Next Article in Journal
Visual Satisfaction of Urban Park Waterfront Environment and Its Landscape Element Characteristics
Previous Article in Journal
Analysis of Precipitation Change and Its Influencing Factors Around the Lop Nor Salt Flat
Previous Article in Special Issue
Numerical Simulation of Fine Particle Migration in Loose Soil Under Groundwater Seepage Based on Computational Fluid Dynamics–Discrete Element Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Early-Time Recession Solution from a Steady-State Initial Condition for the Horizontal Unconfined Aquifer

by
Elias Gravanis
1,
Evangelos Akylas
1 and
Ernestos N. Sarris
2,3,*
1
Department of Civil Engineering and Geomatics, Cyprus University of Technology, 3036 Limassol, Cyprus
2
Department of Mineralogy-Petrology-Economic Geology, Faculty of Sciences, School of Geology, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
3
Oil and Gas Program, Department of Engineering, University of Nicosia, 1700 Nicosia, Cyprus
*
Author to whom correspondence should be addressed.
Water 2025, 17(5), 771; https://doi.org/10.3390/w17050771
Submission received: 7 February 2025 / Revised: 1 March 2025 / Accepted: 5 March 2025 / Published: 6 March 2025

Abstract

:
In this work, we present the semi-analytical solution for the early-time recession phase of the horizontal unconfined aquifer of finite length for steady-state initial conditions. This is a case where self-similarity arguments are not applicable. The solution is built as linear perturbations from the initial steady state. The solution is determined via a Sturm–Liouville eigenvalue problem, which should be solved numerically. On the other hand, the immediate response of the aquifer to the sudden switching off the recharge, i.e., in the earliest times, is obtained by deducing analytically the large eigenvalue asymptotic solutions of the problem. We find analytically that in this time regime, the outflow Q is given by Q = Q0 − 1.4Q05/3L−4/3n−2/3t2/3, where Q0 is the initial outflow rate, L is the length of the aquifer, n is the porosity of the formation and t is the time from the start of the recession. The stated result is very accurate for times t up to ~0.01 nL3/2k−1/2Q0−1/2, where k is the hydraulic conductivity of the formation. The analytical and quantitative relation of the presented solution with the classical recession phase asymptotic solutions derived in the past by Polubarinova (early-time solution) and Boussinesq (separable, late-time solution) is discussed in detail. The presented results can be used as a benchmark solution for modeling or numerical validation purposes.

1. Introduction

The Boussinesq equation is a groundwater flow dynamical equation that models one-dimensional water flows in homogeneous unconfined aquifers over horizontal or uniformly sloping beds [1,2]. The Boussinesq equation is an expression of local conservation of mass (the continuity equation) for saturated flow which follows from the Dupuit–Forchheimer assumptions [3], that the groundwater flows primarily along the flat bed, combined with Darcy’s law. The Dupuit–Forchheimer assumptions imply that the Darcy flux through any cross section of the flow is proportional to the flow depth. This implies that the Boussinesq equation is non-linear (quadratic) with respect to the water depth.
The assumptions about the geometric and hydraulic properties of the flow imply that the Boussinesq equation is a model that describes very approximately the groundwater flow of real aquifers. That means that, as is most usually the case with engineering applications, its primary usefulness rests on simple and clear results that are strongly based on the physical content of the model rather than on complex solutions of the corresponding equation. Such results usually take the form of scaling relations and asymptotic solutions.
For the problem at hand, the recession phase of the unconfined aquifer, there are two classic asymptotic solutions: one derived by Polubarinova for the semi-infinite aquifer with constant boundary depths [4] by imposing self-similarity, and another derived by [5] by imposing separability (which is a special form of self-similarity) on a finite length aquifer. These solutions are the early- and late-time asymptotic solutions of the problem, respectively. The Polubarinova solution also applies to finite aquifers with vanishing outlet depth and corresponds to a horizontal initial profile, see, e.g., [6,7], as we explicitly discuss in this work. Pertinent use of both asymptotic solutions has been made in application-oriented works such as [8].
Self-similar solutions for the recession phase have been constructed over the decades by many authors; see, for example, the discussion in [9]. Barenblatt in [10] derived a power series solution for a self-similar Boussinesq equation assuming a power law time-dependence of the water depth at one end of a semi-infinite horizontal aquifer. Both the Polubarinova and the Barenblatt boundary conditions still generate a certain amount of literature, see, e.g., [11,12,13,14,15,16,17,18,19,20,21,22]. It should be noted that, in contrast, the literature on the early-time solution of the build-up phase is limited; see [23,24] for the exact and approximate solutions to this problem, respectively.
When the initial state of a finite length aquifer is a smooth profile, for example, it may be the steady-state profile of a previous constant recharge period, the Polubarinova problem which leads to a singular outflow rate Q proportional to 1/√t, with t being the recession duration, is no longer the correct response to the sudden change in recharge rate. Boussinesq equation implies that, whenever the depth squared profile is finite at the outlet, there is no reason for Q to start with an (integrable) infinity; Q is perfectly continuous through the sudden change. On the other hand, although the general smooth initial condition is consistent with the outlet boundary condition at start of recession t = 0, it is not consistent dynamically. As we shall see, that requires a vanishing second derivative at the outlet for the depth squared profile. That means that, dQ/dt must exhibit an integrable singularity at t = 0 for the smooth initial conditions. The main objective of this paper is to deduce in the most detailed manner the law of the aquifer response in this problem. For definiteness we focus on the steady-state initial condition, although the formulation is easily generalizable to other initial profiles.
This paper is organized as follows. In Section 2, we set the notation and summarize certain facts necessary for the rest of the work. In Section 3, we develop the analysis of this work. We start by formulating the specific problem one needs to solve, i.e., the dynamics of the perturbations of the initial state, which is then solved both numerically, as well as analytically through asymptotic methods. In Section 4, we summarize our findings and present concluding remarks.

