Finite water-content vadose zone flow method

Last updated

The finite water-content vadose zone flux method [1] [2] represents a one-dimensional alternative to the numerical solution of Richards' equation [3] for simulating the movement of water in unsaturated soils. The finite water-content method solves the advection-like term of the Soil Moisture Velocity Equation, which is an ordinary differential equation alternative to the Richards partial differential equation. The Richards equation is difficult to approximate in general because it does not have a closed-form analytical solution except in a few cases. [4] The finite water-content method, is perhaps the first generic replacement for the numerical solution of the Richards' equation. The finite water-content solution has several advantages over the Richards equation solution. First, as an ordinary differential equation it is explicit, guaranteed to converge [5] and computationally inexpensive to solve. Second, using a finite volume solution methodology it is guaranteed to conserve mass. The finite water content method readily simulates sharp wetting fronts, something that the Richards solution struggles with. [6] The main limiting assumption required to use the finite water-content method is that the soil be homogeneous in layers.

Contents

Finite Water-Content discretization. The porous medium is divided into n uniform "bins" of
D
th
{\displaystyle \Delta \theta }
water content. Fwc domain.png
Finite Water-Content discretization. The porous medium is divided into n uniform "bins" of water content.

The finite water-content vadose zone flux method is derived from the same starting point as the derivation of Richards' equation. However, the derivation employs a hodograph transformation [7] to produce an advection solution that does not include soil water diffusivity, wherein becomes the dependent variable and becomes an independent variable: [2]

where:

is the unsaturated hydraulic conductivity [L T−1],
is the capillary pressure head [L] (negative for unsaturated soil),
is the vertical coordinate [L] (positive downward),
is the water content, (−) and
is time [T].

This equation was converted into a set of three ordinary differential equations (ODEs) [2] using the Method of Lines [8] to convert the partial derivatives on the right-hand side of the equation into appropriate finite difference forms. These three ODEs represent the dynamics of infiltrating water, falling slugs, and capillary groundwater, respectively.

Derivation

A superior derivation was published [9] in 2017, showing that this equation is a diffusion-free version of the Soil Moisture Velocity Equation.

One way to solve this equation is to solve it for and by integration: [10]

Instead, a finite water-content discretization is used and the integrals are replaced with summations:

where is the total number of finite water content bins.

Using this approach, the conservation equation for each bin is:

The method of lines is used to replace the partial differential forms on the right-hand side into appropriate finite-difference forms. This process results in a set of three ordinary differential equations that describe the dynamics of infiltration fronts, falling slugs, and groundwater capillary fronts using a finite water-content discretization.

Method essentials

The finite water-content vadose zone flux calculation method replaces the Richards' equation PDE with a set of three ordinary differential equations (ODEs). These three ODEs are developed in the following sections. Furthermore, because the finite water-content method does not explicitly include soil water diffusivity, it necessitates a separate capillary relaxation step. Capillary relaxation [11] represents a free-energy minimization process at the pore scale that produces no advection beyond the REV scale.

Infiltration fronts

Infiltration fronts in finite water-content domain. Inf fronts.png
Infiltration fronts in finite water-content domain.

With reference to Figure 1, water infiltrating the land surface can flow through the pore space between and . In the context of the method of lines, the partial derivative terms are replaced with:

Given that any ponded depth of water on the land surface is , the Green and Ampt (1911) [12] assumption is employed,

represents the capillary head gradient that is driving the flow. Therefore the finite water-content equation in the case of infiltration fronts is:

Falling slugs

Falling slugs in the finite water-content domain. The water in each bin is considered a separate slug. Falling slugs.png
Falling slugs in the finite water-content domain. The water in each bin is considered a separate slug.

After rainfall stops and all surface water infiltrates, water in bins that contains infiltration fronts detaches from the land surface. Assuming that the capillarity at leading and trailing edges of this 'falling slug' of water is balanced, then the water falls through the media at the incremental conductivity associated with the -th bin:

Capillary groundwater fronts

Groundwater capillary fronts in finite water-content domain. Groundwater fronts.png
Groundwater capillary fronts in finite water-content domain.

