𝑑
𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝑆 𝑑ℎ
𝜕𝑥
(𝐾𝑥𝑥 𝜕ℎ) + 𝜕𝑦 (𝐾𝑦𝑦 𝜕ℎ ) + 𝜕𝑧 (𝐾𝑧𝑧 𝜕ℎ) − 𝑄 = 𝑑 𝑑𝑡 (12)
𝑥 𝑦 𝑧
𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 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)
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 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..
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
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:
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
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.
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.
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.
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.
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.
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.
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.
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.
Leij, F. J., & van Genuchten, M. T. (2002). Solute transport. Soil physics
companion, 189-248.
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.
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.
Poeter, E.P., & Hill, M.C. (1997). Inverse modeling: A necessary next step in
groundwater modeling. Ground Water: 35(2): 250-260.
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.
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.
Scesi, L. T. G., & Gattinoni, P. (2012). Methods and Models to Determine the
Groundwater Flow in Rock Masses: Review and Examples.
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
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.
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.