2. The Boussinesq Equation

We assume uniform and constant hydraulic conductivity k and porosity n throughout the aquifer. The flow will be considered saturated and effectively one-dimensional, and the recharge rate r wherever it appears is regarded as constant. The impermeable bottom of the aquifer is assumed flat and horizontal. The aquifer is assumed to be of finite length L.
The water table height h is governed by the Boussinesq equation.
n h t k x h h x = r
The boundary conditions that we will impose are:
h x ( 0 , t ) = 0 , h ( L , t ) = 0
which expresses that there is no inflow at the point x = 0 and that the water table has zero depth at the outlet of the aquifer.
Let [x] and [h] be the scales of the extent and height of the water table, and [t] a time scale. For r = 0, Equation (1) implies that
k [ h ] 2 [ x ] 2 = n [ h ] [ t ]
We choose [x] = L and for any scales [h] and [t] consistent with (3) we define dimensionless variables X, T, H by
x = L X , h = [ h ] H , t = [ t ] T
Clearly, 0 ≤ X ≤ 1. Then, Equation (1), for zero recharge rate reads
H T = 1 2 2 H 2 X 2
while the boundary conditions take the form
H X ( 0 , T ) = 0 , H ( 1 , T ) = 0
In these variables, the storage and outflow rate read, respectively
S = 0 1 H ( X , T ) d X , Q = 1 2 ( H 2 ) ( 1 , T )

3. Early Time Asymptotic from an Initial Profile with a Continuous Slope

3.1. General Mathematical Analysis

Consider an initial profile H0(X). We assume that this function has a continuous slope and obeys the boundary conditions (6). The slope may become minus infinity at the outlet X = 1 to be consistent with the boundary condition the zero-height boundary condition and a finite or infinite outflow rate Q, given by (7).
By Equation (7), Q is finite if the slope of the squared profile is finite at X = 1. The Polubarinova early-time recession solution [4], reviewed in some detail in Appendix A, is a solution consistent with a uniform initial profile for all X < 1. This means that the square profile has an effectively infinite slope for this solution. This is why Q starts with a singularity in time: Q is proportional to 1/√t.
If the initial profile is, for example, a steady state, or any profile with finite Q before a recession starts, there is no such initial singularity in the flow rate. On the other hand, Equations (5) and (6) say that the outlet boundary condition implies dynamically that the square profile must have a vanishing second derivative at X = 1. This is certainly not a feature of the steady state, or of any generic initial profile with a finite Q. That means that the Boussinesq equation will transform the initial profile into something with this property through a sudden but finite change in ∂H/∂T. That means the second, time derivative of the profile must be infinite for T = 0+. That in turn means that dQ/dT will be infinite for T = 0+. We show here that this is realized by a (smaller than 1) power law departure from the initial value of the outflow rate.
Consider small perturbations δH(Χ,Τ), that is,
S = 0 1 H ( X , T ) d X , Q = 1 2 ( H 2 ) ( 1 , T )
Given that the initial profile obeys the boundary conditions, the perturbation must also obey them:
δ H X ( 0 , T ) = 0 , δ H ( 1 , T ) = 0
with initial condition δH (Χ,0) = 0.
The perturbation is governed by the linear equation
δ H T = 2 ( H 0 δ H ) X 2 + 1 2 2 H 0 2 X 2
which is obtained by substituting (8) in (5) and dropping the quadratic term. The second term on the right hand side does not depend on the field δH, and hence acts as a source/forcing term.
Aiming at an expansion in orthogonal functions, we look for solutions of the form
δ H ( X , T ) = e κ 2 T φ κ ( X )
where κ is a constant, to be determined by the boundary conditions. Substituting into (10) without the source term, we obtain
( H 0 ( X ) φ κ ( X ) ) = κ 2 φ k ( X )
This is an eigenvalue equation, where the eigenfunctions φκ(Χ) obey the boundary conditions
φ κ ( 0 ) = 0 ,       φ κ ( 1 ) = 0
Define the functions Φκ(Χ) = H0(Χ) φκ(Χ). These functions clearly obey the boundary conditions (13) and the equation
Φ κ ( Χ ) = κ 2 1 H 0 ( X ) Φ κ ( X )
Due to the linearity of the eigenvalue Equation (14) the normalization of Φκ(Χ) is left unspecified. We impose the simple condition Φκ(0) = 1.
Equation (14) defines a Sturm–Liouville eigenvalue problem with weight 1/H0(X): The solution of (14) with the given boundary conditions leads to a discrete set of functions Φκ(Χ), labeled by a discrete set of constant rates of decay (eigenvalues) κ2, which form a complete and orthogonal set of functions in space of weighted square integrable functions, which obey the given boundary conditions. The orthogonality relation explicitly reads
Φ κ ( Χ ) = κ 2 1 H 0 ( X ) Φ κ ( X )
for any two values of eigenvalues κ and κ′. In terms of the original functions the orthogonality reads
0 1 φ κ ( X ) φ κ ( X ) H 0 ( X ) d X = N κ δ κ κ , N κ = 0 1 φ κ 2 ( X ) H 0 ( X ) d X
Expand now the general solution of (10) in these orthogonal functions
δ H ( X , T ) = κ c κ ( T ) φ κ ( X )
where the function cκ(Τ) are determined by (10) and the initial condition δH(Χ,0) = 0. This sum is inverted by the orthogonality relation (16):
c κ ( T ) = N κ 1 0 1 δ H ( X , T ) φ κ ( X ) H 0 ( X ) d X
Multiplying (10) with H0(X)φκ(Χ) and integrating over the interval 0 < X < 1, using also the orthogonality relation (16) and Equation (18), we obtain a dynamical equation for cκ(Τ):
c κ ( T ) = κ 2 c κ ( T ) d κ
where dκ are given by
d κ = 1 2 N κ 1 0 1 H 0 ( H 0 2 ) φ κ ( X ) d X = 1 2 N κ 1 0 1 ( H 0 2 ) Φ κ ( X ) d X = constant
with the initial condition cκ(0) = 0. We immediately obtain
c κ ( T ) = d κ κ 2 ( 1 e κ 2 T )
Hence, as long as one is able to solve (14), the solution of the problem is
δ H ( X , T ) = κ d κ κ 2 ( 1 e κ 2 T ) φ κ ( X )
The storage associated with this solution reads
S ( T ) = 0 1 H 0 ( X ) d X + 0 1 δ H ( X , T ) d X = = 0 1 H 0 ( X ) d X κ d κ κ 2 ( 1 e κ 2 T ) e κ , e κ = 0 1 φ κ ( X ) d X = 0 1 Φ κ ( X ) d X H 0 ( X )
Therefore, the outflow rate reads
Q ( T ) = d S ( T ) d T = κ d κ e κ e κ 2 T
Let us now focus on a specific initial condition of this kind, i.e., the steady-state solution. This result is further analyzed below.
Regarding the range of times where the solution (35) is expected to be a reasonably good approximation, one may note that the solution holds well as long as the time-dependent exponentials are small. That is, it is expected to be a good approximation for times t << 1/κMIN2 where κΜΙΝ2 is the smallest eigenvalue.
Without loss of generality, we may write
H 0 ( X ) = 1 X 2
Indeed, the most general steady state is of the form h0(X) = ho√(1 − X2), for some constant ho, in terms of the original (dimensionful variable). We may then set the height scale [h] = ho, which fixes the scale of time by Equation (3):
[ t ] = n k L 2 h o
That of course means that the scales of storage and outflow rate are also fixed
[ S ] = n h o L , [ Q ] = [ S ] [ t ] = k h o 2 L