In this case, the flux of water to the bin occurs between bin j and i. Therefore in the context of the method of lines:

and,

which yields:

The performance of this equation was verified for cases where the groundwater table velocity was less than 0.92 , [13] using a column experiment fashioned after that by Childs and Poulovassilis (1962). [14] Results of that validation showed that the finite water-content vadose zone flux calculation method performed comparably to the numerical solution of Richards' equation.

Capillary relaxation

Because the hydraulic conductivity rapidly increases as the water content moves towards saturation, with reference to Fig.1, right-most bins in both capillary groundwater fronts and infiltration fronts can "out-run" their neighbors to the left. In the finite water content discretization, these shocks [15] are dissipated by the process of capillary relaxation, which represents a pore-scale free-energy minimization process that produces no advection beyond the REV scale [11] Numerically, this process is a numerical sort that places the fronts in monotonically-decreasing magnitude from left-right.

Constitutive relations

The finite water content vadose zone flux method works with any monotonic water retention curve/unsaturated hydraulic conductivity relations such as Brooks and Corey [16] Clapp and Hornberger [17] and van Genuchten-Mualem. [18] The method might work with hysteretic water retention relations- these have not yet been tested.

Limitations

The finite water content method lacks the effect of soil water diffusion. This omission does not affect the accuracy of flux calculations using the method because the mean of the diffusive flux is small. Practically, this means that the shape of the wetting front plays no role in driving the infiltration. The method is thus far limited to 1-D in practical applications. The infiltration equation [2] was extended to 2- and quasi-3 dimensions. [5] More work remains in extending the entire method into more than one dimension.

Awards

The paper describing this method [2] was selected by the Early Career Hydrogeologists Network of the International Association of Hydrogeologists to receive the "Coolest paper Published in 2015" award in recognition of the potential impact of the publication on the future of hydrogeology.

See also

Related Research Articles

<span class="mw-page-title-main">Laplace's equation</span> Second-order partial differential equation

In mathematics and physics, Laplace's equation is a second-order partial differential equation named after Pierre-Simon Laplace, who first studied its properties. This is often written as

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

<span class="mw-page-title-main">Potential flow</span> Velocity field as the gradient of a scalar function

In fluid dynamics, potential flow is the ideal flow pattern of an inviscid fluid. Potential flows are described and determined by mathematical methods.

<span class="mw-page-title-main">Schrödinger equation</span> Description of a quantum-mechanical system

The Schrödinger equation is a linear partial differential equation that governs the wave function of a quantum-mechanical system. Its discovery was a significant landmark in the development of quantum mechanics. It is named after Erwin Schrödinger, who postulated the equation in 1925 and published it in 1926, forming the basis for the work that resulted in his Nobel Prize in Physics in 1933.

Fourier optics is the study of classical optics using Fourier transforms (FTs), in which the waveform being considered is regarded as made up of a combination, or superposition, of plane waves. It has some parallels to the Huygens–Fresnel principle, in which the wavefront is regarded as being made up of a combination of spherical wavefronts whose sum is the wavefront being studied. A key difference is that Fourier optics considers the plane waves to be natural modes of the propagation medium, as opposed to Huygens–Fresnel, where the spherical waves originate in the physical medium.

Geometrical optics, or ray optics, is a model of optics that describes light propagation in terms of rays. The ray in geometrical optics is an abstraction useful for approximating the paths along which light propagates under certain circumstances.

The Grad–Shafranov equation is the equilibrium equation in ideal magnetohydrodynamics (MHD) for a two dimensional plasma, for example the axisymmetric toroidal plasma in a tokamak. This equation takes the same form as the Hicks equation from fluid dynamics. This equation is a two-dimensional, nonlinear, elliptic partial differential equation obtained from the reduction of the ideal MHD equations to two dimensions, often for the case of toroidal axisymmetry. Taking as the cylindrical coordinates, the flux function is governed by the equation,

<span class="mw-page-title-main">Infiltration (hydrology)</span> Process by which water on the ground surface enters the soil and then

, water table, and phreatic or saturated zone. (Source: United States Geological Survey.)]]

