Richards equation

Last updated

The Richards equation represents the movement of water in unsaturated soils, and is attributed to Lorenzo A. Richards who published the equation in 1931. [1] It is a quasilinear partial differential equation; its analytical solution is often limited to specific initial and boundary conditions. [2] Proof of the existence and uniqueness of solution was given only in 1983 by Alt and Luckhaus. [3] The equation is based on Darcy-Buckingham law [1] representing flow in porous media under variably saturated conditions, which is stated as

Contents

where

is the volumetric flux;
is the volumetric water content;
is the liquid pressure head, which is negative for unsaturated porous media;
is the unsaturated hydraulic conductivity;
is the geodetic head gradient, which is assumed as for three-dimensional problems.

Considering the law of mass conservation for an incompressible porous medium and constant liquid density, expressed as

,

where

is the sink term [T], typically root water uptake. [4]

Then substituting the fluxes by the Darcy-Buckingham law the following mixed-form Richards equation is obtained:

.

For modeling of one-dimensional infiltration this divergence form reduces to

.

Although attributed to L. A. Richards, the equation was originally introduced 9 years earlier by Lewis Fry Richardson in 1922. [5] [6]

Formulations

The Richards equation appears in many articles in the environmental literature because it describes the flow in the vadose zone between the atmosphere and the aquifer. It also appears in pure mathematical journals because it has non-trivial solutions. The above-given mixed formulation involves two unknown variables: and . This can be easily resolved by considering constitutive relation , which is known as the water retention curve. Applying the chain rule, the Richards equation may be reformulated as either -form (head based) or -form (saturation based) Richards equation.

Head-based

By applying the chain rule on temporal derivative leads to

,

where is known as the retention water capacity . The equation is then stated as

.

The head-based Richards equation is prone to the following computational issue: the discretized temporal derivative using the implicit Rothe method yields the following approximation:

This approximation produces an error that affects the mass conservation of the numerical solution, and so special strategies for temporal derivatives treatment are necessary. [7]

Saturation-based

By applying the chain rule on the spatial derivative leads to

where , which could be further formulated as , is known as the soil water diffusivity . The equation is then stated as

The saturation-based Richards equation is prone to the following computational issues. Since the limits and , where is the saturated (maximal) water content and is the residual (minimal) water content a successful numerical solution is restricted just for ranges of water content satisfactory below the full saturation (the saturation should be even lower than air entry value) as well as satisfactory above the residual water content. [8]

Parametrization

The Richards equation in any of its forms involves soil hydraulic properties, which is a set of five parameters representing soil type. The soil hydraulic properties typically consist of water retention curve parameters by van Genuchten: [9] (), where is the inverse of air entry value [L−1], is the pore size distribution parameter [-], and is usually assumed as . Further the saturated hydraulic conductivity (which is for non isotropic environment a tensor of second order) should also be provided. Identification of these parameters is often non-trivial and was a subject of numerous publications over several decades. [10] [11] [12] [13] [14] [15]

Limitations

The numerical solution of the Richards equation is one of the most challenging problems in earth science. [16] Richards' equation has been criticized for being computationally expensive and unpredictable [17] [18] because there is no guarantee that a solver will converge for a particular set of soil constitutive relations. Advanced computational and software solutions are required here to over-come this obstacle. The method has also been criticized for over-emphasizing the role of capillarity, [19] and for being in some ways 'overly simplistic' [20] In one dimensional simulations of rainfall infiltration into dry soils, fine spatial discretization less than one cm is required near the land surface, [21] which is due to the small size of the representative elementary volume for multiphase flow in porous media. In three-dimensional applications the numerical solution of the Richards equation is subject to aspect ratio constraints where the ratio of horizontal to vertical resolution in the solution domain should be less than about 7.[ citation needed ]

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, named after French engineer and physicist Claude-Louis Navier and 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).

In mathematics, the Laplace operator or Laplacian is a differential operator given by the divergence of the gradient of a scalar function on Euclidean space. It is usually denoted by the symbols , (where is the nabla operator), or . In a Cartesian coordinate system, the Laplacian is given by the sum of second partial derivatives of the function with respect to each independent variable. In other coordinate systems, such as cylindrical and spherical coordinates, the Laplacian also has a useful form. Informally, the Laplacian Δf (p) of a function f at a point p measures by how much the average value of f over small spheres or balls centered at p deviates from f (p).