3.2. Numerical Treatment of the Problem

It is useful first to solve numerically the eigenvalue problem (14) for large enough eigenvalues κ. (In the next section we shall calculate analytically the critical quantities). By Equation (24), the early-time behavior of the flow rate is dominated by the asymptotic behavior of the various quantities of large eigenvalues. The numerical solution of the eigenvalue Equation (14) is facilitated by the observation that the boundary condition Φ′κ(0) = 0, and normalizing so that Φκ(0) = 1, the first two terms in a power series expansion of the eigenfunctions read
Φ κ ( X ) = 1 1 2 κ 2 X 2
Therefore, Equation (14) can then be solved with initial conditions Φκ(ε) = 1 − (1/2)κ2ε2 and Φ′κ(ε)= −κ2ε, with ε is a small number and the eigenvalues κ2 are determined by the condition that Φκ(1) is zero (within tolerance). We use ε = 10−9 and |Φκ(1)| < 10−8. We obtain results for the (root) eigenvalues κ, and the constants dκ and eκ defined in Equations (20) and (23), respectively, for the first 100 modes.
We find that the smallest root eigenvalue is κ1 = 1.50. That is, the longest mode decay time is 1/κ12 = 0.44. That means that the solution should make sense for times T << 0.44. On the other hand, for a given time T, the series adequately represents the solution of (10) when we include modes such that κ2T >> 1, that is, when we include decay times such that 1/κ2 << Τ. We have included modes up to decay times of order 10−5; hence, we may safely represent times nearly as small as 0.0001.
We may also utilize these results to deduce the behavior of the outflow rate given by Equation (24) for small times, looking for a more explicit expression than Equation (24). To this end, we shall also make explicit certain properties of the eigenvalue and eigenfunctions which define the solution to the problem at hand. First of all, at T = 0+, the flow rate reads
Q ( 0 + ) = κ d κ e κ
Although the time derivative of the profile and hence of the storage is discontinuous at T = 0, the flow rate, defined by (24), is continuous: Q(0+)= Q(0) = 1, using the fact that the steady state (25) produces unit flow rate from the results of the 100 modes we obtain [that is, the error appears to be of order O(100−2)]. In order to determine the behavior of (37) for small T, we need the asymptotic behavior of (square root) eigenvalues κ and of the product eκdκ. As we shall explicitly see below, the behavior of these quantities is needed to deduce the asymptotic early-time behavior of the outflow rate (25).
Indeed, we find that
κ ( i ) = 2.622 i 1.093
where i is the integer index of the eigenfunction corresponding to the number of its nodes. This is obtained by calculating the slope and the intercept of the line κ(i) calculated in a moving window of 10 eigenvalues. The results are plotted in Figure 1 where the slope of the line κ(i) is shown in a dotted line (primary axes) and its intercept is shown dashed line (secondary axes). As we shall see, what matters is the linearity of the relationship between the eigenvalue κ and the integer index i.
The behavior of the product eκdκ is best described by the logarithmic derivative with respect to κ as a function of the node number i, which illustrates a power law dependence of the product on κ. Figure 2 illustrates this dependence for up to node number 100. The result suggests that the product eκdκ exhibits a power law dependence on large κ:
e κ d κ = constant × κ ν , ν = 2.33 7 / 3
Such an estimate allows us to quantify analytically the early-time behavior of the flow rate Q in this problem. Indeed, the small time T behavior of Q given by Equation (24) is dominated by the large κ behavior of the summand. In this limit, the discrete index i, or equivalently via Equation (30), the index κ, is regarded as a continuous variable:
Q ( T ) = constant α κ ν e κ 2 T d κ
where α denotes an (undeterminable) lower-end point of order 1. We have
Q ( T ) = constant α 1 ν v 1 + 1 2 Γ 1 ν 2 T ν 1 2 + O ( T )
for the small times T. We know that the constant term must be 1. We also observe that the dominant term in the expansion of small times is a non-analytic power law. The reason is that the exponent ν ≈ 7/3 does not allow expanding the exponential to obtain even just one term (as the sum does not converge), or equivalently, the sum is not differentiable even once near T = 0, as we expected from our discussion previously. Explicitly, we have
Q ( T ) = 1 c × T 2 / 3 + O ( T )
where the c is a constant of order 1, which is not determined analytically in this work. Equation (34) encodes explicitly the dynamical response of the aquifer to the initial state, which respects the boundary conditions but it is not consistent with its dynamical preservation.
It is clear that Equation (34) applies to even earlier times than the complete result in Equation (24). Upon comparing it with the result of Equation (24), we may determine the numerical constant c. Indeed, for c = 1.414, the result (34) fits very well with the earliest time behavior of (24). The resulting curves are shown in Figure 3. Indeed, for the smallest times of order 0.0001, the relative difference is of order 1:105 while at time 0.05 is 0.5%. It is also worth quoting the result (34) in dimensionful form, using the scales (26) and (27), and the fact that h0 = √(r/k)L for the steady state with uniform and constant recharge rate r. We then have
( dimensionful ) Q ( t ) = Q 0 1.4 ( Q 0 ) 5 / 3 L 4 / 3 n 2 / 3 t 2 / 3 , Q 0 = r L
This solution holds well for dimensionful times t up to 0.01 nL3/2k−1/2Q0−1/2. For example, if we consider a porosity of 30%, aquifer length of 1 km, mean hydraulic conductivity 10−4 m/sec and discharge rate of 10−4 m3/sec, the solution holds very accurately for 11 days since the recharge switch-off.

