Groundwater pollution may occur due to human activities, industrial effluents, cemeteries, mine spoils, etc. This paper deals with one-dimensional mathematical modeling of solute transport in finite aquifers. The governing equation for solute transport by unsteady groundwater flow is solved analytically by the Laplace transform technique. Initially, the aquifer is subjected to the spatially dependent source concentration with zero-order production. One end of the aquifer receives the source concentration and is represented by a mixed-type boundary condition in the splitting time domain. The concentration gradient at the other end of the porous media is assumed to be zero. The temporally dependent velocity and the dispersion coefficients are considered. A numerical solution is obtained by using an explicit finite difference scheme and compared with the analytical result. Accuracy of the solution is discussed by using the root mean square error method. Truncation error is also explored for the parameters like numerical dispersion and velocity terms. The impact of Peclet number is examined. For graphical interpretation, unsteady velocity expressions (i.e., such as exponential, sinusoidal, asymptotic, and algebraic sigmoid) are considered. The work may be used as a preliminary predictive tool for groundwater resource and management.
INTRODUCTION
In India, many aquifers are being contaminated by a host of human activities, such as sewage disposal, refuse dumps, pesticides and chemical fertilizer contamination, industrial effluent discharges, and toxic waste disposal (Rausch et al. 2005; Batu 2006). The traditional advection dispersion equation is based on the conservation of mass and Fick's law of diffusion (Fried & Combarnous 1971; Bear & Verruijt 1987; Chrysikopolous et al. 1990) and constitutes the basis of solute transport models that are used for predicting the movement of contaminants in groundwater systems.
There has been some research on solute transport in groundwater systems. Hunt (1978) analyzed one-, two-, and three-dimensional solutions for instantaneous, continuous, and steady state pollution sources in uniform groundwater flow. Freeze & Cherry (1979) provided a relation between dispersion and groundwater velocity in which the dispersion is proportional to a power of the velocity and experimentally observed that the power ranged between 1 and 2. van Genuchten (1981) explored derivations of analytical solutions using the Laplace transform for the solute transport equation with zero-order production and/or first-order decay subjected to first and third type boundary conditions. Zoppou & Knight (1994) evaluated analytical solutions that are still useful for validating numerical schemes for solving the advection–diffusion equation with spatially variable coefficients. Logan (1996) obtained analytical solutions for a scale-dependent dispersion coefficient increasing exponentially with position up to some constant limiting values. Hantush & Marino (1998) developed analytical solutions using the Laplace and Fourier transform methods and superposition principle for the first-order rate model in an infinite porous medium.
Using Peclet and Courant numbers and a new sink/source dimensionless number, Ataie-Ashtiani et al. (1999) discussed truncation errors associated with finite difference solutions of the advection–dispersion equation (ADE) with first-order reaction. Bedient et al. (1999) presented a mathematical model of the ADE for describing the migration and fate of pollutants in groundwater. Neville et al. (2000) presented semi-analytical solutions for a multi-process non-equilibrium model for describing contaminant concentration distribution patterns. Balla et al. (2002) presented a computational case study using a transport model for pollution of underground water due to damage of the waterproofing system in a waste material depository or sewage sludge composting plant.
Lowry & Li (2005) discussed an improved finite analytical solution method for solving the time-dependent ADE that does not discretize the derivative terms rather solving the equation analytically in the space–time domain. Smedt (2006) presented an analytical solution for solute transport in rivers, including the effect of transient storage and first-order decay. Tkalich (2006) explored the derivation of high-order advection–diffusion schemes by employing the interpolation polynomial method. Hill et al. (2007) proposed upscaling models of solute transport in porous media through genetic programming in heterogeneous porous media. Chen (2007) presented an analytical solution of two-dimensional non-axisymmetrical solute transport in a radially convergent flow tracer test with a diffusion coefficient increasing with travel distance. Yeh & Yeh (2007) derived solutions of the transport equation with a point-source term considered as the point-source solution under the condition that the solute was introduced into the flow system from the boundary that was considered as the boundary-source solution. Kumar et al. (2008) also described transport through a heterogeneous porous medium with a time-dependent dispersivity in solute transport modeling. Zhan et al. (2009) explored two-dimensional solute transport in an aquifer–aquitard system by maintaining the mass conservation at the aquifer–aquitard interface. Gao et al. (2012) explored a mobile–immobile model with an asymptotic dispersivity function of travel distance with the concept of scale-dependent dispersion during solute transport in finite heterogeneous porous media. Rezaei et al. (2013) derived a semi-analytical solution to the two-dimensional conservative solute transport in an aquifer bounded by thin aquitards in the Laplace domain. Singh & Das (2015) explored the analytical and numerical solutions for one-dimensional scale-dependent solute transport in heterogeneous media in which analytical and numerical solutions were compared and found very good agreement among them. The root mean square (RMS) error analysis was made to check the accuracy of the solution.