<span class="mw-page-title-main">Routhian mechanics</span> Formulation of classical mechanics

In classical mechanics, Routh's procedure or Routhian mechanics is a hybrid formulation of Lagrangian mechanics and Hamiltonian mechanics developed by Edward John Routh. Correspondingly, the Routhian is the function which replaces both the Lagrangian and Hamiltonian functions. Routhian mechanics is equivalent to Lagrangian mechanics and Hamiltonian mechanics, and introduces no new physics. It offers an alternative way to solve mechanical problems.

<span class="mw-page-title-main">Water retention curve</span>

Water retention curve is the relationship between the water content, θ, and the soil water potential, ψ. This curve is characteristic for different types of soil, and is also called the soil moisture characteristic.

The Richards equation represents the movement of water in unsaturated soils, and is attributed to Lorenzo A. Richards who published the equation in 1931. It is a quasilinear partial differential equation; its analytical solution is often limited to specific initial and boundary conditions. Proof of the existence and uniqueness of solution was given only in 1983 by Alt and Luckhaus. The equation is based on Darcy-Buckingham law representing flow in porous media under variably saturated conditions, which is stated as

The Gross–Pitaevskii equation describes the ground state of a quantum system of identical bosons using the Hartree–Fock approximation and the pseudopotential interaction model.

<span class="mw-page-title-main">HydroGeoSphere</span>

HydroGeoSphere (HGS) is a 3D control-volume finite element groundwater model, and is based on a rigorous conceptualization of the hydrologic system consisting of surface and subsurface flow regimes. The model is designed to take into account all key components of the hydrologic cycle. For each time step, the model solves surface and subsurface flow, solute and energy transport equations simultaneously, and provides a complete water and solute balance.

In fluid dynamics, the Oseen equations describe the flow of a viscous and incompressible fluid at small Reynolds numbers, as formulated by Carl Wilhelm Oseen in 1910. Oseen flow is an improved description of these flows, as compared to Stokes flow, with the (partial) inclusion of convective acceleration.

<span class="mw-page-title-main">Cnoidal wave</span> Nonlinear and exact periodic wave solution of the Korteweg–de Vries equation

In fluid dynamics, a cnoidal wave is a nonlinear and exact periodic wave solution of the Korteweg–de Vries equation. These solutions are in terms of the Jacobi elliptic function cn, which is why they are coined cnoidal waves. They are used to describe surface gravity waves of fairly long wavelength, as compared to the water depth.

<span class="mw-page-title-main">Potential flow around a circular cylinder</span> Classical solution for inviscid, incompressible flow around a cyclinder

In mathematics, potential flow around a circular cylinder is a classical solution for the flow of an inviscid, incompressible fluid around a cylinder that is transverse to the flow. Far from the cylinder, the flow is unidirectional and uniform. The flow has no vorticity and thus the velocity field is irrotational and can be modeled as a potential flow. Unlike a real fluid, this solution indicates a net zero drag on the body, a result known as d'Alembert's paradox.

Partial-wave analysis, in the context of quantum mechanics, refers to a technique for solving scattering problems by decomposing each wave into its constituent angular-momentum components and solving using boundary conditions.

In physics, the distorted Schwarzschild metric is the metric of a standard/isolated Schwarzschild spacetime exposed in external fields. In numerical simulation, the Schwarzschild metric can be distorted by almost arbitrary kinds of external energy–momentum distribution. However, in exact analysis, the mature method to distort the standard Schwarzschild metric is restricted to the framework of Weyl metrics.

The soil moisture velocity equation describes the speed that water moves vertically through unsaturated soil under the combined actions of gravity and capillarity, a process known as infiltration. The equation is alternative form of the Richardson/Richards' equation. The key difference being that the dependent variable is the position of the wetting front , which is a function of time, the water content and media properties. The soil moisture velocity equation consists of two terms. The first "advection-like" term was developed to simulate surface infiltration and was extended to the water table, which was verified using data collected in a column experimental that was patterned after the famous experiment by Childs & Poulovassilis (1962) and against exact solutions.