3.3. Analytical Treatment of the Problem by Large Eigenvalue Asymptotics

We may complete our analysis by estimating analytically the important results (30) and (31). We start by determining the asymptotic behavior of the large (square root) asymptotic behavior of the solution of (14). This is a well-known type of problem, treated usually by standard asymptotic methods, such as the WKB approximation [25]. It is instructive to first obtain some insight into the problem by looking at the behavior of (14) near the boundary points. We first observe that for X~0, the eigenvalue Equation (14) takes the form
Φ κ ( Χ ) = κ 2 Φ κ ( X )
applying also the boundary condition Φ′κ(0) = 0 and normalization condition Φκ(0) = 1. Then, the solution of (36) is
Φ κ ( X ) = cos ( κ X )
Applying the WKB approximation in its first two orders (which are usually adequate) is essentially equivalent to looking for a solution in the form of the following ansatz
Φ κ ( X ) = A ( X ) cos ( κ ϕ ( X ) )
for some amplitude function A(X) and phase function ϕ ( X ) to be specified. Indeed, by substituting (38) into (14), we find (dividing also everything with κ2):
A ( X ) cos ( κ ϕ ( X ) ) ( ϕ ( X ) ) 2 + 1 H 0 ( X ) + 1 κ sin ( κ ϕ ( X ) ) 2 A ( X ) ϕ ( X ) A ( X ) ϕ ( X ) + 1 κ 2 cos ( κ ϕ ( X ) ) A ( X ) = 0
For large κ, the dominant term with respect to κ implies the following simple equation for the unknown function ϕ ( X ) :
( ϕ ( X ) ) 2 = 1 H 0 ( X )
Hence, we obtain
ϕ ( X ) = 0 X d X H 0 ( X )
Thus, we have determined the wave part of the WKB approximation without yet determining the prefactor A(X). This solution allows us immediately to determine the (asymptotic) form of the eigenvalues. Indeed, applying the boundary condition Φκ(1) = 0 on the ansatz (38) means that it must be cos(κφ(1)) = 0. Hence, we find
κ ( i ) = ( i + 1 2 ) π ϕ ( 1 )
We see that the eigenvalue is indeed a linear function of the integer counting index i. Moreover, let us apply this result to the case of steady-state initial condition (38): H0(X) = √(1 − X2). Then
ϕ ( 1 ) = 0 1 d X ( 1 X 2 ) 1 / 4 = 2 π Γ ( 3 4 ) Γ ( 1 4 ) = 1.198140
Hence, the slope of κ(i) is
κ ( i + 1 ) κ ( i ) = π ϕ ( 1 ) = 2.62206
which is the asymptotic slope we estimated numerically (Figure 1).
Impose now the first subdominant term, of order 1/κ, in (39). We obtain
2 A ( X ) ϕ ( X ) A ( X ) ϕ ( X ) = 0
This implies
A ( X ) = C ϕ ( X ) = C H 0 1 / 4
Applying this equation specifically for the case of the steady-state initial condition and applying the normalization Φκ(0) = 1, we find
A ( X ) = ( 1 X 2 ) 1 / 8
In all, the eigenfunction Φκ(X) under these approximations reads
Φ κ ( X ) = ( 1 X 2 ) 1 / 8 cos ( κ 0 X ( 1 χ 2 ) 1 / 4 d χ )
Note that, by Equation (40),
ϕ ( 1 ) ϕ ( X ) = X 1 ( 1 χ 2 ) 1 / 4 d χ
This allows us to rewrite (48) in the form
Φ κ ( X ) = ( 1 X 2 ) 1 / 8 sin ( κ X 1 ( 1 χ 2 ) 1 / 4 d χ )
where we have dropped a factor sin(κφ(1)) = ±1. Obtaining (50), we used the condition that defines the eigenvalues, i.e., cos(κφ(1)) = 0.
We may now come to the calculation of dκeκ. The quantity dκ is defined as in Equation (20). It involves the quantity Nκ, which is defined in Equation (15). The quantity eκ is defined in Equation (23). Working with the steady-state initial condition, we see that we should estimate the following three integrals for large κ:
0 1 Φ κ ( X ) d X , 0 1 Φ κ 2 ( X ) 1 X 2 d X , 0 1 Φ κ ( X ) 1 X 2 d X
In particular, we need to estimate their scaling with κ.
There are quick but somewhat handwaving ways to determine this scaling based on Equation (50); hence, we shall refrain from writing them down. The robust way to obtain the desired results is to proceed by a pertinent general asymptotic procedure, such as the steepest descent method [25]. The cumbersome chain of details of this calculation does not add anything to our main argument and we present it in Appendix C. Indeed, we find that for large κ
0 1 Φ κ ( X ) d X κ 3 / 2
0 1 Φ κ 2 ( X ) 1 X 2 d X = constant
0 1 Φ κ ( X ) 1 X 2 d X κ 5 / 6
Therefore, using these results in Equations (20) and (23) we find
d κ e κ κ 3 / 2 κ 5 / 6 = κ 7 / 3
which is what we estimated numerically above (Figure 2).
It is instructive to obtain the important result (55) in a less general but more direct way. This can be carried out by looking at the behavior of the eigenvalue Equation (14) near the other end of the interval, X~1. This is a non-trivial point, as the coefficient κ2/H0(Χ) diverges there. Usually, such singularities conceal, and with a bit of effort give easily away, a lot of important information about the solution. Moreover, as we learned explicitly through the steepest descent analysis in Appendix C, the large κ asymptotics rely on the behavior of the solutions in the neighborhood of that singular point. This fact can be understood from the very form of the eigenvalue equation: the coefficient κ2/H0(Χ) is large either by κ being large or by X being close to 1.
Indeed for X~1, and focusing specifically on the steady-state initial condition, we may approximate H0(X) = √(1 − X2) = √2(1 − X). Then, Equation (14) reads
Φ κ ( Χ ) + κ 2 2 ( 1 X ) Φ κ ( X ) = 0
This equation is solved in terms of the Bessel function of the first kind (Arfken et al. [26]):
Φ κ ( X ) = C 1 X J 2 / 3 ( 2 3 2 3 / 4 κ ( 1 X ) 3 / 4 )
where C is an integration constant. There is also a solution in terms of the Bessel function of order –2/3. But this solution is not zero at X = 1; hence, it is not consistent with the boundary condition Φκ(1) = 0, and hence the associated integration constant should vanish. The constant (2/3)23/4 = 1.12 appearing in (57) is an approximation of φ(1) in Equation (57), as a result of linearization of H0(X). We now apply the solution (57) over the whole interval imposing the normalization Φκ(0) = 1 to obtain
Φ κ ( X ) = 1 X J 2 / 3 ( 2 3 2 3 / 4 κ ( 1 X ) 3 / 4 ) J 2 / 3 ( 2 3 2 3 / 4 κ )
We apply the boundary condition Φ′κ(0) = 0. It can only be applied asymptotically for large κ. We obtain
cos ( 2 3 2 3 / 4 κ π 12 ) = 0
which produces again a linear relationship for κ with respect to the counting index. This linear relationship is an approximate form of the one given in (42).
We now have all the information we need to calculate the asymptotics of dκeκ. The three integrals (51) can be easily estimated for large κ. This is carried out in Appendix D, using the asymptotic forms of the Bessel functions and the condition (59). Hence, the result (55) is obtained through a different method. It is worth noting that the linearity of κ with respect to the counting index guaranteed also by (59) together with (55) is all we need to deduce the outflow rate early-time asymptotic in Equation (34). Hence, the near X~1 analysis is an alternative derivation of (34).