In the present work, we focus on one-dimensional solute transport modeling using the ADE in a finite aquifer with first-order decay and zero-order production. To simplify ADE, different transformations were applied. Non-dimensional parameters were employed for reducing the number of parameters of the ADE. To predict the pattern of contaminant concentration, different types of unsteady velocities, such as sinusoidal, exponentially decreasing, asymptotic, and algebraic sigmoid, were considered. They helped describe the nature of contaminant concentration in time and space.
MATHEMATICAL FORMULATION
NUMERICAL SOLUTION
STABILITY CONDITION
The difference approximation equation was stable if the eigenvalues of A had modulus values less than or equal to unity, i.e., , where was the eigenvalue of matrix A.
TRUNCATION ERROR
Numerical dispersion was first quantified by Lantz (1971). Ataie-Ashtiani et al. (1999) explored the expansion of the Taylor series of solute concentration along the ADE used for determining the truncation error in one dimension. Chaudhari (1971) investigated a second-order error through the examination of the truncated Taylor series approximation with explicit finite difference solution of the one-dimensional ADE. We also explored the truncation error for the various parameters, such as dispersion, seepage velocity, first-order decay, and zero-order production term.
Now, after comparison between Equation (52) and the original partial differential equation, we found different forms of the truncation error, as discussed by Ataie-Ashtiani et al. (1996). These errors can be identified as follows:
ACCURACY OF THE SOLUTION
RESULTS AND DISCUSSION
In this paper, the RMS error was used to check the validity of numerical solution against the analytical one, as shown in Tables 1 and 2. The two parameters, and , play an important role to investigate the performance of the numerical solution. In the explicit finite difference scheme is restricted under the stability condition. Thus, in this present study, the accuracy was investigated by selecting different mesh sizes. The RMS error was investigated for , 0.05, and 0.07 for the particular time period 20 years in the time domain for the exponential decreasing and the asymptotic form of the velocity patterns, and 30 years with within the sinusoidal and algebraic sigmoid form of the velocity patterns in sand and clay media, respectively.
Distance . | Analytical result . | Numerical result . | ||
---|---|---|---|---|
ΔZ = 0.02 . | ΔZ = 0.05 . | ΔZ = 0.07 . | ||
Case i: For exponential decreasing form of the velocity pattern | ||||
0.0019 | 0.5125 | 0.0119 | 0.0149 | 0.0169 |
0.0055 | 0.3188 | 0.0159 | 0.0248 | 0.0308 |
0.0093 | 0.1736 | 0.0198 | 0.0348 | 0.0448 |
0.0130 | 0.0837 | 0.0238 | 0.0448 | 0.0588 |
0.0167 | 0.0376 | 0.0278 | 0.0548 | 0.0727 |
RMS error | 0.2719 | 0.2665 | 0.2634 | |
Case ii: For asymptotic form of the velocity pattern | ||||
0.0019 | 0.3613 | 0.0120 | 0.0150 | 0.0170 |
0.0055 | 0.0687 | 0.0160 | 0.0250 | 0.0310 |
0.0093 | 0.0130 | 0.0200 | 0.0350 | 0.0449 |
0.0130 | 0.0094 | 0.0240 | 0.0449 | 0.0589 |
0.0167 | 0.0091 | 0.0280 | 0.0549 | 0.0729 |
RMS error | 0.1583 | 0.1585 | 0.1596 |
Distance . | Analytical result . | Numerical result . | ||
---|---|---|---|---|
ΔZ = 0.02 . | ΔZ = 0.05 . | ΔZ = 0.07 . | ||
Case i: For exponential decreasing form of the velocity pattern | ||||
0.0019 | 0.5125 | 0.0119 | 0.0149 | 0.0169 |
0.0055 | 0.3188 | 0.0159 | 0.0248 | 0.0308 |
0.0093 | 0.1736 | 0.0198 | 0.0348 | 0.0448 |
0.0130 | 0.0837 | 0.0238 | 0.0448 | 0.0588 |
0.0167 | 0.0376 | 0.0278 | 0.0548 | 0.0727 |
RMS error | 0.2719 | 0.2665 | 0.2634 | |
Case ii: For asymptotic form of the velocity pattern | ||||
0.0019 | 0.3613 | 0.0120 | 0.0150 | 0.0170 |
0.0055 | 0.0687 | 0.0160 | 0.0250 | 0.0310 |
0.0093 | 0.0130 | 0.0200 | 0.0350 | 0.0449 |
0.0130 | 0.0094 | 0.0240 | 0.0449 | 0.0589 |
0.0167 | 0.0091 | 0.0280 | 0.0549 | 0.0729 |
RMS error | 0.1583 | 0.1585 | 0.1596 |
Distance . | Analytical result . | Numerical result . | ||
---|---|---|---|---|
ΔZ = 0.02 . | ΔZ = 0.05 . | ΔZ = 0.07 . | ||
Case i: For the sinusoidal form of the velocity pattern | ||||
0.0028 | 0.1094 | 0.0118 | 0.0148 | 0.0168 |