<span class="mw-page-title-main">Spherical harmonics</span> Special mathematical functions defined on the surface of a sphere

In mathematics and physical science, spherical harmonics are special functions defined on the surface of a sphere. They are often employed in solving partial differential equations in many scientific fields.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

In mathematics, the Helmholtz equation is the eigenvalue problem for the Laplace operator. It corresponds to the linear partial differential equation

Used in hydrogeology, the groundwater flow equation is the mathematical relationship which is used to describe the flow of groundwater through an aquifer. The transient flow of groundwater is described by a form of the diffusion equation, similar to that used in heat transfer to describe the flow of heat in a solid. The steady-state flow of groundwater is described by a form of the Laplace equation, which is a form of potential flow and has analogs in numerous fields.

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

Infiltration is the process by which water on the ground surface enters the soil. It is commonly used in both hydrology and soil sciences. The infiltration capacity is defined as the maximum rate of infiltration. It is most often measured in meters per day but can also be measured in other units of distance over time if necessary. The infiltration capacity decreases as the soil moisture content of soils surface layers increases. If the precipitation rate exceeds the infiltration rate, runoff will usually occur unless there is some physical barrier.

In fluid mechanics, potential vorticity (PV) is a quantity which is proportional to the dot product of vorticity and stratification. This quantity, following a parcel of air or water, can only be changed by diabatic or frictional processes. It is a useful concept for understanding the generation of vorticity in cyclogenesis, especially along the polar front, and in analyzing flow in the ocean.

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 mechanics and mathematics, a capillary surface is a surface that represents the interface between two different fluids. As a consequence of being a surface, a capillary surface has no thickness in slight contrast with most real fluid interfaces.

In mathematics, vector spherical harmonics (VSH) are an extension of the scalar spherical harmonics for use with vector fields. The components of the VSH are complex-valued functions expressed in the spherical coordinate basis vectors.

The derivatives of scalars, vectors, and second-order tensors with respect to second-order tensors are of considerable use in continuum mechanics. These derivatives are used in the theories of nonlinear elasticity and plasticity, particularly in the design of algorithms for numerical simulations.

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.

Miniaturizing components has always been a primary goal in the semiconductor industry because it cuts production cost and lets companies build smaller computers and other devices. Miniaturization, however, has increased dissipated power per unit area and made it a key limiting factor in integrated circuit performance. Temperature increase becomes relevant for relatively small-cross-sections wires, where it may affect normal semiconductor behavior. Besides, since the generation of heat is proportional to the frequency of operation for switching circuits, fast computers have larger heat generation than slow ones, an undesired effect for chips manufacturers. This article summaries physical concepts that describe the generation and conduction of heat in an integrated circuit, and presents numerical methods that model heat transfer from a macroscopic point of view.

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

In physics, Lagrangian mechanics is a formulation of classical mechanics founded on the stationary-action principle. It was introduced by the Italian-French mathematician and astronomer Joseph-Louis Lagrange in his presentation to the Turin Academy of Science in 1760 culminating in his 1788 grand opus, Mécanique analytique.

<span class="mw-page-title-main">Finite water-content vadose zone flow method</span>

The finite water-content vadose zone flux method represents a one-dimensional alternative to the numerical solution of Richards' equation 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. 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 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. The main limiting assumption required to use the finite water-content method is that the soil be homogeneous in layers.

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.