4. Early and Late-Time Asymptotics

In what follows, we present the (full) early-time solution, given by Equations (8) and (22) and the numerically obtained eigenfunctions, together with the late-time asymptotic solution introduced by [5]. Boussinesq looked for the simplest kind of self-similar solution, an exact separable solution, which is non-trivial due to the non-linearity of the Boussinesq equation. This important solution is presented in Appendix B. The simultaneous presentation of both asymptotics, together with the numerical solution of the Boussinesq equation, illustrates in a specific manner their extent of validity and the transition between them, expressed in various variables: water table profile, storage and outflow.
A direct test is provided by the time variation in the ratio S2/Q, which is constant for the separable solution. This is given in Figure 4.
The ratio S2/Q is a pure constant with a value of 0.693 for the separable solution, as shown in Appendix B. The numerical solution of the Boussinesq Equation (5) provides the storage by spatial integration, while the outflow rate is deduced by the mass balance, Q(T) = –S′(T). Presumably, these results show first that the separable solution is indeed a ‘late’-time asymptotic: its deviation from the Boussinesq equation solution is 1% at time 0.11, and 1‰ at time 0.29. The early-time asymptotic holds for times much earlier than 0.44 (as expected from our general considerations): the deviation from the solution of the Boussinesq equation appears to be 5% at time 0.05. On the other hand, we shall see below the ratio S2/Q provides a rather conservative criterion for convergence between solutions.
At the level of the time variation in the storage and outflow, the difference between the solutions is much less pronounced, as shown in Figure 5. Regarding the outflow rate, the two asymptotic solutions deviate visibly from the Boussinesq solution in a neighborhood of T = 0.1. In the case of the storage, the early-time solution eventually deviates from the Boussinesq solution, while the separable solution appears to provide a fairly good approximation of the Boussinesq solution at essentially all times.
It should be mentioned that, in order to produce the specific curves for the separable solution, one has to fix the integration constant T0. There is no unique choice for this number, and it is possibly the only remnant of the initial condition. If T0 is fixed too small, the convergence of the solution to the Boussinesq equation solution is delayed; if it is fixed too large, there is a larger disagreement between the solution and the Boussinesq solution at earlier times. To obtain a guiding guess, we may fix T0 so that the separable solution at T = 0 contains the same volume as the initial (steady) state, i.e., π/4, although this is not physically necessary or even reasonable. This gives T0 = (4/π)(4/9J3) = 0.882. A slightly different value T0 = 0.887 appears to give better results and is the one adopted here.
In Figure 6, comparisons of the water table profile are given between asymptotic solutions and the solution of the Boussinesq equation. For clarity purposes, the comparisons are given separately, following the same color coding as in Figure 4 and Figure 5. All curves are drawn for times T = 0.002, 0.02, 0.05, 0.10, 0.15, 0.20, 0.30 and 0.40 (time increases downwards). As expected from our analysis above, the early-time solution deviates visibly after time 0.10, while the separable solution observes well the behavior of the Boussinesq equation after time 0.10.

5. Summary and Conclusions

In the present work, we deduce the early-time response solution of the horizontal, unconfined aquifer to a sudden switching-off of the previously constant recharge. The results are not only novel but in some sense complete the list of fundamental asymptotic solutions of the Boussinesq equation for the recession phase, in addition to the solutions obtained by Boussinesq and Polubarinova-Kochina in the past. The steady-state initial profile is chosen as a simple and naturally smooth initial condition, in contrast to the Polubarinova solution, which starts off with a uniform initial profile in the finite-length aquifer.
As opposed to these classic solutions, the problem at hand cannot be solved by self-similarity arguments. Indeed, the problem is analyzed by solving the dynamics of the depth profile perturbations off the initial condition. This essentially translates to solving an eigenvalue/eigenfunction problem. This problem cannot be solved in closed form. It is either solved numerically for all eigenvalues, or by analytical means for large eigenvalues. The latter takes the form of asymptotic analysis utilizing, e.g., the WKB approximation to solve the eigenfunction problem for large eigenvalues, which carry the behavior of the aquifer for the earliest of times. Indeed, we establish that the outflow Q drops by 2/3 power law of time at the very early stages of the sudden recession. This verifies our argument, given in the introduction, that dQ/dt forms an integrable singularity at the beginning of the recession. We also quantify explicitly the times through which this law holds. The numerical solution of the eigenfunction problem is used to illustrate and compare the perturbations solution against the numerical solution of the full non-linear Boussinesq equation as well as the late time asymptotic by Boussinesq.
The details of our derivations show that with moderate effort it may be possible to establish the early response of the aquifer to the sudden recession for generic initial conditions, or at least, more general than a steady state. This endeavor, which may lead to statements of greater hydrologic interest, that is, set such results against field and laboratory data, will be left for future work. Our present results can certainly be used as a benchmark solution for modeling or numerical validation purposes.