0.0083 | 0.2425 | 0.0158 | 0.0248 | 0.0307 |
0.0138 | 0.2524 | 0.0198 | 0.0347 | 0.0447 |
0.0193 | 0.1869 | 0.0238 | 0.0447 | 0.0586 |
0.0248 | 0.1135 | 0.0278 | 0.0546 | 0.0725 |
RMS error | 0.1726 | 0.1596 | 0.1514 | |
Case ii: For the algebraic sigmoid form of the velocity pattern | ||||
0.0028 | 0.1001 | 0.0119 | 0.0149 | 0.0169 |
0.0083 | 0.1608 | 0.0159 | 0.0249 | 0.0309 |
0.0138 | 0.0868 | 0.0199 | 0.0349 | 0.0449 |
0.0193 | 0.0301 | 0.0239 | 0.0449 | 0.0589 |
0.0248 | 0.0126 | 0.0279 | 0.0549 | 0.0728 |
RMS error | 0.0818 | 0.0780 | 0.0774 |
Distance . | Analytical result . | Numerical result . | ||
---|---|---|---|---|
ΔZ = 0.02 . | ΔZ = 0.05 . | ΔZ = 0.07 . | ||
Case i: For the sinusoidal form of the velocity pattern | ||||
0.0028 | 0.1094 | 0.0118 | 0.0148 | 0.0168 |
0.0083 | 0.2425 | 0.0158 | 0.0248 | 0.0307 |
0.0138 | 0.2524 | 0.0198 | 0.0347 | 0.0447 |
0.0193 | 0.1869 | 0.0238 | 0.0447 | 0.0586 |
0.0248 | 0.1135 | 0.0278 | 0.0546 | 0.0725 |
RMS error | 0.1726 | 0.1596 | 0.1514 | |
Case ii: For the algebraic sigmoid form of the velocity pattern | ||||
0.0028 | 0.1001 | 0.0119 | 0.0149 | 0.0169 |
0.0083 | 0.1608 | 0.0159 | 0.0249 | 0.0309 |
0.0138 | 0.0868 | 0.0199 | 0.0349 | 0.0449 |
0.0193 | 0.0301 | 0.0239 | 0.0449 | 0.0589 |
0.0248 | 0.0126 | 0.0279 | 0.0549 | 0.0728 |
RMS error | 0.0818 | 0.0780 | 0.0774 |
In both the tables, was fixed. Tables 1 and 2 were tabulated for the RMS error in the aquifer (i.e., sand) and aquitard (i.e., clay) for four different types of the velocity patterns, respectively. The RMS error decreases with the increasing grid space for the exponential decreasing form of the velocity pattern for the sand medium, which was observed from Case i of Table 1. In the asymptotic form of the velocity pattern the RMS error increases with respect to the increasing grid space in Case ii of Table 1. The RMS error attains its minimum value with the increasing mesh size for Cases i and ii in Table 2 for the clay medium. The RMS error was evaluated for the accuracy of solution for the sinusoidal form of the velocity pattern tabulated in Case i in Table 2, and for the algebraic sigmoid form of the velocity pattern tabulated in Case ii of Table 2. In both the velocity patterns the result is more accurate for the maximum value of the mesh size, except in the case of the asymptotic form of the velocity pattern where the result is more accurate in the case of minimum value of the mesh size.
VALIDATION OF THE MODEL WITH EXISTING SOLUTION OF LIU ET AL. (1998)
SUMMARY AND CONCLUSIONS
Employing the concept of linear isotherms, analytical solutions for the ADE with respect to the solid and liquid phase are derived. The mixed type boundary condition is employed at the source in the splitting time domain. The contaminant concentration patterns for different types of velocity patterns are evaluated. The following conclusions can be drawn from this study:
The impact of contaminant concentration for linear isotherms with the distribution coefficient is significantly observed in the splitting time domain for different velocity patterns, such as exponentially decreasing, sinusoidally varying, algebraic, and asymptotic forms.
The contaminant concentration values depend upon the decreasing or increasing values of the zero-order production term and first-order decay rate coefficient.
The contaminant concentration distribution behavior is predicted for different geological formations in two time domains, i.e., and .
Comparison of the analytical result with the numerical result is taken into account. Accuracy of the solution is significantly observed by using RMS error. Truncation error of various parameters is also explored, which causes the inconsistency among analytical and numerical results.
The validation of the model is made with the result of an existing solution given by Liu et al. (1998) and the same trend for contaminant concentration was found.
ACKNOWLEDGEMENTS
The authors are grateful to the Indian School of Mines, Dhanbad for providing financial support to PhD candidate under the ISMJRF scheme. The authors are also grateful to the editor and reviewers for the comments which helped improve the quality of the paper.