In fluid dynamics, Hicks equation, sometimes also referred as Bragg–Hawthorne equation or Squire–Long equation, is a partial differential equation that describes the distribution of stream function for axisymmetric inviscid fluid, named after William Mitchinson Hicks, who derived it first in 1898. The equation was also re-derived by Stephen Bragg and William Hawthorne in 1950 and by Robert R. Long in 1953 and by Herbert Squire in 1956. The Hicks equation without swirl was first introduced by George Gabriel Stokes in 1842. The Grad–Shafranov equation appearing in plasma physics also takes the same form as the Hicks equation.

References

  1. Talbot, C.A., and F. L. Ogden (2008), A method for computing infiltration and redistribution in a discretized moisture content domain, Water Resour. Res., 44(8), doi: 10.1029/2008WR006815.
  2. 1 2 3 4 5 Ogden, F. L., W. Lai, R. C. Steinke, J. Zhu, C. A. Talbot, and J. L. Wilson (2015), A new general 1-D vadose zone solution method, Water Resour.Res., 51, doi:10.1002/2015WR017126.
  3. Richards, L. A. (1931), Capillary conduction of liquids through porous mediums, J. Appl. Phys., 1(5), 318–333.
  4. Ross, P.J., and J.-Y. Parlange (1994). Comparing exact and numerical solutions of Richards equation for one-dimensional infiltration and drainage. Soil Sci. Vol 1557, No. 6, pp. 341-345.
  5. 1 2 Yu, H., C. C. Douglas, and F. L. Ogden, (2012), A new application of dynamic data driven system in the Talbot–Ogden model for groundwater infiltration, Procedia Computer Science, 9, 1073–1080.
  6. Tocci, M. D., C. T. Kelley, and C. T. Miller (1997), Accurate and economical solution of the pressure-head form of Richards' equation by the method of lines, Adv. Wat. Resour., 20(1), 1–14.
  7. Philip, J. R. 1957. The theory of infiltration: 1. The infiltration equation and its solution, Soil Sci, 83(5), 345–357.
  8. Griffiths, Graham; Schiesser, William; Hamdi, Samir (2007). "Method of lines". Scholarpedia. 2 (7): 2859. Bibcode:2007SchpJ...2.2859H. doi: 10.4249/scholarpedia.2859 .
  9. Ogden, F.L., M.B. Allen, W.Lai, J. Zhu, C.C. Douglas, M. Seo, and C.A. Talbot, 2017. The Soil Moisture Velocity Equation, J. Adv. Modeling Earth Syst. https://doi.org/10.1002/2017MS000931
  10. Wilson, J. L. (1974), Dispersive mixing in a partially saturated porous medium, PhD dissertation, 355pp., Dept. of Civil Engrg., Mass. Inst. Tech., Cambridge, MA.
  11. 1 2 Moebius, F., D. Canone, and D. Or (2012), Characteristics of acoustic emissions induced by fluid front displacement in porous media, Water Resour. Res., 48(11), W11507, doi:10.1029/2012WR012525.
  12. Green, W. H., and G. A. Ampt (1911), Studies on soil physics, 1, The flow of air and water through soils, J. Agric. Sci., 4(1), 1–24.
  13. Ogden, F. L., W. Lai, R. C. Steinke, and J. Zhu (2015b), Validation of finite water-content vadose zone dynamics method using column experiments with a moving water table and applied surface flux, Water Resour. Res., 10.1002/2014WR016454.
  14. Childs, E. C., and A. Poulovassilis (1962), The moisture profile above a moving water table, J. Soil Sci., 13(2), 271–285.
  15. Smith, R. E. (1983), Approximate soil water movement by kinematic characteristics, Soil Sci. Soc. Am. J., 47(1), 3–8.
  16. Brooks, R.H., and A.T. Corey, 1964. Hydraulic properties of porous media. Hydrol. Pap. 3, Colo. State Univ., Fort Collins, Colorado, USA.
  17. Clapp R.B., and G.M. Hornberger, 1978. Empirical equations for some soil hydraulic properties. Water Resour. Res. 14(4):601–604
  18. van Genuchten, M. Th. (1980). "A Closed-Form Equation for Predicting the Hydraulic Conductivity of Unsaturated Soils" (PDF). Soil Sci. Soc. Am. J., 44 (5): 892-898. doi:10.2136/sssaj1980.03615995004400050002x