Author Contributions

All authors contributed equally to “Conceptualization; methodology; validation; formal analysis; resources; data curation; writing—original draft preparation; writing—review and editing; visualization; supervision; project administration”. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Early-Time Recession Asymptotic Solution: The Polubarinova Solution

Consider a self-similar solution of the form:
H ( X , T ) = f 1 X T   i . e .   H ( X , T ) = f ( Y ) ,   where   Y = 1 X T
that is, we assume a square root scale for X, which also transforms linearly to facilitate imposing the outlet boundary condition. The transformation is explicitly given in the definition of the coordinate Y. This type of self-similarity coordinate is dictated by the diffusion form of the Boussinesq equation.
In order to be consistent with the boundary condition (6) of vanishing depth at outlet X = 1 we impose
f ( 0 ) = 0
Now, (A1) implies that
H ( X , T = 0 + ) = f ( )
for all X < 1. That is, Equation (A1) defines a profile that is uniform at all points, except at the outlet where it vanishes. The uniform value we may set equal to 1, i.e., f(∞) = 1. The discontinuous and non-differentiable initial profile will immediately change to a differentiable square profile but with a very steep slope, leading to an initially infinite outflow rate. It turns out that the singularity, which is associated with the diffusion type of scale for the x coordinate, will be integrable.
For the self-similar solution (A1) the Boussinesq equation takes the following form of an ordinary differential equation
( f 2 ( Y ) ) + Y f ( Y ) = 0
This equation has been dealt with power series by [4] and has been studied and exploited by various authors since, as discussed in the introduction. What matters to us is to obtain the flow rate associated with this self-similar solution.
To this end, it is more convenient to work with the following dependent variable:
g ( Y ) = 1 f 2 ( Y )
In this variable we have
g ( 0 ) = 1 , g ( ) = 0
and
g ( Y ) + Y 2 g ( Y ) 1 g ( Y ) = 0
Given that the Boussinesq equation has not been approximated, Equation (7) still hold and the outflow rate is given by
Q = ( f 2 ) ( 0 ) 2 T = g ( 0 ) 2 T = α T
where in the last equality we have introduced Polubarinova’s α, which is related to the function g by g′(0) = –2α. Therefore, the flow rate depends only on the values of g in the neighborhood of Y = 0. Moreover, the flow rate exhibits a square root (that is, integrable) infinity at T = 0 and. Being integrable allows for a smooth early-time evolution of storage (by Equation (7))
S = 1 2 α T
utilizing the initial condition (A3) which implies that S = 1 at T = 0.
The systematic way to deduce the number α is explained by Polubarinova, i.e., by forming a power series solution, α is determined as a root of a large order polynomial. Having reduced the original problem to the ordinary differential Equation (A7), perhaps the fastest way to determine the value of α is to solve (A7) numerically which is easily carried out.
One may first observe that the fact that g→0 for Y→∞, implies that for large Y it is described by the linear equation
g ( Y ) + Y 2 g ( Y ) = 0
This implies that g decreases very fast for large Y:
g ( Y ) = C e 1 4 Y 2 ,   i . e .   g ( Y ) = C Y e 1 4 Y 2 d Y ~ e 1 4 Y 2 Y
where C is an integration constant.
Hence, we solve numerically the ordinary differential Equation (A7) using the following end-point conditions:
g ( 0 ) = 1 ε , g ( 15 ) = 0

Appendix B. Late Time Recession Asymptotic Solution: The Separable (Boussinesq 1904) Solution

Separable solutions of Equation (5) read
H ( X , T ) = F ( X ) G ( T )
where F and G are functions of a single variable, which obey the equations
G ( T ) G 2 ( T ) = λ 2 = 1 2 ( F 2 ( X ) ) F ( X )
where λ is a constant. One then obtains that
G ( T ) = ( λ T 2 + c 1 ) 1
where c1 is an integration constant, and
F ( X ) = 2 λ 2 3 J 2 w ( X )
where
J = 0 1 w d w 1 w 3 = 5 π Γ ( 2 3 ) Γ ( 1 6 ) = 0.86237
while the new spatial function w(X) is defined by
w d w 1 w 3 = J d X
In the derivation of these formulas the boundary conditions (6) were used, expressed in the form w′(0) = 0 and w(1) = 0. The function w(X) is normalized so that w(0) = 1. Then,
h ( X , T ) = ( T + T 0 ) 1 2 3 J 2 w ( X )
where the constant λ2 cancels out, and we have redefined the constant of integration in the form of a time constant T0. We then have
Q = 4 9 J 3 ( T + T 0 ) 2
utilizing the fact that Equation (B6), together with the boundary condition w(1) = 0, implies that
( w 2 ) ( 1 ) = 2 J
and
S = 4 9 J 3 ( T + T 0 ) 1
using the fact that
0 1 w ( X ) d X = 0 1 w 2 J 1 w 3 d w = 2 3 J
where in the first equality we used (B6) to change variable from X to w.
One should note first that the storage and outflow are related by
d S d T + Q = 0
which is the mass balance relation, dictated and expressed by the Boussinesq equation.
Equations (B8) and (B10) also imply that the storage-outflow relationship is independent of time and the constant T0:
S 2 Q = 4 9 J 3 = 0.693006 = ln 2   ( 3   s . f . )
where the second equality holds in 3 significant figures. This ratio provides a direct way to see that the separable solution is a late time asymptotic of the drainage phase.