References

  1. 1 2 Richards, L.A. (1931). "Capillary conduction of liquids through porous mediums". Physics. 1 (5): 318–333. Bibcode:1931Physi...1..318R. doi:10.1063/1.1745010.
  2. Tracy, F. T. (August 2006). "Clean two- and three-dimensional analytical solutions of Richards' equation for testing numerical solvers: TECHNICAL NOTE". Water Resources Research. 42 (8). doi: 10.1029/2005WR004638 . S2CID   119938184.
  3. Wilhelm Alt, Hans; Luckhaus, Stephan (1 September 1983). "Quasilinear elliptic-parabolic differential equations". Mathematische Zeitschrift. 183 (3): 311–341. doi:10.1007/BF01176474. ISSN   1432-1823. S2CID   120607569.
  4. Feddes, R. A.; Zaradny, H. (1 May 1978). "Model for simulating soil-water content considering evapotranspiration — Comments". Journal of Hydrology. 37 (3): 393–397. Bibcode:1978JHyd...37..393F. doi:10.1016/0022-1694(78)90030-6. ISSN   0022-1694.
  5. Knight, John; Raats, Peter. "The contributions of Lewis Fry Richardson to drainage theory, soil physics, and the soil-plant-atmosphere continuum" (PDF). EGU General Assembly 2016.
  6. Richardson, Lewis Fry (1922). Weather prediction by numerical process. Cambridge, The University press. pp.  262.
  7. Celia, Michael A.; Bouloutas, Efthimios T.; Zarba, Rebecca L. (July 1990). "A general mass-conservative numerical solution for the unsaturated flow equation". Water Resources Research. 26 (7): 1483–1496. Bibcode:1990WRR....26.1483C. doi:10.1029/WR026i007p01483.
  8. Kuráž, Michal; Mayer, Petr; Lepš, Matěj; Trpkošová, Dagmar (2010-04-15). "An adaptive time discretization of the classical and the dual porosity model of Richards' equation". Journal of Computational and Applied Mathematics. Finite Element Methods in Engineering and Science (FEMTEC 2009). 233 (12): 3167–3177. Bibcode:2010JCoAM.233.3167K. doi:10.1016/j.cam.2009.11.056. ISSN   0377-0427.
  9. van Genuchten, M. Th. (September 1980). "A Closed-form Equation for Predicting the Hydraulic Conductivity of Unsaturated Soils". Soil Science Society of America Journal. 44 (5): 892–898. Bibcode:1980SSASJ..44..892V. doi:10.2136/sssaj1980.03615995004400050002x.
  10. Inoue, M.; Šimůnek, J.; Shiozawa, S.; Hopmans, J.W. (June 2000). "Simultaneous estimation of soil hydraulic and solute transport parameters from transient infiltration experiments". Advances in Water Resources. 23 (7): 677–688. Bibcode:2000AdWR...23..677I. doi:10.1016/S0309-1708(00)00011-7.
  11. Fodor, Nándor; Sándor, Renáta; Orfanus, Tomas; Lichner, Lubomir; Rajkai, Kálmán (October 2011). "Evaluation method dependency of measured saturated hydraulic conductivity". Geoderma. 165 (1): 60–68. Bibcode:2011Geode.165...60F. doi:10.1016/j.geoderma.2011.07.004.
  12. Angulo-Jaramillo, Rafael; Vandervaere, Jean-Pierre; Roulier, Stéphanie; Thony, Jean-Louis; Gaudet, Jean-Paul; Vauclin, Michel (May 2000). "Field measurement of soil surface hydraulic properties by disc and ring infiltrometers". Soil and Tillage Research. 55 (1–2): 1–29. doi:10.1016/S0167-1987(00)00098-2.
  13. Köhne, J. Maximilian; Mohanty, Binayak P.; Šimůnek, Jirka (January 2006). "Inverse Dual‐Permeability Modeling of Preferential Water Flow in a Soil Column and Implications for Field‐Scale Solute Transport". Vadose Zone Journal. 5 (1): 59–76. doi:10.2136/vzj2005.0008. ISSN   1539-1663. S2CID   781417.
  14. Younes, Anis; Mara, Thierry; Fahs, Marwan; Grunberger, Olivier; Ackerer, Philippe (3 May 2017). "Hydraulic and transport parameter assessment using column infiltration experiments". Hydrology and Earth System Sciences. 21 (5): 2263–2275. Bibcode:2017HESS...21.2263Y. doi: 10.5194/hess-21-2263-2017 . ISSN   1607-7938.
  15. Kuraz, Michal; Jačka, Lukáš; Ruth Blöcher, Johanna; Lepš, Matěj (1 November 2022). "Automated calibration methodology to avoid convergence issues during inverse identification of soil hydraulic properties". Advances in Engineering Software. 173: 103278. doi:10.1016/j.advengsoft.2022.103278. ISSN   0965-9978. S2CID   252508220.
  16. Farthing, Matthew W., and Fred L. Ogden, (2017). Numerical solution of Richards’ Equation: a review of advances and challenges. Soil Science Society of America Journal, 81(6), pp.1257-1269.
  17. Short, D., W.R. Dawes, and I. White, (1995). The practicability of using Richards' equation for general purpose soil-water dynamics models. Envir. Int'l. 21(5):723-730.
  18. 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.
  19. Germann, P. (2010), Comment on “Theory for source-responsive and free-surface film modeling of unsaturated flow”, Vadose Zone J. 9(4), 1000-1101.
  20. Gray, W. G., and S. Hassanizadeh (1991), Paradoxes and realities in unsaturated flow theory, Water Resour. Res., 27(8), 1847-1854.
  21. Downer, Charles W., and Fred L. Ogden (2003), Hydrol. Proc.,18, pp. 1-22. DOI:10.1002/hyp.1306.

See also