Appendix C

Consider the first integral in (51). Using (50) and changing variables by X = cosθ and χ = cosψ, we obtain
0 1 Φ κ ( X ) d X = 0 π / 2 ( cos θ ) 5 / 4 sin ( κ θ π / 2 cos ψ d ψ ) d θ
We may rewrite this in the form
0 1 Φ κ ( X ) d X = Im 0 π / 2 exp i κ θ π / 2 cos ψ d ψ + 5 4 ln ( cos θ ) d θ
where Im(·) is the imaginary part function of the argument and i is the imaginary unit: i2= –1. This integral is of the general form where the steepest descent applies [25]. Consider an integral of the form
g ( x ) e i κ f ( x ) d x = e i κ f ( x ) + ln g ( x ) d x
(integrating over a specific interval) so that we want its asymptotic form for large positive κ. We first find a stationary point x0 of the exponent:
i κ f ( x 0 ) + g ( x 0 ) g ( x 0 ) = 0
Then, we Taylor expand iκf(x) + lng(x) around the critical point (x = x0 + ε) and perform the arising Gaussian integral with respect to ε taken over a suitable straight line in the complex plane.
In our case the first step is given by
i κ cos θ 0 5 4 sin θ 0 cos θ 0 = 0
It is clear that the solution is of the form θ0 = π/2 − δ for some small number δ. Indeed, we find that
θ 0 = π 2 e i π / 3 5 4 κ 2 / 3
solves (C5) for large κ. Note that the solution is a complex number.
The zeroth order term iκf(x0) + lng(x0) of the Taylor expansion in our case needs performing the integral over ψ which is over is small interval in the neighborhood of π/2. This is easily carried out by expanding the cosine: cosψ= − (ψ − π/2). We finally find, substituting also into the exponential, the factor:
e 6 / 5 e i 5 π / 12 5 4 κ 5 / 6
Then, we have to evaluate the second derivative of iκf(x) + lng(x) at the critical point. In our case this reads
i κ sin θ 0 2 cos θ 0 5 4 1 cos 2 θ 0
Substituting (C6) into this expression we find
e i π / 3 3 2 4 5 1 / 3 κ 4 / 3
Our Taylor expansion around the critical point will be of the form
θ = θ 0 + e i π / 3 ε
with ε being a real number to be integrated over the real line. Thus, we calculate the prefactor of steepest descent by the integral
exp 1 2 e i π / 3 3 2 4 5 1 / 3 κ 4 / 3 e i π / 3 ε 2 d ε = exp 3 4 4 5 1 / 3 κ 4 / 3 ε 2 d ε
which gives
π 3 4 4 5 1 / 3 1 / 2 κ 2 / 3
Putting the factors (C6) and (C11) together we find (dropping real numerical constant factors)
e i 5 π / 12 κ 5 / 6 κ 2 / 3 = e i 5 π / 12 κ 3 / 2
The desired result is given by the imaginary part of this quantity, which is not zero. Hence, we obtain Equation (52). The result (54) follows by the same steps, the difference being that the analogous relation to (C1) contains a factor (cosθ)1/4. The result (53) follows from the fact that cosine squared contains a non-trigonometric term (cos2x = 1/2 + 1/2cos2x) that leads to a constant asymptotic.

Appendix D

We shall be content to derive the result (52). We have
0 1 Φ κ ( X ) d X = 0 1 1 X J 2 / 3 ( 2 3 2 3 / 4 κ ( 1 X ) 3 / 4 ) J 2 / 3 ( 2 3 2 3 / 4 κ ) d X
The large argument asymptotics of the Bessel functions of the first kind of order p [26] reads:
J p ( x ) ~ π 2 x sin ( x p π 2 + π 4 )
That means that the denominator in (D1) takes the form
J 2 / 3 ( 2 3 2 3 / 4 κ ) = κ 1 / 2 3 2 3 / 8 π sin ( 2 3 2 3 / 4 κ π 12 ) κ 1 / 2
where we have used the condition (59) which provides the specific form of the eigenvalues in this approximation.
Next change variables in (D1) by 1–X = (κu)4/3, so that, using also (D3) to read
0 1 Φ κ ( X ) d X κ 3 / 2 0 κ u J 2 / 3 ( 2 3 2 3 / 4 u ) d u
where a factor κ2 came from the change in variable to produce, together with (D3), the factor κ−3/2. Now, if the integral is convergent for large κ, we may let κ go to infinity and hence obtain the desired result. Indeed, using the asymptotics (D2) we may verify this convergence as follows. First, we obtain
0 κ u J 2 / 3 ( 2 3 2 3 / 4 u ) d u ~ κ u sin ( 2 3 2 3 / 4 u π 12 ) d u
where we focus on the upper limit of the integral, dropping also all irrelevant numerical factors. Integrating once by parts we obtain
κ u sin ( 2 3 2 3 / 4 u π 12 ) d u ~ κ cos ( 2 3 2 3 / 4 κ π 12 ) κ 1 2 u cos ( 2 3 2 3 / 4 u π 12 ) d u
We see that the first, potentially divergent term, vanishes identically by condition (59). The remaining integral is perfectly convergent.

References

  1. Boussinesq, J. Essai sur la theorie des eaux courantes du mouvement nonpermanent des eaux souterraines. Acad. Sci. Inst. Fr. 1877, 23, 252–260. [Google Scholar]
  2. Childs, E.C. Drainage of groundwater resting on a sloping bed. Water Resour. Res. 1971, 7, 1256–1263. [Google Scholar] [CrossRef]
  3. Dupuit, J. Etudes Theoriques et Practiques Sur le Mouvement des Eaux Dans les Canaux Decouverts et a Travers les Terrains Permeables, 2nd ed.; Dunod: Paris, France, 1863. [Google Scholar]
  4. Polubarinova-Kochina, P.Y. Theory of Ground Water Movement; Princeton University Press: Princeton, NJ, USA, 1962. [Google Scholar]
  5. Boussinesq, J. Recherches theoriques sur l’ecoulement des nappes d’eau infiltrees dans le sol et sur debit de sources. J. Math. Pures Appl. 1904, 10, 5–78. [Google Scholar]
  6. Lockington, D.A. Response of unconfined aquifer to sudden change in boundary head. J. Irrig. Drain. Eng. 1997, 123, 24–27. [Google Scholar] [CrossRef]
  7. Parlange, J.; Stagnitti, F.; Heilig, A.; Szilagyi, J.; Parlange, M.B.; Steenhuis, T.S.; Hogarth, W.L.; Barry, D.A.; Li, L. Sudden drawdown and drainage of a horizontal aquifer. Water Resour. Res. 2001, 37, 2097–2101. [Google Scholar] [CrossRef]
  8. Brutsaert, W.; Lopez, J.P. Basin-scale geohydrologic drought flow features of riparian aquifers in the southern Great Plains. Water Resour. Res. 1998, 34, 233–240. [Google Scholar] [CrossRef]
  9. Bear, J. Dynamics of Fluids in Porous Media; Dover Publications: New York, NY, USA, 1972. [Google Scholar]
  10. Barenblatt, G.I.; Entov, V.M.; Ryzhik, V.M. Theory of Fluid Flows Through Natural Rocks; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1990. [Google Scholar]
  11. Parlange, J.Y.; Hogarth, W.L.; Govindaraju, R.S.; Parlange, M.B.; Lockington, D. On an exact analytical solution of the Boussinesq equation. Transp. Porous Media 2000, 39, 339–345. [Google Scholar] [CrossRef]
  12. Lockington, D.; Parlange, J.-Y.; Parlange, M.; Selker, J. Similarity solution of the Boussinesq equation. Adv. Water Resour. 2000, 23, 725–729. [Google Scholar] [CrossRef]
  13. Telyakovskiy, A.S.; Allen, M.B. Polynomial approximate solutions to the Boussinesq equation. Adv. Water Resour. 2006, 29, 1767–1779. [Google Scholar] [CrossRef]
  14. Telyakovskiy, A.S.; Braga, G.A.; Kurita, S.; Mortensen, J. On a power series solution to the Boussinesq equation. Adv. Water Resour. 2010, 33, 1128–1129. [Google Scholar] [CrossRef]
  15. Moutsopoulos, K.N. The analytical solution of the Boussinesq equation for flow induced by a step change of the water table elevation revisited. Transp. Porous Media 2010, 85, 919–940. [Google Scholar] [CrossRef]
  16. Moutsopoulos, K.N. Solutions of the Boussinesq equation subject to a nonlinear Robin boundary condition. Water Resour. Res. 2013, 49, 7–18. [Google Scholar] [CrossRef]
  17. Olsen, J.S.; Telyakovskiy, A.S. Polynomial approximate solutions of a generalized Boussinesq equation. Water Resour. Res. 2013, 49, 3049–3053. [Google Scholar] [CrossRef]
  18. Chor, T.; Dias, N.L.; de Zárate, A.R. An exact series and improved numerical and approximate solutions for the Boussinesq equation. Water Resour. Res. 2013, 49, 7380–7387. [Google Scholar] [CrossRef]
  19. Dias, N.L.; Chor, T.L.; de Zárate, A.R. A semianalytical solution for the Boussinesq equation with nonhomogeneous constant boundary conditions. Water Resour. Res. 2014, 50, 6549–6556. [Google Scholar] [CrossRef]
  20. Jiang, Q.; Tang, Y. A general approximate method for the groundwater response problem caused by water level variation. J. Hydrol. 2015, 529, 398–409. [Google Scholar] [CrossRef]
  21. Hayek, M. Accurate approximate semi-analytical solutions to the Boussinesq groundwater flow equation for recharging and discharging of horizontal unconfined aquifers. J. Hydrol. 2019, 570, 411–422. [Google Scholar] [CrossRef]
  22. Akylas, E.; Gravanis, E. Approximate Solutions of the Boussinesq Equation for Horizontal Unconfined Aquifers During Pure Drainage Phase. Water 2024, 16, 2984. [Google Scholar] [CrossRef]
  23. Gravanis, E.; Akylas, E. Early-time solution of the horizontal unconfined aquifer in the buildup phase. Water Resour. Res. 2017, 53, 8310–8326. [Google Scholar] [CrossRef]
  24. Gravanis, E.; Akylas, E.; Sarris, E.N. Approximate Solutions for Horizontal Unconfined Aquifers in the Buildup Phase. Water 2024, 16, 1031. [Google Scholar] [CrossRef]
  25. Bender, C.M.; Orszag, S.A. Advanced Mathematical Methods for Scientists and Engineers I: Asymptotic Methods and Perturbation Theory; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
  26. Arfken, G.B.; Weber, H.J.; Harris, F.E. Mathematical Methods for Physicists: A Comprehensive Guide; Academic Press: Cambridge, MA, USA, 2011. [Google Scholar]
Figure 1. Dotted line (primary axes): slope of κ(i). Dashed–dotted line (secondary axes): (Minus the) intercept of κ(i).
Figure 1. Dotted line (primary axes): slope of κ(i). Dashed–dotted line (secondary axes): (Minus the) intercept of κ(i).
Water 17 00771 g001
Figure 2. The logarithmic derivative of the product eκdκ as function of the number of nodes.
Figure 2. The logarithmic derivative of the product eκdκ as function of the number of nodes.
Water 17 00771 g002
Figure 3. Comparison between the early-time asymptotic (24) and the power law asymptotic (34) for the flow rate Q.
Figure 3. Comparison between the early-time asymptotic (24) and the power law asymptotic (34) for the flow rate Q.
Water 17 00771 g003
Figure 4. The ratio of storage squared over outflow rate for the asymptotic solutions and numerical solution of the Boussinesq equation.
Figure 4. The ratio of storage squared over outflow rate for the asymptotic solutions and numerical solution of the Boussinesq equation.
Water 17 00771 g004
Figure 5. Outflow rate (left) and storage (right) time variation for the asymptotic solutions and numerical solution of the Boussinesq equation.
Figure 5. Outflow rate (left) and storage (right) time variation for the asymptotic solutions and numerical solution of the Boussinesq equation.
Water 17 00771 g005
Figure 6. Comparison of the water table time variation between the early-time solution (left) and the separable solution (right) with the numerical solution of the Boussinesq equation.
Figure 6. Comparison of the water table time variation between the early-time solution (left) and the separable solution (right) with the numerical solution of the Boussinesq equation.
Water 17 00771 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gravanis, E.; Akylas, E.; Sarris, E.N. Early-Time Recession Solution from a Steady-State Initial Condition for the Horizontal Unconfined Aquifer. Water 2025, 17, 771. https://doi.org/10.3390/w17050771

AMA Style

Gravanis E, Akylas E, Sarris EN. Early-Time Recession Solution from a Steady-State Initial Condition for the Horizontal Unconfined Aquifer. Water. 2025; 17(5):771. https://doi.org/10.3390/w17050771

Chicago/Turabian Style

Gravanis, Elias, Evangelos Akylas, and Ernestos N. Sarris. 2025. "Early-Time Recession Solution from a Steady-State Initial Condition for the Horizontal Unconfined Aquifer" Water 17, no. 5: 771. https://doi.org/10.3390/w17050771

APA Style

Gravanis, E., Akylas, E., & Sarris, E. N. (2025). Early-Time Recession Solution from a Steady-State Initial Condition for the Horizontal Unconfined Aquifer. Water, 17(5), 771. https://doi.org/10.3390/w17050771

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop